Next Article in Journal
Enteric Viruses in Turkeys: A Systematic Review and Comparative Data Analysis
Previous Article in Journal
Primary Sequence-Intrinsic Immune Evasion by Viral Proteins Guides CTL-Based Vaccine Strategies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Era of Gene Therapy: The Advancement of Lentiviral Vectors and Their Pseudotyping

by
Bat-Erdene Jargalsaikhan
1,2,*,
Masanaga Muto
1 and
Masatsugu Ema
1,3
1
Department of Stem Cells and Human Disease Models, Research Center for Animal Life Science, Shiga University of Medical Science, Seta, Tsukinowa-cho, Otsu 520-2192, Japan
2
Medical Genome Center, National Cerebral and Cardiovascular Center, 6-1 Kishibe-Shinmachi, Suita 564-8565, Japan
3
Institute for the Advanced Study of Human Biology (ASHBi), Kyoto University, Yoshida-Konoe-cho, Sakyo-ku, Kyoto 606-8501, Japan
*
Author to whom correspondence should be addressed.
Viruses 2025, 17(8), 1036; https://doi.org/10.3390/v17081036
Submission received: 5 June 2025 / Revised: 22 July 2025 / Accepted: 23 July 2025 / Published: 24 July 2025
(This article belongs to the Section Human Virology and Viral Diseases)

Abstract

Over 35 years of history, the field of gene therapy has undergone much progress. The initial concept—the replacement of dysfunctional genes with correct ones—has advanced to the next stage and reached the level of precise genome editing. Dozens of gene therapy products based on viral and non-viral delivery platforms have been approved, marking the dawn of the gene therapy era. These viral vector strategies rely on adenoviruses, adeno-associated viruses, lentivirus-derived tools, and so on. From the middle of the gene therapy transition, despite the challenges and serious negative consequences, the lentiviral vector has emerged as a cornerstone and demonstrated benefits in fields ranging from basic science to gene therapy. Therefore, we outline the importance of lentiviral vectors in the gene therapy era by focusing on their roles in the clinical usage, derivation, and development of next-generation platforms, as well as their pseudotyping.

1. The Derivation of Lentiviral Vectors

Acquired immunodeficiency syndrome (AIDS) was first recognized as a new disease in 1981 [1]. Luc Montagnier and his colleagues soon discovered that the cause of AIDS is a novel virus in the Retroviridae family, which was officially named human immunodeficiency virus (HIV) in 1986 [2]. HIV-1 and HIV-2, the main types of HIV, are positive-sense single-stranded RNA (ssRNA) viruses that contain reverse transcriptase (RT). The HIV-1 genome is approximately 9.7 kb and contains several cis-acting elements and 9 open reading frames that allow it to produce three main polyprotein precursors (Gag, Gag-Pol, and envelope), two regulatory proteins (Tat and Rev), and four accessory (Vif, Vpr, Vpu, and Nef) proteins [3,4] (Figure 1A).
An MLV-derived gamma-retroviral vector from the Retroviridae family was first developed in the early 1980s, based on transient transfection of separated plasmid vectors into producer cells [5,6], which showed the ability to transduce dividing cells and integrate transgenes into the host genome [7]. It can be pseudotyped using heterologous envelope proteins, such as vesicular stomatitis virus glycoprotein G (VSV-G), resulting in a high titer and broad tropism [8]. The gamma-retroviral vector was first introduced into gene therapy trials in the 1990s [9,10].
The first-generation LV system was developed in 1996 by Naldini and colleagues using the same principles as those used for the gamma-retroviral vectors as above, and is capable of transducing both dividing and non-dividing cells [11], offering exciting new potential for inserting genes into the genome of non-proliferating cells. The system contains three plasmid vectors, as illustrated in Figure 1B: (I) the packaging plasmid encodes all of the proteins excluding the HIV envelope protein and Vpu accessory protein, with the HIV-1 packaging signal and cis-acting elements deleted from the untranslated region; (II) the envelope plasmid encodes VSV-G protein or heterologous envelope protein; and (III) the transfer plasmid carries the transgene expression cassette under an internal promoter flanked by the HIV-1 long terminal repeat (LTR), containing the HIV-1 packaging signal and required cis-elements [11].
The second-generation LV system was subsequently developed in 1997 by Zufferey and colleagues [12], in which sequences encoding the accessory proteins Vif, Vpr, Vpu, and Nef were eliminated (Figure 1C). Although these accessory proteins confer a survival advantage with respect to lentivirus replication in vivo, they are not vital for viral growth in vitro [12,13].
Just a year after that, third-generation LV systems were reported in 1998 about the same time by Miyoshi [14] and Dull [15] and their respective colleagues. In this generation, (I) regulatory and accessory proteins were completely eliminated from the packaging plasmid vector; (II) the envelope plasmid encodes the VSV-G protein or heterologous envelope protein; (III) the transfer vector contains the CMV promoter-driven truncated 5′-LTR, required cis-acting elements, transgene expression cassette, partially deleted self-inactivating ΔU3-LTR (SIN-LTR) and polyadenylation (pA) sequences; and (IV) a separated regulatory plasmid vector which expresses only the Rev protein. A 133–400 bp deletion in the promoter and enhancer region of the 3′-LTR enables self-inactivation by eliminating LTR-driven gene expression in the integrated provirus [14,15]. During reverse transcription, the ΔU3-LTR is transferred to the 5′-LTR, allowing the proviral DNA to be flanked by the ΔU3-LTR [14,15,16]. Moreover, the U3 region of the 5′-LTR was replaced with a strong heterologous constitutive promoter (RSV or CMV) in the transfer plasmid vector, providing Tat-independent vRNA transcription during production [14,15,16]. The SIN-LTR reduces the risks of vector mobilization, replication-competent vectors, and unintended activation of nearby genes at the integration site [17,18,19]. The third-generation system eventually became a versatile tool [20], following the subsequent improvement of the transfer vector construct: first, through inclusion of the post-transcriptional regulatory element of woodchuck hepatitis virus (WPRE) before the self-inactivating LTR (SIN-LTR) [21]; and second, through insertion of the central polypurine tract/central termination sequence (cPPT/CTS) between the Rev-responsive element (RRE) and an internal promoter sequence [22] (Figure 1D). The WPRE significantly improves gene expression by enhancing mRNA processing and export in both intronless and spliced mRNAs, regardless of the cell cycle state of the transduced cell [21]. Nuclear localization of the DNA flap after reverse transcription of the viral genome is dependent on the cPPT/CTS, allowing the lentiviral vector to efficiently infect the host cells—especially non-dividing cells [22,23]—when compared with the MLV-derived gamma-retroviral vector that relies on mitosis [24]. Another attractive and useful aspect is integrase-defective lentiviral vectors (IDLVs). Several integrase-defective mutants at functional domains of IN, such as class I (e.g., D64V, D116N, E152A) and class II (e.g., D167K, D167A, Q168A, 262RRK), have been reported and investigated, reducing proviral integration [25,26,27,28]. Mutations on IN that primarily diminish the integration step are categorized as class I, while class II mutants mainly hamper particle assembly, reverse transcription, and nuclear import of the pre-integration complex (PIC) [25,26,27,28]. The D64V variant had reduced integration activity by up to 1/10,000 compared to the wild-type virus [25], unintegrated vectors maintain transient gene expression from episomal DNA up to several months, and the D64V is most common variant regarding the usage of IDLVs [20,26,27,28]. Finally, the third-generation lentiviral vector system, with a capacity of up to 10 kb cargo, is now considered a cornerstone in gene and protein delivery technology, having been utilized from basic science to gene therapy [20].

2. Clinical Use of Lentiviral Vectors

The first Phase 1 lentiviral vector clinical trial was conducted in 2003, which assessed the safety of autologous CD4+ T cells that were ex vivo transduced using an HIV-1-derived lentiviral vector encoding a 937-base antisense sequence against the HIV envelope. The trial involved five patients with HIV-1 who had previously experienced treatment failures with antiretroviral therapy. Among the participants, three patients demonstrated significant reductions in viral load, along with stable or increased CD4+ T cell counts [29,30]. Additionally, for all 65 subjects in Phases 1 and 2, no adverse events related to the therapy were recorded [31]. Over the past two decades, lentiviral vectors have emerged as powerful tool and there have been remarkable advances in the clinical application of such vectors, highlighting their enormous potential for gene therapy aimed at a wide range of inherited and acquired diseases.
We comprehensively summarized 701 registered clinical trials of the lentiviral vector-based gene therapies shown in Supplementary Table S1. Of those, 209 clinical trials involved lentiviral vector-mediated gene transfer therapies, with lentivirus-like particles used to deliver CRISPR/Cas in ocular diseases in 4 clinical trial registries, 44 trials for T cell receptor (TCR) cell therapy, and the rest of the registries involving chimeric antigen receptor (CAR) cell therapy. Overall, the clinical applications of lentiviral vectors can be divided into five main uses: (I) correction of faulty or mutated genes by delivering a functional copy to ex vivo HSCs, (II) delivery of CAR or TCR to the ex vivo T cells, (III) in vivo direct gene transfer, (IV) antisense or shRNA delivery, and (V) delivery of gene editing tools, indicating broad-range utilization of lentiviral vectors in the field. The major human diseases in clinical trials include neurological, ocular, hematological, metabolic, cancer, immunological, skin, viral infection, and lung diseases, as well as long-term follow-up for viral vector gene therapies (Figure 2).
Gene therapy—especially the cutting-edge approaches of CAR-T and TCR-T cell therapy—has recently achieved remarkable success in treating cancers that have failed to respond to conventional treatments [32,33,34]. The concept is that genetic modification of T cells can increase their ability to specifically target cancer cells [35,36]. CAR-T or TCR-T cell therapy involves collecting T cells from patients or donors, selecting a suitable subclass (e.g., Treg, CD3, CD4, CD8, CD4/CD8), and performing ex vivo genetic modification by transducing them with a lentiviral vector carrying CAR gene or TCR gene expression cassette, followed by expansion and quality control of the transduced cells, as well as cryopreservation of the CAR-T or TCR-T cells. The cryopreserved cells can be stored for up to several months, ready for infusion into the patients to fight against cancer [37,38]. CAR-T directly targets cancer cells via CAR, which consists of a cancer antigen-specific external single-chain variable fragment (scFv), a transmembrane domain, an intracellular signaling domain, and a costimulatory domain from receptors (e.g., CD28, OX40, CD137), while TCR-T recognizes intracellular specific antigens presented by major histocompatibility complex (MHC) molecules of cancer cells through the transduced natural TCR (α- and β-chains non-covalently linked to the CD3 complex on the surface), which leads to the activation of T cells and destruction of cancer cells [39,40].
TCR-T cell therapy is considered valuable against solid tumors. For example, it has been shown to be effective in melanoma by targeting the New York esophageal squamous cell carcinoma (NY-ESO-1) antigen [41]; effective for synovial sarcoma, ovarian cancer, and head and neck cancer treatment by recognizing melanoma antigen gene A4 (MAGE-A4) antigen [42,43]; and promising against hepatocellular carcinoma (HCC) via identification of alpha-fetoprotein (AFP) antigen [44]. Moreover, mesothelin (MESO)-targeting TCR-T cell therapy is encouraging and is under clinical trials for various cancers including mesothelioma, lung cancer, ovarian cancer, and other solid tumors [45].
CAR-T cell therapy has demonstrated outstanding performance in treating hematologic malignancies, with approximately 80% malignancy response of complete or partial remission [46]. Several CAR-T gene therapy products are available that recognize and destroy CD19-positive B cell malignancies, which are notable for treating specific B-cell non-Hodgkin lymphomas, B-cell ALL, and CLL [47,48,49,50,51,52,53]. Furthermore, plasma cell-associated protein B-cell maturation antigen (BCMA)-specific CAR-T cells are effective in the treatment of multiple myeloma [54,55,56,57,58]. In clinical trials, CAR-T cell therapies targeting various tumor-associated antigens (TAAs), such as CD19/CD20, C-type lectin-like molecule-1 (CLL-1), CD22, CD133, and CD171, have been assessed [59,60,61,62,63,64].
Lentiviral vector-mediated gene transfer into CD34+ HSCs has been utilized in the treatment of several genetic disorders, including X-linked severe combined immunodeficiency (X-SCID) caused by mutations in the IL2RG gene [65]; sickle cell disease [66] and β-thalassemia [67], both resulting from mutations in the HBB gene; Fanconi anemia caused by mutations in the FANCA gene [68]; metachromatic leukodystrophy (MLD) due to mutations in the ARSA gene [69]; cerebral adrenoleukodystrophy (CALD) associated with mutations in the ABCD1 gene [70]; hemophilia A arising from mutations in the F8 gene [71]; and Wiskott–Aldrich Syndrome (WAS), which is caused by mutations in the WAS gene [72].
Moreover, lentiviral vector-based gene therapy products directly injected into the target tissue include tyrosine hydroxylase (TH), aromatic L-amino acid decarboxylase (AADC), and GTP-cyclohydrolase 1 (CH1) gene delivery in brain striatum for Parkinson’s disease [73]; endostatin and angiostatin gene transfer via subretinal injection to treat neovascular age-related macular degeneration (AMD) [74]; and ABCA4 gene delivery to the eyes of Stargardt Disease patients [75], all resulting in well-tolerated and sustained transgene expression.
Despite these successful stories, lentiviral vector-mediated gene therapy presents some adverse effects, such as cytokine release syndrome, neurotoxicity, severe infection, and prolonged cytopenia in the short-term as well as a theoretical risk of insertional secondary malignancy in the long-term [76]. Besides, in vivo CRISPR/Cas-mediated gene editing therapy trials have utilized lentivirus-like particles to treat ocular diseases, including primary open-angle glaucoma with MYOC gene mutations and refractory herpetic viral keratitis [77], expanding their potential.

3. The Gene Therapy Era and Lentiviral Vectors

The United States (US) Food and Drug Administration (FDA) broadly defines human gene therapy as a therapeutic technique that modifies or manipulates the expression of a gene or alters the biological properties of living cells. The guidelines also include the use of gene therapy products to achieve a therapeutic or preventive effect via the transcription or translation of transferred genetic material or altering human genetic sequences [78]. According to the regulatory document of the European Medicines Agency (EMA), gene therapy medicinal products contain recombinant nucleic acids or a substance used to regulate, repair, replace, add, or delete a genetic sequence for therapeutic, preventive, or diagnostic purposes. Its effect is directly related to the genetic sequence or its expression product [79]. Both the FDA and EMA guidelines widely consider recombinant nucleic acids (e.g., plasmids DNA, RNA), genetically modified microorganisms (e.g., viruses, bacteria, fungi), engineered site-specific nucleases, and ex vivo genetically modified human cells to be human gene therapy products, except for vaccines against infectious diseases [78,79]. At the end of 2024, 55 gene therapy products had been approved [47,48,49,50,51,52,53,55,56,57,58,66,67,69,70,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120], excluding non-genetically engineered cell therapy and vector vaccines. These have been categorized into three developmental stages, as shown in Figure 3.
In 1989, Rosenberg and his colleagues demonstrated the feasibility of using autologous human cells transduced with a gamma-retroviral vector derived from murine leukemia virus (MLV) for gene therapy [10], which marked the beginning of the gene therapy era. The original concept of gene therapy was to treat hereditary disorders caused by faulty genes by delivering a functional copy of the gene to the affected cells using non-viral and viral vectors [121,122,123]. Therefore, the first therapeutic clinical trial involving the autologous infusion of T cells with the gamma-retroviral vector-mediated adenosine deaminase (ADA) gene transfer was conducted in two children with Severe Combined Immunodeficiency due to adenosine deaminase deficiency (ADA-SCID) monogenic disease in 1990 [9,124]. Gene therapy holds great promise and, since then, hundreds of gene therapy clinical trials have been conducted worldwide using potential viral and non-viral gene delivery methods, laying the foundation for further development. For instance, several clinical trials were conducted to introduce a functional copy of the cystic fibrosis membrane conductance regulator (CFTR) gene into the airway cells of patients with cystic fibrosis caused by CFTR mutations in the early 1990s [125,126,127,128,129]. These trials used different tools for therapeutic purposes. In the UK and US, researchers have used cationic liposomes combined with complementary DNA [125,126]. In the US and France, a human Adenovirus serotype 5 (Ad5) viral vector has been used [127,128]. In the US, a human Adeno-associated virus serotype 2 (AAV2) viral vector [129] was also used to transfer a functional copy of the CFTR gene into the airway epithelium. However, these attempts did not provide sustained lung function for cystic fibrosis patients due to ineffective gene transfer and the inhibition of AAV2, cationic lipofection, and Ad5 via bronchial secretion and local immune responses [130,131,132]. Unfortunately, in 1999, the first notorious adverse effect of gene therapy was the death of an 18-year-old male with partial ornithine transcarbamylase (OTC) deficiency, due to a severe immune reaction after an intra-arterial infusion of human adenovirus type 5, deleted in E1 and E4, containing human OTC cDNA [133]. Later, at the end of 2002, patients treated for X-linked SCID (X-SCID) using the ex vivo genetically modified autologous CD34+ hematopoietic stem cells (HSCs) with MLV-derived gamma-retroviral vector-based transfer of IL2RG expression cassette started to develop lymphoid T-cell leukemia [134]. This gene therapy-related secondary malignancy was caused by insertional activation of the proto-oncogenes LMO-2, BMI1, and CCND2 in almost half of the patients [135,136]. These initial clinical trials revealed serious treatment-related toxicities, including inflammatory responses to the vectors, as well as secondary malignancies resulting from the viral vector-mediated activation of proto-oncogenes; therefore, similar trials were halted by the governments of France and the US [137]. These early attempts—with all the setbacks and lessons learned—can be considered the first stage of human gene therapy development [138].
The second stage of gene therapy was initiated in early 2000s, stimulated by more basic research in virology, immunology, cell biology, model development, and targeted diseases, which led to the development of improved gene therapy products and their successful application in gene therapy [139]. While the UK Gene Therapy Advisory Committee was inviting feedback on the use of the MLV-derived gamma-retroviral vectors—particularly regarding the potential improvements associated with self-inactivating (SIN) vectors as an alternative to gamma-retroviral vectors [137]—the first clinical trial using a human immunodeficiency virus 1 (HIV-1)-derived lentiviral vector was started in January of 2003 in the US, offering a safer and more efficient viral vector [140]. The MLV-derived gamma-retroviral vector integration sites are typically located in the transcription region, while the HIV-1-derived lentiviral vector insertions were significantly reduced in that region and evenly distributed throughout the rest of the gene region [141,142]. Additionally, lentiviral vectors can transduce both dividing and non-dividing cells [143,144], suggesting a safer and more suitable choice than gamma-retroviral vectors. From 2010, the use of the lentiviral vectors in gene therapy clinical trials has increased dramatically, surpassing gamma-retroviral vectors [145]. Furthermore, due to the improved safety achieved by removing enhancer-promoter elements from the LTR, MLV-derived SIN gamma-retroviral vectors for gene therapy are used to this day [48,49,92,93,96] and have been formally approved (Figure 3). Moreover, SIN lentiviral vectors (both integrative and non-integrative) have been used in clinical trials for various therapeutic applications—particularly ex vivo transduction—including in hereditary genetic disease by introducing functional genes to HSCs; in cancers by delivering chimeric antigen receptor (CAR) and tumor specific T-cell receptor (TCR) to the T cells; and in RNA and protein delivery to target cells [20]. Subsequently, the first regulatory approval of a lentiviral vector gene therapy product occurred in 2017; namely, HIV-1-derived lentiviral vector-mediated chimeric antigen receptor (CAR)-T cells targeting CD19 to treat B cell malignancies [146]. In the treatment of X-SCID via ex vivo genetically modified CD34+ HSCs, the usage of SIN lentiviral vector demonstrates benefits over the gamma-retroviral vector, such as providing multilineage engraftment and effective immune reconstitution [65]. Moreover, three decades after the earlier venture to treat cystic fibrosis via gene therapy, the first human trial involving CFTR-targeting therapy with a lentiviral vector began in 2024, utilizing a Simian immunodeficiency virus (SIV)-derived lentiviral vector pseudotyped with Sendai virus (SeV) fusion (F) and hemagglutinin-neuraminidase (HN) envelope proteins [147]. The SeV-F/HN pseudotyped SIV-derived lentiviral vector is ideal for introducing transgenes into airway cells due to its high affinity for α2,3 sialylated N-acetyllactosamine (LacNAc) [148,149].
In the case of AAV vectors, the discovery of natural serotypes beyond AAV5 (such as AAV8, 9, and rh74), the development of engineered capsids that evade immune responses, and the design of vectors for prolonged transgene expression and improved safety have led to the next generation, with AAV gene therapy products having received official approval since 2012 [150]. Furthermore, the earliest antisense oligonucleotides (ASOs) that entered clinical trials consisted of a phosphodiester or phosphorothioate backbone without any modifications [151]. They were transformed with modifications including 2′-O-Methyl (2′-O-Me), 2′-O-Methoxyethyl (2′-O-MOE), and 2′-Fluoro (2′-F), as well as chemical modification of the furanose ring with next-generation oligonucleotides (ASO and siRNA), consequently demonstrating clinical efficacy and safety and also gaining approval [152,153]. ASOs are single-stranded synthetic oligonucleotides, typically 15-30 bases long, designed to bind the target RNA (pre-mRNA or mRNA) to regulate gene expression through the mechanism of either RNase H-mediated RNA degradation or altering its processing [81,153,154]. In contrast, siRNAs—which are typically 20-25 bp long double-stranded RNA molecules (modified or non-modified)—lose one strand (the passenger strand) after Dicer cleavage in the cytoplasm, while the remaining strand (the guide strand) leads to target mRNA degradation by forming an RISC (RNA-induced silencing complex) [81,152,154]. Further, potential viral and non-viral methods for gene transfer are still being explored, such as a novel Anellovector that is under development from the human native Anelloviridae viruses, which do not cause any known diseases [155]. The majority of gene therapy products have advanced to the next generation at this stage and have reached official approval, offering safer and more effective approaches for the delivery of genetic material (DNA or RNA) or proteins into human cells for therapeutic purposes.
From the 2010s, the potential for the precise replacement or editing of mutated genes with exact corrections appeared with the achievement and development of admirable gene editing tools based on engineered site-specific nucleases, such as Zinc-finger nuclease (ZNF) [156], Transcription Activator-Like Effector Nuclease (TALEN) [157,158], and Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR)/CRISPR-associated protein (Cas) (CRISPR/Cas) [159]. Moreover, site-specific nuclease-based next-generational base editing [160] and prime editing [161] provide tools to edit DNA without double-strand break, which evolved promptly and play a crucial role in the field [162,163]. After only a short period, the CRISPR/Cas genetically modified cell therapy astonishingly received official approval in 2023 [164]. Additionally, gene therapy needs to be highly effective, with minimal toxicity and low cost [46]. Delivering the CRISPR/Cas system to target cells is challenging, and one of the most important delivery tools in both in vitro and in vivo applications is lentiviral vectors [165,166,167]. Essentially, lentiviral vectors can be used to deliver CRISPR/Cas to cells in two main ways. First, integrative and non-integrative lentiviral vectors that simply carry CRISPR/Cas and sgRNA expression cassettes have become an important tool in biomedical research [167,168]. Another method is to utilize lentivirus-like particles that are packaged with engineered site-specific nucleases or cargo proteins [166,169,170,171]. Current in vivo gene editing therapy trials employ lentivirus-like particles to deliver the CRISPR/Cas system [77], suggesting the continued potential of recombinant lentiviral vectors.
Gene therapy is believed to be the medicine of the 21st century [172] and tremendous progress has been made in recent years, with 6,333 clinical trials including a range of promising treatments for a variety of diseases registered by the end of 2024 [173]. Additionally, 21 non-viral and 34 viral gene therapy products have been approved (Figure 3); for a detailed overview, see Supplementary Table S2. Notably, 14 are lentiviral vector gene therapies, highlighting their significance within the area. Lentiviral vectors have emerged as precious tools, having entered clinical application from the second-stage of the gene therapy era and demonstrating benefits.

4. Current Issues and Development of Next-Generation Lentiviral Vectors

The third-generation lentiviral vector system is still considered state-of-the-art, even two decades after it obtained its current form and began to be utilized in gene therapy [20]. Although CAR-T has achieved remarkable triumphs against blood cancer [54], there remain concerns due to the possibility of insertional oncogenesis [174]. Furthermore, the US FDA has investigated CAR-T therapy-related secondary malignancies and revealed 33 cases among the 30,000 people who had been treated with it, as of the first quarter of 2024 [175]; however, there is a debate regarding whether the danger is due to the original blood cancer or the cutting-edge therapy. However, the benefits outweighed the risks and saved thousands of lives [176]. In particular, deleting the enhancer and promoter sequences from the LTR, SIN gamma-retroviral, and lentiviral vectors minimized the risk of unintended activation of nearby promoters at the integration sites [17,18,19] (Figure 4A,B).
In hundreds of clinical trials over two decades using lentiviral vectors for gene therapy, no serious complications had been reported until recently. Unfortunately, 7 of 67 patients who suffered CALD—a rare genetic disorder that damages the brain and spinal cord caused by ABCD1 gene mutation—developed hematological malignancies after infusion of lentiviral vector-mediated ex vivo genetically modified autologous HSCs [177]. The issue is likely due to activation of the proto-oncogenes near the integration sites through the Myeloproliferative sarcoma virus enhancer, Negative control region deleted (MNDU3) internal promoter in the SIN lentiviral vector that drives ABCD1 cDNA expression [178], as shown in Figure 4C. The lentiviral vectors—which used human cellular gene promoters in at least 16 other diseases—have shown a strong safety record over the past 15 years [178], suggesting that vector design is vital. Additionally, the MND-LTR promoter had stronger and more persistent transgene expression in hematopoietic cells of all lineages compared to the MLV-LTR promoter [179]. Usage of the LTR promoter-enhancer derived from gamma retroviruses as an internal promoter in the SIN lentiviral vector, as well as inclusion of the gamma-retroviral LTR promoter-enhancer at the LTR sequence of the lentiviral vector, can lead to increased activation of nearby genes, potentially contributing to tumorigenic processes [180,181].
Simultaneously, several next-generation lentiviral vector platforms are being developed and/or are ready for deployment. The first next-generation lentiviral vector, named Lenti-X fourth-generation, which has been commercially available since around the end of the 2010s, is a product of Takara Bio Inc (Japan) which has been widely used in basic research [182,183,184,185]. Transfer and envelope plasmid vectors remain intact in the Lenti-X and, in order to produce ultra-high titers of lentiviral vectors, a complicated packaging system was introduced compared with that in the previous generation (Figure 5A,B).
The new system contains Tetracycline-controlled transactivator (tTA), which binds to the tetracycline response element (TRE) in the promoter region, leading to a high level of gene expression in the absence of tetracycline/doxycycline [186]; the tTA is shown as Tet-off under the CMV promoter in Figure 5B. The packaging construct gag-pol was divided into two separated plasmids—(I) the TRE promoter-driven gag-pro expression cassette, and (II) vpr accessory and pol under LTR from HIV-2 (LTR-2)—ensuring safety and reducing the risk of replication-competent vectors. When the gag-pro-pol sequence was divided into two separate components, gag-pro and pol, the frequency of recombination-dependent DNA mobilization was remarkably reduced when compared with the intact plasmid vector [187]. Moreover, high-titer lentiviral vector production is achieved via intense expression of viral essential proteins through the TRE (can be disabled with Tet-off) and LTR-2 (driven by Tat-transactivator) promoters.
Vink and his colleagues developed the LTR1 fourth-generation lentiviral vector system by rearranging the cis-acting elements in the transfer plasmid vector (Figure 5C) that controls the complex replication process of the lentiviral genome in the host, and demonstrated that this next-generation system could produce sufficient titers for preclinical gene therapy in a hemophilia B mouse model [188,189]. The packaging, envelope, and regulatory plasmid vectors were the same as in the previous version. Although SIN vectors were thought to have no LTR-driven transcription due to deletion of the promoter/enhancer elements, full-length packageable vRNA was exhibited from proviral DNA, indicating some risk [190]. While traditional SIN vectors have both a 5′ and 3′ LTR flanking the viral genome, the LTR1 fourth-generation vector features PBS at the 5′ end and a single complete SIN-LTR followed by PBS, the packaging signal (ψ), and RRE at the 3′ end. As a result of this modification, the LTR1 requires only a single DNA transfer jump (unlike the classical two-step DNA transfer jump), enabling removal of the packaging signal (ψ) and RRE during reverse transcription (Figure 6A,B). Hence, even if full-length vRNA leakage occurs in LTR1 provirus, the elimination of the packaging signal (ψ) significantly reduces vector mobilization.
Another novel platform, the TetraVecta fourth-generation lentiviral vector system, was launched by Oxford Biomedica (UK) in 2023 and is characterized by enhanced safety, quality, and large payload capacity by introducing four systematic modifications on the transfer plasmid [191,192], as illustrated in Figure 5D. Firstly, 2KO modification at the major splicing donor (SD) site, optimized SD-inactivating sequences, along with a new class of vRNA enhancers based on modified U1 snRNA [192], provides unspliced viral genomic RNA (vgRNA) during production, as HIV-1 splicing at the SD site is dependent on U1 snRNA [193,194]. Moreover, approximately 95% of the vgRNA undergoes mRNA splicing during production [195]. Secondly, viral sequences with the RRE element are removed, allowing approximately 1 kb of additional space for the payload. The system is Rev-independent and does not need a regulatory plasmid vector. Thirdly, a bacterial tryptophan RNA-binding attenuation protein (TRAP) sequence is introduced near the transcriptional starting site of the transgene—called the transgene repression in vector production (TRiP) system—which inhibits translation during viral vector production [196,197] and is beneficial for CAR-T cell production or cytotoxic payloads. The last refinement is bidirectional poly A sequences (supA) at the SIN-LTR, which are 50-fold stronger than the previously used sequences, thus improving transcriptional insulation [191,192].
It is generally known that third-generation lentiviral vector systems adopt roughly 20% of the HIV-1 genome sequence, which integrates into the host genome along with the transgene. Next-generation systems substantially decrease the HIV-1 genome sequence in the integrated provirus, down to approximately 5% for LTR1 and 10% for TetraVecta platforms. As current issues, it has been suggested that vector design and the choice of internal promoter in the SIN vector are critical; however, the fourth-generation systems have unique features and aim for enhanced safety, quality, production efficiency, and payload capacity, as well as minimized viral backbone and toxicity.

5. Structure of the Lentiviral Particle and Its Pseudotyping

The mature HIV-1 virion is a spherical particle, approximately 100 nm, surrounded by a lipid membrane with glycoproteins containing a conical capsid, viral genomic ssRNA, and proteins [198]. As the lentiviral vector is derived from HIV-1, it resembles the virion in structure (Figure 7A); however, the incorporation of regulatory and accessory proteins depends on the method of lentiviral vector generation.
The protease (PR) is incorporated into particles as part of the monomer Gag-Pol polyprotein and then dimerizes to become active during viral budding, leading to the maturation of viral particles [199]. The capsid is a protein shell known as p24 that surrounds ribonucleoprotein (RNP) particles and includes two copies of viral genomic ssRNA, ncRNAs (e.g., 7SL RNA [200,201], tRNA lys [202,203]), viral enzymes (IN and RT), and viral nucleocapsid proteins. Cyclophilin A (CypA) is a host cell protein that is incorporated and binds to the HIV-1 capsid [204]. Moreover, cytoskeletal proteins from the host cell—such as actin, myosin, and ezrin—are incorporated into the particle [205]. Although regulatory and accessory proteins could be incorporated, Vpr is a predominant component in the viral particle that is critical for viral replication in target cells [206]. Furthermore, depending on the host cell type, different types of proteins from the host are incorporated into the viral particles to varying degrees [207]. The matrix (MA) protein is inside the lipid membrane, which is crucial for viral replication, maturation, and the structure of the particle [208]. The outer layer of the viral particle is a lipid membrane with envelope proteins. Interestingly, the HIV-1 gp120-gp41 envelope protein is incorporated in the envelope membrane along with cellular molecules such as MHC class I and II, CD molecules (e.g., CD54, CD63, CD11), and integrins (e.g., integrin α4β7) [207,209,210]. The host cell molecules assembled in the envelope play a role in evading the immune response and facilitating cell-to-cell transmission [207,209,210]. The HIV-1 virion enters cells using the envelope proteins in two main ways: (I) direct fusion with the cell membrane [3,4,211,212] or (II) endocytosis and fusion with intracellular compartments [212,213]. Then, the lentiviral genome integrates into the genome of the target cell through a sequential process that reverses transcription via RT and tRNA lys, entering into the nucleus through the nuclear pore complex (NCP), using viral IN and cellular LEDGF/p75 as co-factors [203,214].
It is worth noting that the gp120-gp41 envelope protein can be replaced with heterologous glycoproteins—a common technique recognized as pseudotyping, which is analogous to adapting one’s attire to different contexts [215,216]. The interaction between the envelope protein and the target receptor on the cell membrane determines viral entry, while pseudotyping alters tropism [215,216]. The viral envelope glycoproteins of enveloped viruses can be incorporated into lentiviral vectors through several strategies, including the use of a wild-type, a truncated form, modified cytoplasmic tail, or an engineered envelope protein [148,217,218,219]. Single-envelope protein pseudotyping is the most familiar scenario, such as VSV-G, MLV-A, MLV-E, and RD114 [219]. Additionally, dual or multiple pseudotyping is possible, through which two or more envelope proteins can be incorporated into a single lentiviral particle. Phenotypically unmixed pseudotyping of lentiviral particles is characterized by the incorporation of viral envelope proteins from a single origin, whereas phenotypically mixed pseudotyped lentiviral particles incorporate multiple (usually dual or triple) envelope proteins from distinct origins on the same viral particle surface [217,218,220,221,222]. Table 1 lists representative examples of pseudotypes exhibiting phenotypically unmixed and mixed characteristics. The heterologous envelope proteins are well adopted in lentiviral vectors for phenotypically unmixed pseudotyping, while a compatible combination is required for phenotypically mixed pseudotyping [218,223]. Moreover, we illustrate currently applied pseudotyping apporaches for HIV-1-derived lentiviral vectors in Figure 7B.
Regarding envelope proteins from Retroviridae family viruses, HIV-1 envelope protein gp120-gp41 binds the CD4 receptor and co-receptors (CCR5 and CXCR4), infecting CD4+ cells, allowing for HIV-1 investigation or gene delivery to target cells [211,225]. Murine leukemia virus amphotropic (MLV-A) envelope proteins (4070A, which recognizes Pit2 and 10A1 binds both Pit1 and Pit2) and murine leukemia virus ecotropic (MLV-E) envelope protein bind CAT1, thus expanding tropism in murine and human cells [226,227,228,229]. Gibbon Ape Leukemia Virus envelope protein (GALV) targets GLVR-1, and its pseudotyping efficiently transduces human T and B cells [230]. RD114 is a feline endogenous retrovirus-derived envelope protein that targets the ASCT2 receptor and confers high infectivity in CD34+ cells [231]. Avian sarcoma leukosis virus (ASLV) envelope proteins allow for the effective transduction of lentiviral vectors to neurons via recognizing TVA and TVB receptors [232]. Jaagsiekte sheep retrovirus (JSRV) envelope protein can bind glycosylphosphatidylinositol (GPI)-anchored Hyaluronidase 2 (Hyal2), and its pseudotyping is suitable for lung epithelial cell transduction [233]. Modified foamy viral envelope protein has shown benefits for gene transfer efficiency in CD34+ cells [234]. While Baboon endogenous virus (BaEV)-envelope protein pseudotyped lentiviral vector was shown to be effective in cytokine-stimulated NK cells [235], Koala retrovirus (KoRV) envelope protein pseudotyping was predominant in freshly isolated immune cells [236]. In addition, the R peptide-truncated BaEV (BaEVRless) envelope protein pseudotyped lentiviral vector retained infectivity in human serum, and had a superior ability to transduce cytokine-unstimulated CD34+ HSCs cells by targeting both ASCT1 and ASCT2 receptors in the presence of RetroNectin (a transduction enhancer), offering potential for ex vivo gene delivery targeting CD34+ HSCs [237,238,239].
Regarding envelope proteins from Rhabdoviridae family viruses, the VSV-G envelope glycoprotein targets low-density lipoprotein receptor (LDL-R) and is the most popular pseudotyping protein, having been used in both research and clinical applications due to its broad tropism, efficient packaging, stability, and robustness [11,219,224]; furthermore, it is inactivated by human serum [240]. Moreover, the envelope glycoprotein G of viruses in this family—such as Rabies virus (RABV), Chandipura vesiculo virus (CNV), Piry vesiculo virus (PRV), Maraba virus (MARV), Cocal virus (COCV), and Mokola virus (MOKV)—can be used in lentiviral pseudotyping, providing neuronal target delivery, reduced immunogenicity, and increased specificity [229,241,242,243,244].
Regarding the envelope proteins from Baculoviridae family viruses, the Baculovirus GP64 envelope glycoprotein pseudotyping for lentiviral vectors is an alternative option to replace standard VSV-G pseudotyping in general usage, due to its reduced immunity, lower toxicity, high stability, and broad tropism [245,246,247]. Heparan sulfate proteoglycan receptors and lipid rafts are recognized potential targets of the GP64 envelope protein [248].
Regarding envelope proteins from Paramyxovirinae subfamily viruses (Paramyxoviridae family), the modified hemagglutinin (H) and fusion (F) glycoprotein phenotypically unmixed dual-pseudotyped lentiviral vectors (H/F-LV) of the measles virus are considered a promising tool to introduce transgenes into immune cells and unstimulated CD34+ HSCs [249,250,251,252]. The phenotypically unmixed dual-pseudotyped lentiviral vectors with Sendai virus-modified HN and truncated F glycoproteins are considered ideal tools for delivering transgenes into airway cells, which are currently being employed in clinical trials [147,148,149]. Moreover, lentiviral vectors can be pseudotyped with HN and F of the human parainfluenza virus [253,254]. Furthermore, while some highly pathogenic Nipah and Hendra viruses are restricted to biosafety level-4 (BSL-4), the henipavirus F and G envelope glycoprotein pseudotyped lentiviral vector system provides a practical and low-risk technique for virology and biomedical research [255]. The modified F and wild-type G envelope proteins of the Nipah virus incorporate efficiently into the lentiviral vector and target ephrinB2+ cells, serving as a promising in vivo or in vitro delivery platform [256].
Regarding envelope proteins from Pneumovirinae subfamily viruses (Paramyxoviridae family), although human respiratory syncytial virus (RSV) is contagious, causing infection in the lungs and respiratory tract [257], a phenotypically unmixed heterologous triple-pseudotyped lentiviral vector (SH/G/F-LV) using the RSV envelope proteins SH, G, and F is considered useful in RSV virology research [220].
Regarding envelope proteins from Hepadnaviridae family viruses, a phenotypically unmixed heterologous triple-pseudotyped lentiviral vector (L/M/S-LV) using the L, M, and S envelope proteins of hepatitis B virus (HBV) has shown potential for liver-specific gene therapy and helped to elucidate HBV’s attachment and entry mechanisms [222].
Regarding envelope proteins from Orthomyxoviridae family viruses, a lentiviral vector pseudotyped with Hemagglutinin (HA) and neuraminidase (NA) envelope proteins from influenza viruses is considered as an alternative and attractive source for the study of influenza virology, in order to evaluate neutralizing antibodies and viral entry [258,259].
Regarding envelope proteins from Filoviridae family viruses, the severe hemorrhagic fever outbreak caused by Ebola and Marburg viruses—both BSL-4 pathogens—can lead to organ failure and death [260,261,262]. Therefore, the utilization of lentiviral vectors pseudotyped with the GP1 and GP2 envelope proteins from these threatening pathogens provides safer tools, allowing for better understanding of the underlying viral pathogenesis and efficient gene delivery [229,263,264,265].
Regarding envelope proteins from Coronaviridae family viruses, spike (S) and E envelope proteins from the SARS-CoV-1, MERS-CoV, or SARS-CoV-2 viruses pseudotyped lentiviral vectors have greatly contributed to the investigation of their viral entrance mechanisms, the evaluation of neutralizing antibodies, and the development of therapeutic approaches or vaccines [266,267].
Regarding envelope proteins from Togaviridae family viruses, this family comprises a range of zoonotic viruses—including chikungunya virus (CHIKV), sindbis virus (SINV), ross river virus (RRV), and semliki forest virus (SFV)—which cause a variety of symptoms, ranging from fever and headache to rash [268,269,270,271,272]. The generation and usage of the lentiviral vectors pseudotyped with E1 and E2 envelope proteins from these viruses has led to a better understanding of their basic virology and the immune responses associated with infection [273,274,275,276,277].
Regarding envelope proteins from Flaviviridae family viruses, the viruses in this family consist of important human pathogens, including zika virus (ZIKV), dengue virus (DENV), yellow fever virus (RFV), hepatitis C (HCV), and Japanese encephalitis virus (JEV) [268,278,279,280]. Pseudotyped lentiviral vectors using the envelope glycoproteins E1 and E2 (e.g., ZIKV, HCV, JEV) have shown benefits in investigating the attachment and entrance of the pathogens, the development of vaccines, and alteration of tropism [281,282,283].
Regarding viruses belonging to the Bunyavirales order, some cause zoonotic diseases; such as Rift Valley Fever Virus (RVFV) from the Phenuiviridae family, Crimean-Congo Hemorrhagic Fever Virus (CCHFV) from the Nairoviridae family, and Andes virus (ANDV) from the Hantaviridae family [284,285,286]. Lentiviral vectors pseudotyped with the Gn and Gc glycoproteins from these viruses have been utilized to study them and develop associated vaccines and treatments [287,288,289]. Moreover, several viruses from Arenaviridae family (Bunyavirales order)—including Lassa virus (LASV), Lujo virus (LUJV), Junin virus (JUNV), Machupo virus (MACV), Guanarito virus (GTOV), Tacaribe virus (TCRV), Chapare virus (CHAPV), and Sabiá virus—are responsible for causing zoonotic diseases and are classified as BSL-3 and BSL-4 pathogens [262,290,291,292]. Hence, lentiviral vectors pseudotyped with GP1 and GP2 envelope proteins from these zoonotic viruses are considered precious tools in the biomedical area [291,293]. In addition, the envelope proteins of non-pathogenic Arenaviridae family viruses, including Lymphocytic choriomeningitis virus (LCMV) and Pichinde virus (PICV), have also been used to pseudotype lentiviral vectors for virology research and tropism-altering purposes [229,294,295,296].
Regarding envelope proteins fused with scFV or nanobodies, the use of engineered envelope proteins for the lentiviral vectors (e.g., VSV-G glycoprotein fused with scFV or nanobodies, and Sindbis E1 and E2 glycoprotein fused with scFV or nanobodies) has demonstrated highly specific target delivery both in vivo and in vitro [297,298,299,300]. The VSV-G envelope protein fused with anti-CD30 scFV pseudotyping targets CD30+ lymphocytes [301], while the VSV-G fused with anti-EGFR scFV pseudotyping targets EGFR+ tumor cells [302]; furthermore, with poloxamer-based adjuvants, higher infectivity to the target cells than typical VSV-G pseudotyped LV was observed, offering potential in terms of cancer and immunotherapy. Moreover, binding-defective and fusion-competent VSV-G fused with antigen-presenting cell (APC)-specific nanobody pseudotyped LV was shown to selectively infect dendritic cells and macrophages both in vitro and in situ, demonstrating higher target delivery when compared with the VSV-G pseudotyped LV [298]. The CD90-targeting lentiviral vector, pseudotyped with either binding-defective and fusion-competent VSV-G with anti-CD90 scFV or binding-defective and fusion-competent measles H/F with anti-CD90 scFV, achieved success with lower off-target effects for ex vivo delivery in true HSCs, laying the foundation for in vivo gene therapy [303]. Thus, engineered envelope proteins—obtained via scFV and Nanobody display technology—provide the primary advantage of precise targeting both in vivo and ex vivo, which offers benefits including diminished off-target effects when compared to most envelope proteins with broad natural tropism, thus advancing the potential of lentiviral vectors.
In summary, pseudotyped lentiviral vectors with phenotypically unmixed or mixed heterologous envelope glycoproteins and engineered envelope proteins provide several advantages, such as expanding or altering viral tropism, enhancing stability, improving transduction efficiency, specific target delivery, lower risk, and reduced immune responses, providing particualr benefits in the context of gene therapy, vaccine development, basic science, and virology research.

6. Conclusions

Viruses have an innate ability to replicate by transferring genetic information and proteins into cells. Among viral vector platforms with delivery ability, lentiviral vectors have demonstrated promising capabilities in clinical applications due to their capacity to integrate genetic information into host cells’ DNA with minimal side effects, generally outweighing their disadvantages. Ultimately, the development of next-generation vectors focused on the viral genome and pseudotyping with heterologous or engineered envelope glycoproteins have made such systems versatile and beneficial tools in various applications, emphasizing their significant potential.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/v17081036/s1, Table S1: Current gene therapy clinical trials employing lentiviral vectors for human diseases; Table S2: Regulatory approved gene therapy products at the end of 2024.

Author Contributions

Conceptualization, B.-E.J.; visualization, B.-E.J.; writing—original draft preparation, B.-E.J.; writing—review and editing, B.-E.J., M.M., and M.E.; supervision, M.E. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data supporting this review article are available in the Supplementary Files linked above.

Acknowledgments

We are deeply grateful to the entire team at the Department of Stem Cells and Human Disease Models for their invaluable support and encouragement during the development of this work.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

AAVAdeno-associated virus
AdAdenovirus
ADAAdenosine deaminase
ADA-SCIDSevere combined immunodeficiency due to the adenosine deaminase deficiency
AFPAlpha-fetoprotein
AIDSAcquired immunodeficiency syndrome
AMDAge-related macular degeneration
ASOAntisense oligonucleotides
BCMAB-cell maturation antigen
CACapsid
CALDCerebral adrenoleukodystrophy
CARChimeric antigen receptor
CasCRISPR-associated protein
CFTRCystic fibrosis membrane conductance regulator
CMVCytomegalovirus
cPPT/CTSCentral polypurine tract/central termination sequence
CRISPRClustered regularly interspaced short palindromic repeats
DNADeoxyribonucleic acid
EMAEuropean Medicines Agency
FFusion
FDAFood and Drug Administration
HBBHemoglobin subunit beta
HCCHepatocellular carcinoma
HIVHuman immunodeficiency virus
HNHemagglutinin-neuraminidase
HSCsHematopoietic stem cells
HSVHerpes Simplex Virus
IL2RGInterleukin 2 receptor subunit gamma
INIntegrase
LTRLong terminal repeat
MAMatrix
MAGE-A4Melanoma antigen gene A4
MERS-CoVMiddle east respiratory syndrome coronavirus
MESOMesothelin
MHCMajor Histocompatibility Complex
MLDMetachromatic leukodystrophy
MLVMurine leukemia virus
mRNA Messenger RNA
NCNucleocapsid
NY-ESO-1New York esophageal squamous cell carcinoma
OTCOrnithine transcarbamylase
PRProtease
RNARibonucleic acid
RRERev response element
RTReverse transcriptase
SARS-CoVSevere acute respiratory syndrome coronavirus
scFVSingle-chain variable fragment
SeVSendai virus
sgRNASingle guide RNA
shRNAShort hairpin RNA
siRNASmall interfering RNA
SIN-LTRSelf-inactivating LTR
SIVSimian immunodeficiency virus
ssRNASingle-stranded RNA
TALENTranscription activator-like effector nuclease
TCRT-cell receptor
TRETetracycline response element
vgRNAViral genomic RNA
vRNAViral RNA
VSV-GVesicular stomatitis virus glycoprotein G
WASWiskott–Aldrich Syndrome
WPREWoodchuck Hepatitis Virus Post-Transcriptional Regulatory Element
X-SCIDX-linked severe combined immunodeficiency
ZNFZinc-finger nuclease

References

  1. Centers for Disease Control (CDC). Kaposi’s Sarcoma and Pneumocystis Pneumonia among Homosexual Men—New York City and California. MMWR Morb. Mortal. Wkly. Rep. 1981, 30, 305–308. [Google Scholar]
  2. Shampo, M.A.; Kyle, R.A. Luc Montagnier—Discoverer of the AIDS Virus. Mayo Clin. Proc. 2002, 77, 506. [Google Scholar] [CrossRef] [PubMed]
  3. Engelman, A.; Cherepanov, P. The Structural Biology of HIV-1: Mechanistic and Therapeutic Insights. Nat. Rev. Microbiol. 2012, 10, 279–290. [Google Scholar] [CrossRef] [PubMed]
  4. Dwivedi, R.; Prakash, P.; Kumbhar, B.V.; Balasubramaniam, M.; Dash, C. HIV-1 Capsid and Viral DNA Integration. mBio 2024, 15, e0021222. [Google Scholar] [CrossRef] [PubMed]
  5. Mann, R.; Mulligan, R.C.; Baltimore, D. Construction of a Retrovirus Packaging Mutant and Its Use to Produce Helper-Free Defective Retrovirus. Cell 1983, 33, 153–159. [Google Scholar] [CrossRef] [PubMed]
  6. Morgenstern, J.P.; Land, H. Advanced Mammalian Gene Transfer: High Titre Retroviral Vectors with Multiple Drug Selection Markers and a Complementary Helper-Free Packaging Cell Line. Nucleic Acids Res. 1990, 18, 3587–3596. [Google Scholar] [CrossRef] [PubMed]
  7. Miller, D.G.; Adam, M.A.; Miller, A.D. Gene Transfer by Retrovirus Vectors Occurs Only in Cells That Are Actively Replicating at the Time of Infection. Mol. Cell. Biol. 1990, 10, 4239–4242. [Google Scholar] [CrossRef] [PubMed]
  8. Burns, J.C.; Friedmann, T.; Driever, W.; Burrascano, M.; Yee, J.K. Vesicular Stomatitis Virus G Glycoprotein Pseudotyped Retroviral Vectors: Concentration to Very High Titer and Efficient Gene Transfer into Mammalian and Nonmammalian Cells. Proc. Natl. Acad. Sci. USA 1993, 90, 8033–8037. [Google Scholar] [CrossRef] [PubMed]
  9. Blaese, R.M.; Culver, K.W.; Miller, A.D.; Carter, C.S.; Fleisher, T.; Clerici, M.; Shearer, G.; Chang, L.; Chiang, Y.; Tolstoshev, P.; et al. T Lymphocyte-Directed Gene Therapy for ADA—SCID: Initial Trial Results After 4 Years. Science 1995, 270, 475–480. [Google Scholar] [CrossRef] [PubMed]
  10. Rosenberg, S.A.; Aebersold, P.; Cornetta, K.; Kasid, A.; Morgan, R.A.; Moen, R.; Karson, E.M.; Lotze, M.T.; Yang, J.C.; Topalian, S.L. Gene Transfer into Humans—Immunotherapy of Patients with Advanced Melanoma, Using Tumor-Infiltrating Lymphocytes Modified by Retroviral Gene Transduction. N. Engl. J. Med. 1990, 323, 570–578. [Google Scholar] [CrossRef] [PubMed]
  11. Naldini, L.; Blömer, U.; Gallay, P.; Ory, D.; Mulligan, R.; Gage, F.H.; Verma, I.M.; Trono, D. In Vivo Gene Delivery and Stable Transduction of Nondividing Cells by a Lentiviral Vector. Science 1996, 272, 263–267. [Google Scholar] [CrossRef] [PubMed]
  12. Zufferey, R.; Nagy, D.; Mandel, R.J.; Naldini, L.; Trono, D. Multiply Attenuated Lentiviral Vector Achieves Efficient Gene Delivery In Vivo. Nat. Biotechnol. 1997, 15, 871–875. [Google Scholar] [CrossRef] [PubMed]
  13. Li, L.; Li, H.S.; Pauza, C.D.; Bukrinsky, M.; Zhao, R.Y. Roles of HIV-1 Auxiliary Proteins in Viral Pathogenesis and Host-Pathogen Interactions. Cell Res. 2005, 15, 923–934. [Google Scholar] [CrossRef] [PubMed]
  14. Miyoshi, H.; Blömer, U.; Takahashi, M.; Gage, F.H.; Verma, I.M. Development of a Self-Inactivating Lentivirus Vector. J. Virol. 1998, 72, 8150–8157. [Google Scholar] [CrossRef] [PubMed]
  15. Dull, T.; Zufferey, R.; Kelly, M.; Mandel, R.J.; Nguyen, M.; Trono, D.; Naldini, L. A Third-Generation Lentivirus Vector with a Conditional Packaging System. J. Virol. 1998, 72, 8463–8471. [Google Scholar] [CrossRef] [PubMed]
  16. Zufferey, R.; Dull, T.; Mandel, R.J.; Bukovsky, A.; Quiroz, D.; Naldini, L.; Trono, D. Self-Inactivating Lentivirus Vector for Safe and Efficient In Vivo Gene Delivery. J. Virol. 1998, 72, 9873–9880. [Google Scholar] [CrossRef] [PubMed]
  17. Hanawa, H.; Persons, D.A.; Nienhuis, A.W. Mobilization and Mechanism of Transcription of Integrated Self-Inactivating Lentiviral Vectors. J. Virol. 2005, 79, 8410–8421. [Google Scholar] [CrossRef] [PubMed]
  18. Xu, W.; Russ, J.L.; Eiden, M. V Evaluation of Residual Promoter Activity in γ-Retroviral Self-Inactivating (SIN) Vectors. Mol. Ther. 2012, 20, 84–90. [Google Scholar] [CrossRef] [PubMed]
  19. Cavazza, A.; Cocchiarella, F.; Bartholomae, C.; Schmidt, M.; Pincelli, C.; Larcher, F.; Mavilio, F. Self-Inactivating MLV Vectors Have a Reduced Genotoxic Profile in Human Epidermal Keratinocytes. Gene Ther. 2013, 20, 949–957. [Google Scholar] [CrossRef] [PubMed]
  20. Wolff, J.H.; Mikkelsen, J.G. Delivering Genes with Human Immunodeficiency Virus-Derived Vehicles: Still State-of-the-Art after 25 Years. J. Biomed. Sci. 2022, 29, 79. [Google Scholar] [CrossRef] [PubMed]
  21. Zufferey, R.; Donello, J.E.; Trono, D.; Hope, T.J. Woodchuck Hepatitis Virus Posttranscriptional Regulatory Element Enhances Expression of Transgenes Delivered by Retroviral Vectors. J. Virol. 1999, 73, 2886–2892. [Google Scholar] [CrossRef] [PubMed]
  22. Follenzi, A.; Ailles, L.E.; Bakovic, S.; Geuna, M.; Naldini, L. Gene Transfer by Lentiviral Vectors Is Limited by Nuclear Translocation and Rescued by HIV-1 Pol Sequences. Nat. Genet. 2000, 25, 217–222. [Google Scholar] [CrossRef] [PubMed]
  23. Zennou, V.; Petit, C.; Guetard, D.; Nerhbass, U.; Montagnier, L.; Charneau, P. HIV-1 Genome Nuclear Import Is Mediated by a Central DNA Flap. Cell 2000, 101, 173–185. [Google Scholar] [CrossRef] [PubMed]
  24. Roe, T.; Reynolds, T.C.; Yu, G.; Brown, P.O. Integration of Murine Leukemia Virus DNA Depends on Mitosis. EMBO J. 1993, 12, 2099–2108. [Google Scholar] [CrossRef] [PubMed]
  25. Leavitt, A.D.; Robles, G.; Alesandro, N.; Varmus, H.E. Human Immunodeficiency Virus Type 1 Integrase Mutants Retain In Vitro Integrase Activity yet Fail to Integrate Viral DNA Efficiently during Infection. J. Virol. 1996, 70, 721–728. [Google Scholar] [CrossRef] [PubMed]
  26. Yáñez-Muñoz, R.J.; Balaggan, K.S.; MacNeil, A.; Howe, S.J.; Schmidt, M.; Smith, A.J.; Buch, P.; MacLaren, R.E.; Anderson, P.N.; Barker, S.E.; et al. Effective Gene Therapy with Nonintegrating Lentiviral Vectors. Nat. Med. 2006, 12, 348–353. [Google Scholar] [CrossRef] [PubMed]
  27. Philippe, S.; Sarkis, C.; Barkats, M.; Mammeri, H.; Ladroue, C.; Petit, C.; Mallet, J.; Serguera, C. Lentiviral Vectors with a Defective Integrase Allow Efficient and Sustained Transgene Expression In Vitro and In Vivo. Proc. Natl. Acad. Sci. USA 2006, 103, 17684–17689. [Google Scholar] [CrossRef] [PubMed]
  28. Qamar Saeed, M.; Dufour, N.; Bartholmae, C.; Sieranska, U.; Knopf, M.; Thierry, E.; Thierry, S.; Delelis, O.; Grandchamp, N.; Pilet, H.; et al. Comparison Between Several Integrase-Defective Lentiviral Vectors Reveals Increased Integration of an HIV Vector Bearing a D167H Mutant. Mol. Ther. Nucleic Acids 2014, 3, e213. [Google Scholar] [CrossRef] [PubMed]
  29. MacGregor, R.R. Clinical Protocol. A Phase 1 Open-Label Clinical Trial of the Safety and Tolerability of Single Escalating Doses of Autologous CD4 T Cells Transduced with VRX496 in HIV-Positive Subjects. Hum. Gene Ther. 2001, 12, 2028–2029. [Google Scholar] [PubMed]
  30. Levine, B.L.; Humeau, L.M.; Boyer, J.; MacGregor, R.-R.; Rebello, T.; Lu, X.; Binder, G.K.; Slepushkin, V.; Lemiale, F.; Mascola, J.R.; et al. Gene Transfer in Humans Using a Conditionally Replicating Lentiviral Vector. Proc. Natl. Acad. Sci. USA 2006, 103, 17372–17377. [Google Scholar] [CrossRef] [PubMed]
  31. McGarrity, G.J.; Hoyah, G.; Winemiller, A.; Andre, K.; Stein, D.; Blick, G.; Greenberg, R.N.; Kinder, C.; Zolopa, A.; Binder-Scholl, G.; et al. Patient Monitoring and Follow-up in Lentiviral Clinical Trials. J. Gene Med. 2013, 15, 78–82. [Google Scholar] [CrossRef] [PubMed]
  32. Yu, W.-L.; Hua, Z.-C. Chimeric Antigen Receptor T-Cell (CAR T) Therapy for Hematologic and Solid Malignancies: Efficacy and Safety—A Systematic Review with Meta-Analysis. Cancers 2019, 11, 47. [Google Scholar] [CrossRef] [PubMed]
  33. Yu, X.-J.; Liu, C.; Hu, S.-Z.; Yuan, Z.-Y.; Ni, H.-Y.; Sun, S.-J.; Hu, C.-Y.; Zhan, H.-Q. Application of CAR-T Cell Therapy in B-Cell Lymphoma: A Meta-Analysis of Randomized Controlled Trials. Clin. Transl. Oncol. 2024, 27, 2700–2709. [Google Scholar] [CrossRef] [PubMed]
  34. Yarza, R.; Bover, M.; Herrera-Juarez, M.; Rey-Cardenas, M.; Paz-Ares, L.; Lopez-Martin, J.A.; Haanen, J. Efficacy of T-Cell Receptor-Based Adoptive Cell Therapy in Cutaneous Melanoma: A Meta-Analysis. Oncologist 2023, 28, e406–e415. [Google Scholar] [CrossRef] [PubMed]
  35. Gross, G.; Waks, T.; Eshhar, Z. Expression of Immunoglobulin-T-Cell Receptor Chimeric Molecules as Functional Receptors with Antibody-Type Specificity. Proc. Natl. Acad. Sci. USA 1989, 86, 10024–10028. [Google Scholar] [CrossRef] [PubMed]
  36. Blüthmann, H.; Kisielow, P.; Uematsu, Y.; Malissen, M.; Krimpenfort, P.; Berns, A.; von Boehmer, H.; Steinmetz, M. T-Cell-Specific Deletion of T-Cell Receptor Transgenes Allows Functional Rearrangement of Endogenous Alpha- and Beta-Genes. Nature 1988, 334, 156–159. [Google Scholar] [CrossRef] [PubMed]
  37. Ramesh, P.; Hui, H.Y.L.; Brownrigg, L.M.; Fuller, K.A.; Erber, W.N. Chimeric Antigen Receptor T-cells: Properties, Production, and Quality Control. Int. J. Lab. Hematol. 2023, 45, 425–435. [Google Scholar] [CrossRef] [PubMed]
  38. Blache, U.; Popp, G.; Dünkel, A.; Koehl, U.; Fricke, S. Potential Solutions for Manufacture of CAR T Cells in Cancer Immunotherapy. Nat. Commun. 2022, 13, 5225. [Google Scholar] [CrossRef] [PubMed]
  39. June, C.H.; O’Connor, R.S.; Kawalekar, O.U.; Ghassemi, S.; Milone, M.C. CAR T Cell Immunotherapy for Human Cancer. Science 2018, 359, 1361–1365. [Google Scholar] [CrossRef] [PubMed]
  40. Golikova, E.A.; Alshevskaya, A.A.; Alrhmoun, S.; Sivitskaya, N.A.; Sennikov, S.V. TCR-T Cell Therapy: Current Development Approaches, Preclinical Evaluation, and Perspectives on Regulatory Challenges. J. Transl. Med. 2024, 22, 897. [Google Scholar] [CrossRef] [PubMed]
  41. Rapoport, A.P.; Stadtmauer, E.A.; Binder-Scholl, G.K.; Goloubeva, O.; Vogl, D.T.; Lacey, S.F.; Badros, A.Z.; Garfall, A.; Weiss, B.; Finklestein, J.; et al. NY-ESO-1-Specific TCR-Engineered T Cells Mediate Sustained Antigen-Specific Antitumor Effects in Myeloma. Nat. Med. 2015, 21, 914–921. [Google Scholar] [CrossRef] [PubMed]
  42. Hong, D.S.; Van Tine, B.A.; Biswas, S.; McAlpine, C.; Johnson, M.L.; Olszanski, A.J.; Clarke, J.M.; Araujo, D.; Blumenschein, G.R.; Kebriaei, P.; et al. Autologous T Cell Therapy for MAGE-A4+ Solid Cancers in HLA-A*02+ Patients: A Phase 1 Trial. Nat. Med. 2023, 29, 104–114. [Google Scholar] [CrossRef] [PubMed]
  43. Chandora, K.; Chandora, A.; Saeed, A.; Cavalcante, L. Adoptive T Cell Therapy Targeting MAGE-A4. Cancers 2025, 17, 413. [Google Scholar] [CrossRef] [PubMed]
  44. Ozer, M.; Goksu, S.Y.; Akagunduz, B.; George, A.; Sahin, I. Adoptive Cell Therapy in Hepatocellular Carcinoma: A Review of Clinical Trials. Cancers 2023, 15, 1808. [Google Scholar] [CrossRef] [PubMed]
  45. Hassan, R.; Butler, M.; O’Cearbhaill, R.E.; Oh, D.Y.; Johnson, M.; Zikaras, K.; Smalley, M.; Ross, M.; Tanyi, J.L.; Ghafoor, A.; et al. Mesothelin-Targeting T Cell Receptor Fusion Construct Cell Therapy in Refractory Solid Tumors: Phase 1/2 Trial Interim Results. Nat. Med. 2023, 29, 2099–2109. [Google Scholar] [CrossRef] [PubMed]
  46. Brudno, J.N.; Maus, M.V.; Hinrichs, C.S. CAR T Cells and T-Cell Therapies for Cancer: A Translational Science Review. JAMA 2024, 332, 1924–1935. [Google Scholar] [CrossRef] [PubMed]
  47. Schuster, S.J.; Bishop, M.R.; Tam, C.S.; Waller, E.K.; Borchmann, P.; McGuirk, J.P.; Jäger, U.; Jaglowski, S.; Andreadis, C.; Westin, J.R.; et al. Tisagenlecleucel in Adult Relapsed or Refractory Diffuse Large B-Cell Lymphoma. N. Engl. J. Med. 2019, 380, 45–56. [Google Scholar] [CrossRef] [PubMed]
  48. Locke, F.L.; Miklos, D.B.; Jacobson, C.A.; Perales, M.-A.; Kersten, M.-J.; Oluwole, O.O.; Ghobadi, A.; Rapoport, A.P.; McGuirk, J.; Pagel, J.M.; et al. Axicabtagene Ciloleucel as Second-Line Therapy for Large B-Cell Lymphoma. N. Engl. J. Med. 2022, 386, 640–654. [Google Scholar] [CrossRef] [PubMed]
  49. Wang, Y.; Jain, P.; Locke, F.L.; Maurer, M.J.; Frank, M.J.; Munoz, J.L.; Dahiya, S.; Beitinjaneh, A.M.; Jacobs, M.T.; Mcguirk, J.P.; et al. Brexucabtagene Autoleucel for Relapsed or Refractory Mantle Cell Lymphoma in Standard-of-Care Practice: Results from the US Lymphoma CAR T Consortium. J. Clin. Oncol. 2023, 41, 2594–2606. [Google Scholar] [CrossRef] [PubMed]
  50. Abramson, J.S.; Solomon, S.R.; Arnason, J.; Johnston, P.B.; Glass, B.; Bachanova, V.; Ibrahimi, S.; Mielke, S.; Mutsaers, P.; Hernandez-Ilizaliturri, F.; et al. Lisocabtagene Maraleucel as Second-Line Therapy for Large B-Cell Lymphoma: Primary Analysis of the Phase 3 TRANSFORM Study. Blood 2023, 141, 1675–1684. [Google Scholar] [CrossRef] [PubMed]
  51. Ying, Z.; Yang, H.; Guo, Y.; Li, W.; Zou, D.; Zhou, D.; Wang, Z.; Zhang, M.; Wu, J.; Liu, H.; et al. Relmacabtagene Autoleucel (Relma-Cel) CD19 CAR-T Therapy for Adults with Heavily Pretreated Relapsed/Refractory Large B-Cell Lymphoma in China. Cancer Med. 2021, 10, 999–1011. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, Y.; Lv, L.; Song, Y.; Wei, X.; Zhou, H.; Liu, Q.; Xu, K.; Yan, D.; Zhang, C.; Liu, S.; et al. Inaticabtagene Autoleucel in Adult Relapsed or Refractory B-Cell Acute Lymphoblastic Leukemia. Blood Adv. 2025, 9, 836–843. [Google Scholar] [CrossRef] [PubMed]
  53. Roddie, C.; Sandhu, K.S.; Tholouli, E.; Logan, A.C.; Shaughnessy, P.; Barba, P.; Ghobadi, A.; Guerreiro, M.; Yallop, D.; Abedi, M.; et al. Obecabtagene Autoleucel in Adults with B-Cell Acute Lymphoblastic Leukemia. N. Engl. J. Med. 2024, 391, 2219–2230. [Google Scholar] [CrossRef] [PubMed]
  54. Sheykhhasan, M.; Ahmadieh-Yazdi, A.; Vicidomini, R.; Poondla, N.; Tanzadehpanah, H.; Dirbaziyan, A.; Mahaki, H.; Manoochehri, H.; Kalhor, N.; Dama, P. CAR T Therapies in Multiple Myeloma: Unleashing the Future. Cancer Gene Ther. 2024, 31, 667–686. [Google Scholar] [CrossRef] [PubMed]
  55. Munshi, N.C.; Anderson, L.D.; Shah, N.; Madduri, D.; Berdeja, J.; Lonial, S.; Raje, N.; Lin, Y.; Siegel, D.; Oriol, A.; et al. Idecabtagene Vicleucel in Relapsed and Refractory Multiple Myeloma. N. Engl. J. Med. 2021, 384, 705–716. [Google Scholar] [CrossRef] [PubMed]
  56. Martin, T.; Usmani, S.Z.; Berdeja, J.G.; Agha, M.; Cohen, A.D.; Hari, P.; Avigan, D.; Deol, A.; Htut, M.; Lesokhin, A.; et al. Ciltacabtagene Autoleucel, an Anti-B-Cell Maturation Antigen Chimeric Antigen Receptor T-Cell Therapy, for Relapsed/Refractory Multiple Myeloma: CARTITUDE-1 2-Year Follow-Up. J. Clin. Oncol. 2023, 41, 1265–1274. [Google Scholar] [CrossRef] [PubMed]
  57. Keam, S.J. Equecabtagene Autoleucel: First Approval. Mol. Diagn. Ther. 2023, 27, 781–787. [Google Scholar] [CrossRef] [PubMed]
  58. Dhillon, S. Zevorcabtagene Autoleucel: First Approval. Mol. Diagn. Ther. 2024, 28, 501–506. [Google Scholar] [CrossRef] [PubMed]
  59. Wei, J.; Han, X.; Bo, J.; Han, W. Target Selection for CAR-T Therapy. J. Hematol. Oncol. 2019, 12, 62. [Google Scholar] [CrossRef] [PubMed]
  60. Wang, Y.; Chen, M.; Wu, Z.; Tong, C.; Dai, H.; Guo, Y.; Liu, Y.; Huang, J.; Lv, H.; Luo, C.; et al. CD133-Directed CAR T Cells for Advanced Metastasis Malignancies: A Phase I Trial. Oncoimmunology 2018, 7, e1440169. [Google Scholar] [CrossRef] [PubMed]
  61. Kulczycka, M.; Derlatka, K.; Tasior, J.; Lejman, M.; Zawitkowska, J. CAR T-Cell Therapy in Children with Solid Tumors. J. Clin. Med. 2023, 12, 2326. [Google Scholar] [CrossRef] [PubMed]
  62. Larson, S.M.; Walthers, C.M.; Ji, B.; Ghafouri, S.N.; Naparstek, J.; Trent, J.; Chen, J.M.; Roshandell, M.; Harris, C.; Khericha, M.; et al. CD19/CD20 Bispecific Chimeric Antigen Receptor (CAR) in Naive/Memory T Cells for the Treatment of Relapsed or Refractory Non-Hodgkin Lymphoma. Cancer Discov. 2023, 13, 580–597. [Google Scholar] [CrossRef] [PubMed]
  63. Frank, M.J.; Baird, J.H.; Kramer, A.M.; Srinagesh, H.K.; Patel, S.; Brown, A.K.; Oak, J.S.; Younes, S.F.; Natkunam, Y.; Hamilton, M.P.; et al. CD22-Directed CAR T-Cell Therapy for Large B-Cell Lymphomas Progressing after CD19-Directed CAR T-Cell Therapy: A Dose-Finding Phase 1 Study. Lancet 2024, 404, 353–363. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, X.; Lv, H.; Xiao, X.; Bai, X.; Liu, P.; Pu, Y.; Meng, J.; Zhu, H.; Wang, Z.; Zhang, H.; et al. A Phase I Clinical Trial of CLL-1 CAR-T Cells for the Treatment of Relapsed/Refractory Acute Myeloid Leukemia in Adults. Blood 2023, 142, 2106. [Google Scholar] [CrossRef]
  65. Mamcarz, E.; Zhou, S.; Lockey, T.; Abdelsamed, H.; Cross, S.J.; Kang, G.; Ma, Z.; Condori, J.; Dowdy, J.; Triplett, B.; et al. Lentiviral Gene Therapy Combined with Low-Dose Busulfan in Infants with SCID-X1. N. Engl. J. Med. 2019, 380, 1525–1534. [Google Scholar] [CrossRef] [PubMed]
  66. Kanter, J.; Chawla, A.; Thompson, A.A.; Kwiatkowski, J.L.; Parikh, S.; Mapara, M.Y.; Rifkin-Zenenberg, S.; Aygun, B.; Kasow, K.A.; Gupta, A.O.; et al. Lovotibeglogene Autotemcel Gene Therapy for Sickle Cell Disease: 60 Months Follow-Up. J. Sick. Cell Dis. 2024, 1, yoae002.002. [Google Scholar] [CrossRef]
  67. Locatelli, F.; Thompson, A.A.; Kwiatkowski, J.L.; Porter, J.B.; Thrasher, A.J.; Hongeng, S.; Sauer, M.G.; Thuret, I.; Lal, A.; Algeri, M.; et al. Betibeglogene Autotemcel Gene Therapy for Non–β00 Genotype β-Thalassemia. N. Engl. J. Med. 2022, 386, 415–427. [Google Scholar] [CrossRef] [PubMed]
  68. Czechowicz, A.D.; Sevilla, J.; Booth, C.; Zubicaray, J.; Rio, P.; Navarro, S.; Cherry, K.; O’Toole, G.; Xu-Bayford, J.; Ancliff, P.; et al. Lentiviral-Mediated Gene Therapy for Patients with Fanconi Anemia [Group a]: Updated Results from Global RP-L102 Clinical Trials. Blood 2024, 144, 7463. [Google Scholar] [CrossRef]
  69. Fumagalli, F.; Calbi, V.; Natali Sora, M.G.; Sessa, M.; Baldoli, C.; Rancoita, P.M.V.; Ciotti, F.; Sarzana, M.; Fraschini, M.; Zambon, A.A.; et al. Lentiviral Haematopoietic Stem-Cell Gene Therapy for Early-Onset Metachromatic Leukodystrophy: Long-Term Results from a Non-Randomised, Open-Label, Phase 1/2 Trial and Expanded Access. Lancet 2022, 399, 372–383. [Google Scholar] [CrossRef] [PubMed]
  70. Eichler, F.; Duncan, C.N.; Musolino, P.L.; Lund, T.C.; Gupta, A.O.; De Oliveira, S.; Thrasher, A.J.; Aubourg, P.; Kühl, J.-S.; Loes, D.J.; et al. Lentiviral Gene Therapy for Cerebral Adrenoleukodystrophy. N. Engl. J. Med. 2024, 391, 1302–1312. [Google Scholar] [CrossRef] [PubMed]
  71. Srivastava, A.; Abraham, A.; Aboobacker, F.; Singh, G.; Geevar, T.; Kulkarni, U.; Selvarajan, S.; Korula, A.; Dave, R.G.; Shankar, M.; et al. Lentiviral Gene Therapy with CD34+ Hematopoietic Cells for Hemophilia A. N. Engl. J. Med. 2025, 392, 450–457. [Google Scholar] [CrossRef] [PubMed]
  72. Magnani, A.; Semeraro, M.; Adam, F.; Booth, C.; Dupré, L.; Morris, E.C.; Gabrion, A.; Roudaut, C.; Borgel, D.; Toubert, A.; et al. Long-Term Safety and Efficacy of Lentiviral Hematopoietic Stem/Progenitor Cell Gene Therapy for Wiskott-Aldrich Syndrome. Nat. Med. 2022, 28, 71–80. [Google Scholar] [CrossRef] [PubMed]
  73. Palfi, S.; Gurruchaga, J.M.; Ralph, G.S.; Lepetit, H.; Lavisse, S.; Buttery, P.C.; Watts, C.; Miskin, J.; Kelleher, M.; Deeley, S.; et al. Long-Term Safety and Tolerability of ProSavin, a Lentiviral Vector-Based Gene Therapy for Parkinson’s Disease: A Dose Escalation, Open-Label, Phase 1/2 Trial. Lancet 2014, 383, 1138–1146. [Google Scholar] [CrossRef] [PubMed]
  74. Campochiaro, P.A.; Lauer, A.K.; Sohn, E.H.; Mir, T.A.; Naylor, S.; Anderton, M.C.; Kelleher, M.; Harrop, R.; Ellis, S.; Mitrophanous, K.A. Lentiviral Vector Gene Transfer of Endostatin/Angiostatin for Macular Degeneration (GEM) Study. Hum. Gene Ther. 2017, 28, 99–111. [Google Scholar] [CrossRef] [PubMed]
  75. Parker, M.A.; Erker, L.R.; Audo, I.; Choi, D.; Mohand-Said, S.; Sestakauskas, K.; Benoit, P.; Appelqvist, T.; Krahmer, M.; Ségaut-Prévost, C.; et al. Three-Year Safety Results of SAR422459 (EIAV-ABCA4) Gene Therapy in Patients with ABCA4-Associated Stargardt Disease: An Open-Label Dose-Escalation Phase I/IIa Clinical Trial, Cohorts 1–5. Am. J. Ophthalmol. 2022, 240, 285–301. [Google Scholar] [CrossRef] [PubMed]
  76. Ottaviano, G.; Qasim, W. Current Landscape of Vector Safety and Genotoxicity after Hematopoietic Stem or Immune Cell Gene Therapy. Leukemia 2025, 39, 1325–1333. [Google Scholar] [CrossRef] [PubMed]
  77. Zhao, Q.; Wei, L.; Chen, Y. From Bench to Bedside: Developing CRISPR/Cas-Based Therapy for Ocular Diseases. Pharmacol. Res. 2025, 213, 107638. [Google Scholar] [CrossRef] [PubMed]
  78. U.S. Food and Drug Administration. Long Term Follow-Up After Administration of Human Gene Therapy Products Guidance for Industry. Available online: https://www.fda.gov/media/113768/download (accessed on 28 January 2025).
  79. European Medicines Agency. Guideline on the Quality, Non-Clinical and Clinical Aspects of Gene Therapy Medicinal Products. Available online: https://www.ema.europa.eu/en/documents/scientific-guideline/guideline-quality-non-clinical-and-clinical-aspects-gene-therapy-medicinal-products_en.pdf (accessed on 28 January 2025).
  80. Vitravene Study Group. A Randomized Controlled Clinical Trial of Intravitreous Fomivirsen for Treatment of Newly Diagnosed Peripheral Cytomegalovirus Retinitis in Patients with AIDS. Am. J. Ophthalmol. 2002, 133, 467–474. [Google Scholar] [CrossRef]
  81. Barresi, V.; Musmeci, C.; Rinaldi, A.; Condorelli, D.F. Transcript-Targeted Therapy Based on RNA Interference and Antisense Oligonucleotides: Current Applications and Novel Molecular Targets. Int. J. Mol. Sci. 2022, 23, 8875. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, W.-W.; Li, L.; Li, D.; Liu, J.; Li, X.; Li, W.; Xu, X.; Zhang, M.J.; Chandler, L.A.; Lin, H.; et al. The First Approved Gene Therapy Product for Cancer Ad-p53 (Gendicine): 12 Years in the Clinic. Hum. Gene Ther. 2018, 29, 160–179. [Google Scholar] [CrossRef] [PubMed]
  83. Ng, E.W.M.; Shima, D.T.; Calias, P.; Cunningham, E.T.; Guyer, D.R.; Adamis, A.P. Pegaptanib, a Targeted Anti-VEGF Aptamer for Ocular Vascular Disease. Nat. Rev. Drug Discov. 2006, 5, 123–132. [Google Scholar] [CrossRef] [PubMed]
  84. Mondal, M.; Guo, J.; He, P.; Zhou, D. Recent Advances of Oncolytic Virus in Cancer Therapy. Hum. Vaccin. Immunother. 2020, 16, 2389–2402. [Google Scholar] [CrossRef] [PubMed]
  85. Chawla, S.P.; Bruckner, H.; Morse, M.A.; Assudani, N.; Hall, F.L.; Gordon, E.M. A Phase I-II Study Using Rexin-G Tumor-Targeted Retrovector Encoding a Dominant-Negative Cyclin G1 Inhibitor for Advanced Pancreatic Cancer. Mol. Ther.—Oncol. 2019, 12, 56–67. [Google Scholar] [CrossRef] [PubMed]
  86. Deev, R.; Plaksa, I.; Bozo, I.; Mzhavanadze, N.; Suchkov, I.; Chervyakov, Y.; Staroverov, I.; Kalinin, R.; Isaev, A. Results of 5-Year Follow-up Study in Patients with Peripheral Artery Disease Treated with PL-VEGF165 for Intermittent Claudication. Ther. Adv. Cardiovasc. Dis. 2018, 12, 237–246. [Google Scholar] [CrossRef] [PubMed]
  87. Bryant, L.M.; Christopher, D.M.; Giles, A.R.; Hinderer, C.; Rodriguez, J.L.; Smith, J.B.; Traxler, E.A.; Tycko, J.; Wojno, A.P.; Wilson, J.M. Lessons Learned from the Clinical Development and Market Authorization of Glybera. Hum. Gene Ther. Clin. Dev. 2013, 24, 55–64. [Google Scholar] [CrossRef] [PubMed]
  88. Geary, R.S.; Baker, B.F.; Crooke, S.T. Clinical and Preclinical Pharmacokinetics and Pharmacodynamics of Mipomersen (Kynamro®): A Second-Generation Antisense Oligonucleotide Inhibitor of Apolipoprotein B. Clin. Pharmacokinet. 2015, 54, 133–146. [Google Scholar] [CrossRef] [PubMed]
  89. Ferrucci, P.F.; Pala, L.; Conforti, F.; Cocorocchio, E. Talimogene Laherparepvec (T-VEC): An Intralesional Cancer Immunotherapy for Advanced Melanoma. Cancers 2021, 13, 1383. [Google Scholar] [CrossRef] [PubMed]
  90. Lim, K.R.Q.; Maruyama, R.; Yokota, T. Eteplirsen in the Treatment of Duchenne Muscular Dystrophy. Drug Des. Dev. Ther. 2017, 11, 533–545. [Google Scholar] [CrossRef] [PubMed]
  91. Sumner, C.J.; Crawford, T.O. Two Breakthrough Gene-Targeted Treatments for Spinal Muscular Atrophy: Challenges Remain. J. Clin. Investig. 2018, 128, 3219–3227. [Google Scholar] [CrossRef] [PubMed]
  92. Migliavacca, M.; Barzaghi, F.; Fossati, C.; Rancoita, P.M.V.; Gabaldo, M.; Dionisio, F.; Giannelli, S.; Salerio, F.A.; Ferrua, F.; Tucci, F.; et al. Long-Term and Real-World Safety and Efficacy of Retroviral Gene Therapy for Adenosine Deaminase Deficiency. Nat. Med. 2024, 30, 488–497. [Google Scholar] [CrossRef] [PubMed]
  93. Mohty, M.; Labopin, M.; Velardi, A.; van Lint, M.T.; Bunjes, D.; Bruno, B.; Santarone, S.; Tischer, J.; Koc, Y.; Wu, D.; et al. Allogeneic Genetically Modified T Cells (HSV-TK) As Adjunctive Treatment in Haploidentical Hematopoietic Stem-Cell Transplantation (Haplo-HSCT) of Adult Patients with High-Risk Hematological Malignancies: A Pair-Matched Analysis from the Acute Leukemia Wo. Blood 2016, 128, 672. [Google Scholar] [CrossRef]
  94. Mitchell, W.M. Efficacy of Rintatolimod in the Treatment of Chronic Fatigue Syndrome/Myalgic Encephalomyelitis (CFS/ME). Expert Rev. Clin. Pharmacol. 2016, 9, 755–770. [Google Scholar] [CrossRef] [PubMed]
  95. Testa, F.; Bacci, G.; Falsini, B.; Iarossi, G.; Melillo, P.; Mucciolo, D.P.; Murro, V.; Salvetti, A.P.; Sodi, A.; Staurenghi, G.; et al. Voretigene Neparvovec for Inherited Retinal Dystrophy Due to RPE65 Mutations: A Scoping Review of Eligibility and Treatment Challenges from Clinical Trials to Real Practice. Eye 2024, 38, 2504–2515. [Google Scholar] [CrossRef] [PubMed]
  96. Kim, M.-K.; Ha, C.-W.; In, Y.; Cho, S.-D.; Choi, E.-S.; Ha, J.-K.; Lee, J.-H.; Yoo, J.-D.; Bin, S.-I.; Choi, C.-H.; et al. A Multicenter, Double-Blind, Phase III Clinical Trial to Evaluate the Efficacy and Safety of a Cell and Gene Therapy in Knee Osteoarthritis Patients. Hum. Gene Ther. Clin. Dev. 2018, 29, 48–59. [Google Scholar] [CrossRef] [PubMed]
  97. Maurer, M.S.; Kale, P.; Fontana, M.; Berk, J.L.; Grogan, M.; Gustafsson, F.; Hung, R.R.; Gottlieb, R.L.; Damy, T.; González-Duarte, A.; et al. Patisiran Treatment in Patients with Transthyretin Cardiac Amyloidosis. N. Engl. J. Med. 2023, 389, 1553–1565. [Google Scholar] [CrossRef] [PubMed]
  98. Benson, M.D.; Waddington-Cruz, M.; Berk, J.L.; Polydefkis, M.; Dyck, P.J.; Wang, A.K.; Planté-Bordeneuve, V.; Barroso, F.A.; Merlini, G.; Obici, L.; et al. Inotersen Treatment for Patients with Hereditary Transthyretin Amyloidosis. N. Engl. J. Med. 2018, 379, 22–31. [Google Scholar] [CrossRef] [PubMed]
  99. Strauss, K.A.; Farrar, M.A.; Muntoni, F.; Saito, K.; Mendell, J.R.; Servais, L.; McMillan, H.J.; Finkel, R.S.; Swoboda, K.J.; Kwon, J.M.; et al. Onasemnogene Abeparvovec for Presymptomatic Infants with Two Copies of SMN2 at Risk for Spinal Muscular Atrophy Type 1: The Phase III SPR1NT Trial. Nat. Med. 2022, 28, 1381–1389. [Google Scholar] [CrossRef] [PubMed]
  100. Komatsuno, T. The First Gene Therapy Product in Japan Collategene® Intramuscular Injection. Drug Deliv. Syst. 2019, 34, 305–308. [Google Scholar] [CrossRef]
  101. Dickey, A.K.; Leaf, R.K. Givosiran: A Targeted Treatment for Acute Intermittent Porphyria. Hematol. Am. Soc. Hematol. Educ. Progr. 2024, 2024, 426–433. [Google Scholar] [CrossRef] [PubMed]
  102. Scaglioni, D.; Catapano, F.; Ellis, M.; Torelli, S.; Chambers, D.; Feng, L.; Beck, M.; Sewry, C.; Monforte, M.; Harriman, S.; et al. The Administration of Antisense Oligonucleotide Golodirsen Reduces Pathological Regeneration in Patients with Duchenne Muscular Dystrophy. Acta Neuropathol. Commun. 2021, 9, 7. [Google Scholar] [CrossRef] [PubMed]
  103. Witztum, J.L.; Gaudet, D.; Freedman, S.D.; Alexander, V.J.; Digenio, A.; Williams, K.R.; Yang, Q.; Hughes, S.G.; Geary, R.S.; Arca, M.; et al. Volanesorsen and Triglyceride Levels in Familial Chylomicronemia Syndrome. N. Engl. J. Med. 2019, 381, 531–542. [Google Scholar] [CrossRef] [PubMed]
  104. Roshmi, R.R.; Yokota, T. Pharmacological Profile of Viltolarsen for the Treatment of Duchenne Muscular Dystrophy: A Japanese Experience. Clin. Pharmacol. 2021, 13, 235–242. [Google Scholar] [CrossRef] [PubMed]
  105. Garrelfs, S.F.; Frishberg, Y.; Hulton, S.A.; Koren, M.J.; O’Riordan, W.D.; Cochat, P.; Deschênes, G.; Shasha-Lavsky, H.; Saland, J.M.; Van’t Hoff, W.G.; et al. Lumasiran, an RNAi Therapeutic for Primary Hyperoxaluria Type 1. N. Engl. J. Med. 2021, 384, 1216–1226. [Google Scholar] [CrossRef] [PubMed]
  106. Wagner, K.R.; Kuntz, N.L.; Koenig, E.; East, L.; Upadhyay, S.; Han, B.; Shieh, P.B. Safety, Tolerability, and Pharmacokinetics of Casimersen in Patients with Duchenne Muscular Dystrophy Amenable to Exon 45 Skipping: A Randomized, Double-Blind, Placebo-Controlled, Dose-Titration Trial. Muscle Nerve 2021, 64, 285–292. [Google Scholar] [CrossRef] [PubMed]
  107. Ray, K.K.; Troquay, R.P.T.; Visseren, F.L.J.; Leiter, L.A.; Scott Wright, R.; Vikarunnessa, S.; Talloczy, Z.; Zang, X.; Maheux, P.; Lesogor, A.; et al. Long-Term Efficacy and Safety of Inclisiran in Patients with High Cardiovascular Risk and Elevated LDL Cholesterol (ORION-3): Results from the 4-Year Open-Label Extension of the ORION-1 Trial. Lancet Diabetes Endocrinol. 2023, 11, 109–119. [Google Scholar] [CrossRef] [PubMed]
  108. Frampton, J.E. Teserpaturev/G47Δ: First Approval. BioDrugs 2022, 36, 667–672. [Google Scholar] [CrossRef] [PubMed]
  109. Fontana, M.; Berk, J.L.; Gillmore, J.D.; Witteles, R.M.; Grogan, M.; Drachman, B.; Damy, T.; Garcia-Pavia, P.; Taubel, J.; Solomon, S.D.; et al. Vutrisiran in Patients with Transthyretin Amyloidosis with Cardiomyopathy. N. Engl. J. Med. 2025, 392, 33–44. [Google Scholar] [CrossRef] [PubMed]
  110. Tai, C.-H.; Lee, N.-C.; Chien, Y.-H.; Byrne, B.J.; Muramatsu, S.-I.; Tseng, S.-H.; Hwu, W.-L. Long-Term Efficacy and Safety of Eladocagene Exuparvovec in Patients with AADC Deficiency. Mol. Ther. 2022, 30, 509–518. [Google Scholar] [CrossRef] [PubMed]
  111. Lee, A. Nadofaragene Firadenovec: First Approval. Drugs 2023, 83, 353–357. [Google Scholar] [CrossRef] [PubMed]
  112. Pipe, S.W.; Leebeek, F.W.G.; Recht, M.; Key, N.S.; Castaman, G.; Miesbach, W.; Lattimore, S.; Peerlinck, K.; Van der Valk, P.; Coppens, M.; et al. Gene Therapy with Etranacogene Dezaparvovec for Hemophilia B. N. Engl. J. Med. 2023, 388, 706–718. [Google Scholar] [CrossRef] [PubMed]
  113. Ozelo, M.C.; Mahlangu, J.; Pasi, K.J.; Giermasz, A.; Leavitt, A.D.; Laffan, M.; Symington, E.; Quon, D.V.; Wang, J.-D.; Peerlinck, K.; et al. Valoctocogene Roxaparvovec Gene Therapy for Hemophilia A. N. Engl. J. Med. 2022, 386, 1013–1025. [Google Scholar] [CrossRef] [PubMed]
  114. Frangoul, H.; Locatelli, F.; Sharma, A.; Bhatia, M.; Mapara, M.; Molinari, L.; Wall, D.; Liem, R.I.; Telfer, P.; Shah, A.J.; et al. Exagamglogene Autotemcel for Severe Sickle Cell Disease. N. Engl. J. Med. 2024, 390, 1649–1662. [Google Scholar] [CrossRef] [PubMed]
  115. Dhillon, S. Beremagene Geperpavec: First Approval. Drugs 2023, 83, 1131–1135. [Google Scholar] [CrossRef] [PubMed]
  116. Syed, Y.Y. Nedosiran: First Approval. Drugs 2023, 83, 1729–1733. [Google Scholar] [CrossRef] [PubMed]
  117. Wiesenfarth, M.; Dorst, J.; Brenner, D.; Elmas, Z.; Parlak, Ö.; Uzelac, Z.; Kandler, K.; Mayer, K.; Weiland, U.; Herrmann, C.; et al. Effects of Tofersen Treatment in Patients with SOD1-ALS in a “Real-World” Setting—A 12-Month Multicenter Cohort Study from the German Early Access Program. EClinicalMedicine 2024, 69, 102495. [Google Scholar] [CrossRef] [PubMed]
  118. Hoy, S.M. Delandistrogene Moxeparvovec: First Approval. Drugs 2023, 83, 1323–1329. [Google Scholar] [CrossRef] [PubMed]
  119. D’Angelo, S.P.; Araujo, D.M.; Abdul Razak, A.R.; Agulnik, M.; Attia, S.; Blay, J.-Y.; Carrasco Garcia, I.; Charlson, J.A.; Choy, E.; Demetri, G.D.; et al. Afamitresgene Autoleucel for Advanced Synovial Sarcoma and Myxoid Round Cell Liposarcoma (SPEARHEAD-1): An International, Open-Label, Phase 2 Trial. Lancet 2024, 403, 1460–1471. [Google Scholar] [CrossRef] [PubMed]
  120. Cuker, A.; Kavakli, K.; Frenzel, L.; Wang, J.-D.; Astermark, J.; Cerqueira, M.H.; Iorio, A.; Katsarou-Fasouli, O.; Klamroth, R.; Shapiro, A.D.; et al. Gene Therapy with Fidanacogene Elaparvovec in Adults with Hemophilia B. N. Engl. J. Med. 2024, 391, 1108–1118. [Google Scholar] [CrossRef] [PubMed]
  121. Friedmann, T. A Brief History of Gene Therapy. Nat. Genet. 1992, 2, 93–98. [Google Scholar] [CrossRef] [PubMed]
  122. Wirth, T.; Parker, N.; Ylä-Herttuala, S. History of Gene Therapy. Gene 2013, 525, 162–169. [Google Scholar] [CrossRef] [PubMed]
  123. Landhuis, E. The Definition of Gene Therapy Has Changed. Nature 2021. [Google Scholar] [CrossRef] [PubMed]
  124. Muul, L.M.; Tuschong, L.M.; Soenen, S.L.; Jagadeesh, G.J.; Ramsey, W.J.; Long, Z.; Carter, C.S.; Garabedian, E.K.; Alleyne, M.; Brown, M.; et al. Persistence and Expression of the Adenosine Deaminase Gene for 12 Years and Immune Reaction to Gene Transfer Components: Long-Term Results of the First Clinical Gene Therapy Trial. Blood 2003, 101, 2563–2569. [Google Scholar] [CrossRef] [PubMed]
  125. Caplen, N.J.; Alton, E.W.F.W.; Mddleton, P.G.; Dorin, J.R.; Stevenson, B.J.; Gao, X.; Durham, S.R.; Jeffery, P.K.; Hodson, M.E.; Coutelle, C.; et al. Liposome-Mediated CFTR Gene Transfer to the Nasal Epithelium of Patients with Cystic Fibrosis. Nat. Med. 1995, 1, 39–46. [Google Scholar] [CrossRef] [PubMed]
  126. Zabner, J.; Cheng, S.H.; Meeker, D.; Launspach, J.; Balfour, R.; Perricone, M.A.; Morris, J.E.; Marshall, J.; Fasbender, A.; Smith, A.E.; et al. Comparison of DNA-Lipid Complexes and DNA Alone for Gene Transfer to Cystic Fibrosis Airway Epithelia In Vivo. J. Clin. Investig. 1997, 100, 1529–1537. [Google Scholar] [CrossRef] [PubMed]
  127. Crystal, R.G.; McElvaney, N.G.; Rosenfeld, M.A.; Chu, C.S.; Mastrangeli, A.; Hay, J.G.; Brody, S.L.; Jaffe, H.A.; Eissa, N.T.; Danel, C. Administration of an Adenovirus Containing the Human CFTR CDNA to the Respiratory Tract of Individuals with Cystic Fibrosis. Nat. Genet. 1994, 8, 42–51. [Google Scholar] [CrossRef] [PubMed]
  128. Bellon, G.; Michel-Calemard, L.; Thouvenot, D.; Jagneaux, V.; Poitevin, F.; Malcus, C.; Accart, N.; Layani, M.P.; Aymard, M.; Bernon, H.; et al. Aerosol Administration of a Recombinant Adenovirus Expressing CFTR to Cystic Fibrosis Patients: A Phase I Clinical Trial. Hum. Gene Ther. 1997, 8, 15–25. [Google Scholar] [CrossRef] [PubMed]
  129. Flotte, T.; Carter, B.; Conrad, C.; Guggino, W.; Reynolds, T.; Rosenstein, B.; Taylor, G.; Walden, S.; Wetzel, R. A Phase I Study of an Adeno-Associated Virus-CFTR Gene Vector in Adult CF Patients with Mild Lung Disease. Johns Hopkins Children’s Center, Baltimore, Maryland. Hum. Gene Ther. 1996, 7, 1145–1159. [Google Scholar] [CrossRef] [PubMed]
  130. Virella-Lowell, I.; Poirier, A.; Chesnut, K.A.; Brantly, M.; Flotte, T.R. Inhibition of Recombinant Adeno-Associated Virus (RAAV) Transduction by Bronchial Secretions from Cystic Fibrosis Patients. Gene Ther. 2000, 7, 1783–1789. [Google Scholar] [CrossRef] [PubMed]
  131. Fasbender, A.; Zabner, J.; Zeiher, B.; Welsh, M. A Low Rate of Cell Proliferation and Reduced DNA Uptake Limit Cationic Lipid-Mediated Gene Transfer to Primary Cultures of Ciliated Human Airway Epithelia. Gene Ther. 1997, 4, 1173–1180. [Google Scholar] [CrossRef] [PubMed]
  132. Knowles, M.R.; Hohneker, K.W.; Zhou, Z.; Olsen, J.C.; Noah, T.L.; Hu, P.-C.; Leigh, M.W.; Engelhardt, J.F.; Edwards, L.J.; Jones, K.R.; et al. A Controlled Study of Adenoviral-Vector–Mediated Gene Transfer in the Nasal Epithelium of Patients with Cystic Fibrosis. N. Engl. J. Med. 1995, 333, 823–831. [Google Scholar] [CrossRef] [PubMed]
  133. Raper, S.E.; Chirmule, N.; Lee, F.S.; Wivel, N.A.; Bagg, A.; Gao, G.; Wilson, J.M.; Batshaw, M.L. Fatal Systemic Inflammatory Response Syndrome in a Ornithine Transcarbamylase Deficient Patient Following Adenoviral Gene Transfer. Mol. Genet. Metab. 2003, 80, 148–158. [Google Scholar] [CrossRef] [PubMed]
  134. Hacein-Bey-Abina, S.; von Kalle, C.; Schmidt, M.; Le Deist, F.; Wulffraat, N.; McIntyre, E.; Radford, I.; Villeval, J.-L.; Fraser, C.C.; Cavazzana-Calvo, M.; et al. A Serious Adverse Event after Successful Gene Therapy for X-Linked Severe Combined Immunodeficiency. N. Engl. J. Med. 2003, 348, 255–256. [Google Scholar] [CrossRef] [PubMed]
  135. Hacein-Bey-Abina, S.; Von Kalle, C.; Schmidt, M.; McCormack, M.P.; Wulffraat, N.; Leboulch, P.; Lim, A.; Osborne, C.S.; Pawliuk, R.; Morillon, E.; et al. LMO2-Associated Clonal T Cell Proliferation in Two Patients after Gene Therapy for SCID-X1. Science 2003, 302, 415–419. [Google Scholar] [CrossRef] [PubMed]
  136. Hacein-Bey-Abina, S.; Garrigue, A.; Wang, G.P.; Soulier, J.; Lim, A.; Morillon, E.; Clappier, E.; Caccavelli, L.; Delabesse, E.; Beldjord, K.; et al. Insertional Oncogenesis in 4 Patients after Retrovirus-Mediated Gene Therapy of SCID-X1. J. Clin. Investig. 2008, 118, 3132–3142. [Google Scholar] [CrossRef] [PubMed]
  137. Edelstein, M.L.; Abedi, M.R.; Wixon, J. Gene Therapy Clinical Trials Worldwide to 2007—An Update. J. Gene Med. 2007, 9, 833–842. [Google Scholar] [CrossRef] [PubMed]
  138. Friedmann, T. The Road toward Human Gene Therapy-A 25-Year Perspective. Ann. Med. 1997, 29, 575–577. [Google Scholar] [CrossRef] [PubMed]
  139. Dunbar, C.E.; High, K.A.; Joung, J.K.; Kohn, D.B.; Ozawa, K.; Sadelain, M. Gene Therapy Comes of Age. Science 2018, 359, eaan4672. [Google Scholar] [CrossRef] [PubMed]
  140. Dropulic, B.; Humeau, L.; Slepushkin, V.; Lu, X.; Manilla, P.; Rebello, T.; Afable, C.; Lacy, K.; Ybarra, C.; Levine, B.; et al. Establishing Safety in the Clinic for the First Lentiviral Vector To Be Tested in Humans. Blood 2004, 104, 411. [Google Scholar] [CrossRef]
  141. Cattoglio, C.; Pellin, D.; Rizzi, E.; Maruggi, G.; Corti, G.; Miselli, F.; Sartori, D.; Guffanti, A.; Di Serio, C.; Ambrosi, A.; et al. High-Definition Mapping of Retroviral Integration Sites Identifies Active Regulatory Elements in Human Multipotent Hematopoietic Progenitors. Blood 2010, 116, 5507–5517. [Google Scholar] [CrossRef] [PubMed]
  142. Moiani, A.; Suerth, J.D.; Gandolfi, F.; Rizzi, E.; Severgnini, M.; De Bellis, G.; Schambach, A.; Mavilio, F. Genome-Wide Analysis of Alpharetroviral Integration in Human Hematopoietic Stem/Progenitor Cells. Genes 2014, 5, 415–429. [Google Scholar] [CrossRef] [PubMed]
  143. Goldfarb, D.S. HIV-1 Virology. Simply Marvelous Nuclear Transport. Curr. Biol. 1995, 5, 570–573. [Google Scholar] [CrossRef] [PubMed]
  144. Gallay, P.; Hope, T.; Chin, D.; Trono, D. HIV-1 Infection of Nondividing Cells through the Recognition of Integrase by the Importin/Karyopherin Pathway. Proc. Natl. Acad. Sci. USA 1997, 94, 9825–9830. [Google Scholar] [CrossRef] [PubMed]
  145. Johnson, N.M.; Alvarado, A.F.; Moffatt, T.N.; Edavettal, J.M.; Swaminathan, T.A.; Braun, S.E. HIV-Based Lentiviral Vectors: Origin and Sequence Differences. Mol. Ther. Methods Clin. Dev. 2021, 21, 451–465. [Google Scholar] [CrossRef] [PubMed]
  146. Braendstrup, P.; Levine, B.L.; Ruella, M. The Long Road to the First FDA-Approved Gene Therapy: Chimeric Antigen Receptor T Cells Targeting CD19. Cytotherapy 2020, 22, 57–69. [Google Scholar] [CrossRef] [PubMed]
  147. Davies, J.C.; Polineni, D.; Boyd, A.C.; Donaldson, S.; Gill, D.R.; Griesenbach, U.; Hyde, S.C.; Jain, R.; McLachlan, G.; Mall, M.A.; et al. Lentiviral Gene Therapy for Cystic Fibrosis: A Promising Approach and First-in-Human Trial. Am. J. Respir. Crit. Care Med. 2024, 210, 1398–1408. [Google Scholar] [CrossRef] [PubMed]
  148. Kobayashi, M.; Iida, A.; Ueda, Y.; Hasegawa, M. Pseudotyped Lentivirus Vectors Derived from Simian Immunodeficiency Virus SIVagm with Envelope Glycoproteins from Paramyxovirus. J. Virol. 2003, 77, 2607–2614. [Google Scholar] [CrossRef] [PubMed]
  149. Munday, R.J.; Coradin, T.; Nimmo, R.; Lad, Y.; Hyde, S.C.; Mitrophanos, K.; Gill, D.R. Sendai F/HN Pseudotyped Lentiviral Vector Transduces Human Ciliated and Non-Ciliated Airway Cells Using α 2,3 Sialylated Receptors. Mol. Ther.—Methods Clin. Dev. 2022, 26, 239–252. [Google Scholar] [CrossRef] [PubMed]
  150. Wang, J.-H.; Gessler, D.J.; Zhan, W.; Gallagher, T.L.; Gao, G. Adeno-Associated Virus as a Delivery Vector for Gene Therapy of Human Diseases. Signal Transduct. Target. Ther. 2024, 9, 78. [Google Scholar] [CrossRef] [PubMed]
  151. Azad, R.F.; Driver, V.B.; Tanaka, K.; Crooke, R.M.; Anderson, K.P. Antiviral Activity of a Phosphorothioate Oligonucleotide Complementary to RNA of the Human Cytomegalovirus Major Immediate-Early Region. Antimicrob. Agents Chemother. 1993, 37, 1945–1954. [Google Scholar] [CrossRef] [PubMed]
  152. Hu, B.; Zhong, L.; Weng, Y.; Peng, L.; Huang, Y.; Zhao, Y.; Liang, X.-J. Therapeutic SiRNA: State of the Art. Signal Transduct. Target. Ther. 2020, 5, 101. [Google Scholar] [CrossRef] [PubMed]
  153. Lauffer, M.C.; van Roon-Mom, W.; Aartsma-Rus, A. Possibilities and Limitations of Antisense Oligonucleotide Therapies for the Treatment of Monogenic Disorders. Commun. Med. 2024, 4, 6. [Google Scholar] [CrossRef] [PubMed]
  154. Watts, J.K.; Corey, D.R. Silencing Disease Genes in the Laboratory and the Clinic. J. Pathol. 2012, 226, 365–379. [Google Scholar] [CrossRef] [PubMed]
  155. Liou, S.; Boggavarapu, R.; Cohen, N.R.; Zhang, Y.; Sharma, I.; Zeheb, L.; Mukund Acharekar, N.; Rodgers, H.D.; Islam, S.; Pitts, J.; et al. Structure of Anellovirus-like Particles Reveal a Mechanism for Immune Evasion. Nat. Commun. 2024, 15, 7219. [Google Scholar] [CrossRef] [PubMed]
  156. Tebas, P.; Stein, D.; Tang, W.W.; Frank, I.; Wang, S.Q.; Lee, G.; Spratt, S.K.; Surosky, R.T.; Giedlin, M.A.; Nichol, G.; et al. Gene Editing of CCR5 in Autologous CD4 T Cells of Persons Infected with HIV. N. Engl. J. Med. 2014, 370, 901–910. [Google Scholar] [CrossRef] [PubMed]
  157. Ding, Q.; Lee, Y.-K.; Schaefer, E.A.K.; Peters, D.T.; Veres, A.; Kim, K.; Kuperwasser, N.; Motola, D.L.; Meissner, T.B.; Hendriks, W.T.; et al. A TALEN Genome-Editing System for Generating Human Stem Cell-Based Disease Models. Cell Stem Cell 2013, 12, 238–251. [Google Scholar] [CrossRef] [PubMed]
  158. Qasim, W.; Amrolia, P.J.; Samarasinghe, S.; Ghorashian, S.; Zhan, H.; Stafford, S.; Butler, K.; Ahsan, G.; Gilmour, K.; Adams, S.; et al. First Clinical Application of Talen Engineered Universal CAR19 T Cells in B-ALL. Blood 2015, 126, 2046. [Google Scholar] [CrossRef]
  159. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome Engineering Using the CRISPR-Cas9 System. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef] [PubMed]
  160. Komor, A.C.; Kim, Y.B.; Packer, M.S.; Zuris, J.A.; Liu, D.R. Programmable Editing of a Target Base in Genomic DNA without Double-Stranded DNA Cleavage. Nature 2016, 533, 420–424. [Google Scholar] [CrossRef] [PubMed]
  161. Anzalone, A.V.; Randolph, P.B.; Davis, J.R.; Sousa, A.A.; Koblan, L.W.; Levy, J.M.; Chen, P.J.; Wilson, C.; Newby, G.A.; Raguram, A.; et al. Search-and-Replace Genome Editing without Double-Strand Breaks or Donor DNA. Nature 2019, 576, 149–157. [Google Scholar] [CrossRef] [PubMed]
  162. Ashmore-Harris, C.; Fruhwirth, G.O. The Clinical Potential of Gene Editing as a Tool to Engineer Cell-Based Therapeutics. Clin. Transl. Med. 2020, 9, 15. [Google Scholar] [CrossRef] [PubMed]
  163. Lu, Y.; Xue, J.; Deng, T.; Zhou, X.; Yu, K.; Deng, L.; Huang, M.; Yi, X.; Liang, M.; Wang, Y.; et al. Safety and Feasibility of CRISPR-Edited T Cells in Patients with Refractory Non-Small-Cell Lung Cancer. Nat. Med. 2020, 26, 732–740. [Google Scholar] [CrossRef] [PubMed]
  164. Philippidis, A. CASGEVY Makes History as FDA Approves First CRISPR/Cas9 Genome Edited Therapy. Hum. Gene Ther. 2024, 35, 1–4. [Google Scholar] [CrossRef] [PubMed]
  165. Lino, C.A.; Harper, J.C.; Carney, J.P.; Timlin, J.A. Delivering CRISPR: A Review of the Challenges and Approaches. Drug Deliv. 2018, 25, 1234–1257. [Google Scholar] [CrossRef] [PubMed]
  166. Choi, J.G.; Dang, Y.; Abraham, S.; Ma, H.; Zhang, J.; Guo, H.; Cai, Y.; Mikkelsen, J.G.; Wu, H.; Shankar, P.; et al. Lentivirus Pre-Packed with Cas9 Protein for Safer Gene Editing. Gene Ther. 2016, 23, 627–633. [Google Scholar] [CrossRef] [PubMed]
  167. Uchida, N.; Drysdale, C.M.; Nassehi, T.; Gamer, J.; Yapundich, M.; DiNicola, J.; Shibata, Y.; Hinds, M.; Gudmundsdottir, B.; Haro-Mora, J.J.; et al. Cas9 Protein Delivery Non-Integrating Lentiviral Vectors for Gene Correction in Sickle Cell Disease. Mol. Ther. Methods Clin. Dev. 2021, 21, 121–132. [Google Scholar] [CrossRef] [PubMed]
  168. Sanjana, N.E.; Shalem, O.; Zhang, F. Improved Vectors and Genome-Wide Libraries for CRISPR Screening. Nat. Methods 2014, 11, 783–784. [Google Scholar] [CrossRef] [PubMed]
  169. Cai, Y.; Bak, R.O.; Krogh, L.B.; Staunstrup, N.H.; Moldt, B.; Corydon, T.J.; Schrøder, L.D.; Mikkelsen, J.G. DNA Transposition by Protein Transduction of the PiggyBac Transposase from Lentiviral Gag Precursors. Nucleic Acids Res. 2014, 42, e28. [Google Scholar] [CrossRef] [PubMed]
  170. Cai, Y.; Bak, R.O.; Mikkelsen, J.G. Targeted Genome Editing by Lentiviral Protein Transduction of Zinc-Finger and TAL-Effector Nucleases. eLife 2014, 3, e01911. [Google Scholar] [CrossRef] [PubMed]
  171. Strebinger, D.; Frangieh, C.J.; Friedrich, M.J.; Faure, G.; Macrae, R.K.; Zhang, F. Cell Type-Specific Delivery by Modular Envelope Design. Nat. Commun. 2023, 14, 5141. [Google Scholar] [CrossRef] [PubMed]
  172. Verma, I.M.; Weitzman, M.D. GENE THERAPY: Twenty-First Century Medicine. Annu. Rev. Biochem. 2005, 74, 711–738. [Google Scholar] [CrossRef] [PubMed]
  173. Meng, X.; Du, Y.; Liu, C.; Zhai, Z.; Pan, J. GTO: A Comprehensive Gene Therapy Omnibus. Nucleic Acids Res. 2025, 53, D1393–D1403. [Google Scholar] [CrossRef] [PubMed]
  174. Berg, P.; Ruppert-Seipp, G.; Müller, S.; Maurer, G.D.; Hartmann, J.; Holtick, U.; Buchholz, C.J.; Funk, M.B. CAR T-Cell-Associated Secondary Malignancies Challenge Current Pharmacovigilance Concepts. EMBO Mol. Med. 2025, 17, 211–218. [Google Scholar] [CrossRef] [PubMed]
  175. Willyard, C. Do Cutting-Edge CAR-T-Cell Therapies Cause Cancer? What the Data Say. Nature 2024, 629, 22–24. [Google Scholar] [CrossRef] [PubMed]
  176. Dolgin, E. FDA Investigating CAR-Related T-Cell Malignancies. Cancer Discov. 2024, 14, 9–10. [Google Scholar] [CrossRef]
  177. Duncan, C.N.; Bledsoe, J.R.; Grzywacz, B.; Beckman, A.; Bonner, M.; Eichler, F.S.; Kühl, J.-S.; Harris, M.H.; Slauson, S.; Colvin, R.A.; et al. Hematologic Cancer after Gene Therapy for Cerebral Adrenoleukodystrophy. N. Engl. J. Med. 2024, 391, 1287–1301. [Google Scholar] [CrossRef] [PubMed]
  178. Kohn, D.B.; Booth, C.; Naldini, L. Myelodysplasia after Lentiviral Gene Therapy. N. Engl. J. Med. 2024, 391, 2382. [Google Scholar] [CrossRef] [PubMed]
  179. Halene, S.; Wang, L.; Cooper, R.M.; Bockstoce, D.C.; Robbins, P.B.; Kohn, D.B. Improved Expression in Hematopoietic and Lymphoid Cells in Mice after Transplantation of Bone Marrow Transduced with a Modified Retroviral Vector. Blood 1999, 94, 3349–3357. [Google Scholar] [CrossRef] [PubMed]
  180. Montini, E.; Cesana, D.; Schmidt, M.; Sanvito, F.; Bartholomae, C.C.; Ranzani, M.; Benedicenti, F.; Sergi, L.S.; Ambrosi, A.; Ponzoni, M.; et al. The Genotoxic Potential of Retroviral Vectors Is Strongly Modulated by Vector Design and Integration Site Selection in a Mouse Model of HSC Gene Therapy. J. Clin. Investig. 2009, 119, 964–975. [Google Scholar] [CrossRef] [PubMed]
  181. Espinoza, D.A.; Fan, X.; Yang, D.; Cordes, S.F.; Truitt, L.L.; Calvo, K.R.; Yabe, I.M.; Demirci, S.; Hope, K.J.; Hong, S.G.; et al. Aberrant Clonal Hematopoiesis Following Lentiviral Vector Transduction of HSPCs in a Rhesus Macaque. Mol. Ther. 2019, 27, 1074–1086. [Google Scholar] [CrossRef] [PubMed]
  182. Kim, C.Y.; Zhang, X.; Witola, W.H. Small GTPase Immunity-Associated Proteins Mediate Resistance to Toxoplasma Gondii Infection in Lewis Rat. Infect. Immun. 2018, 86, 10-1128. [Google Scholar] [CrossRef] [PubMed]
  183. Johnson, S.; Karpova, Y.; Guo, D.; Ghatak, A.; Markov, D.A.; Tulin, A.V. PARG Suppresses Tumorigenesis and Downregulates Genes Controlling Angiogenesis, Inflammatory Response, and Immune Cell Recruitment. BMC Cancer 2022, 22, 557. [Google Scholar] [CrossRef] [PubMed]
  184. Alvandi, Z.; Opas, M. C-Src Kinase Inhibits Osteogenic Differentiation via Enhancing STAT1 Stability. PLoS ONE 2020, 15, e0241646. [Google Scholar] [CrossRef] [PubMed]
  185. Zhang, X.; Liu, Y.; Liu, J.; Bailey, A.L.; Plante, K.S.; Plante, J.A.; Zou, J.; Xia, H.; Bopp, N.E.; Aguilar, P.V.; et al. A Trans-Complementation System for SARS-CoV-2 Recapitulates Authentic Viral Replication without Virulence. Cell 2021, 184, 2229–2238.e13. [Google Scholar] [CrossRef] [PubMed]
  186. Gossen, M.; Bujard, H. Tight Control of Gene Expression in Mammalian Cells by Tetracycline-Responsive Promoters. Proc. Natl. Acad. Sci. USA 1992, 89, 5547–5551. [Google Scholar] [CrossRef] [PubMed]
  187. Wu, X.; Wakefield, J.K.; Liu, H.; Xiao, H.; Kralovics, R.; Prchal, J.T.; Kappes, J.C. Development of a Novel Trans-Lentiviral Vector That Affords Predictable Safety. Mol. Ther. 2000, 2, 47–55. [Google Scholar] [CrossRef] [PubMed]
  188. Vink, C.A.; Counsell, J.R.; Perocheau, D.P.; Karda, R.; Buckley, S.M.K.; Brugman, M.H.; Galla, M.; Schambach, A.; McKay, T.R.; Waddington, S.N.; et al. Eliminating HIV-1 Packaging Sequences from Lentiviral Vector Proviruses Enhances Safety and Expedites Gene Transfer for Gene Therapy. Mol. Ther. 2017, 25, 1790–1804. [Google Scholar] [CrossRef] [PubMed]
  189. Berkhout, B. A Fourth Generation Lentiviral Vector: Simplifying Genomic Gymnastics. Mol. Ther. 2017, 25, 1741–1743. [Google Scholar] [CrossRef] [PubMed]
  190. Logan, A.C.; Haas, D.L.; Kafri, T.; Kohn, D.B. Integrated Self-Inactivating Lentiviral Vectors Produce Full-Length Genomic Transcripts Competent for Encapsidation and Integration. J. Virol. 2004, 78, 8421–8436. [Google Scholar] [CrossRef] [PubMed]
  191. Wright, J.; Alberts, B.; Mitrophanous, K.; Clarkson, N.; Farley, D.; Kazarian, K.; Hove, S. The TetravectaTM System: A New Tool Kit Enhancing Lentiviral Vector Production for the Next Generation of Gene Therapies. Cytotherapy 2024, 26, S208. [Google Scholar] [CrossRef]
  192. Wright, J.; Alberts, B.M.; Hood, A.J.; Nogueira, C.; Miskolczi, Z.; Vieira, C.R.; Chipchase, D.; Lamont, C.M.; Goodyear, O.; Moyce, L.J.; et al. Improved Production and Quality of Lentiviral Vectors By Major-Splice-Donor Mutation and Co-Expression of a Novel U1 Snrna-Based Enhancer. SSRN, 2024; preprint. [Google Scholar] [CrossRef]
  193. Mueller, N.; van Bel, N.; Berkhout, B.; Das, A.T. HIV-1 Splicing at the Major Splice Donor Site Is Restricted by RNA Structure. Virology 2014, 468–470, 609–620. [Google Scholar] [CrossRef] [PubMed]
  194. Emery, A.; Swanstrom, R. HIV-1: To Splice or Not to Splice, That Is the Question. Viruses 2021, 13, 181. [Google Scholar] [CrossRef] [PubMed]
  195. Sertkaya, H.; Ficarelli, M.; Sweeney, N.P.; Parker, H.; Vink, C.A.; Swanson, C.M. HIV-1 Sequences in Lentiviral Vector Genomes Can Be Substantially Reduced without Compromising Transduction Efficiency. Sci. Rep. 2021, 11, 12067. [Google Scholar] [CrossRef] [PubMed]
  196. Antson, A.A.; Dodson, E.J.; Dodson, G.; Greaves, R.B.; Chen, X.; Gollnick, P. Structure of the Trp RNA-Binding Attenuation Protein, TRAP, Bound to RNA. Nature 1999, 401, 235–242. [Google Scholar] [CrossRef] [PubMed]
  197. Maunder, H.E.; Wright, J.; Kolli, B.R.; Vieira, C.R.; Mkandawire, T.T.; Tatoris, S.; Kennedy, V.; Iqball, S.; Devarajan, G.; Ellis, S.; et al. Enhancing Titres of Therapeutic Viral Vectors Using the Transgene Repression in Vector Production (TRiP) System. Nat. Commun. 2017, 8, 14834. [Google Scholar] [CrossRef] [PubMed]
  198. German Advisory Committee Blood (Arbeitskreis Blut). Subgroup ‘Assessment of Pathogens Transmissible by Blood’. Human Immunodeficiency Virus (HIV). Transfus. Med. Hemother. 2016, 43, 203–222. [Google Scholar] [CrossRef] [PubMed]
  199. Tabler, C.O.; Wegman, S.J.; Chen, J.; Shroff, H.; Alhusaini, N.; Tilton, J.C. The HIV-1 Viral Protease Is Activated during Assembly and Budding Prior to Particle Release. J. Virol. 2022, 96, e02198-21. [Google Scholar] [CrossRef] [PubMed]
  200. Itano, M.S.; Arnion, H.; Wolin, S.L.; Simon, S.M. Recruitment of 7SL RNA to Assembling HIV-1 Virus-like Particles. Traffic 2018, 19, 36–43. [Google Scholar] [CrossRef] [PubMed]
  201. Onafuwa-Nuga, A.A.; Telesnitsky, A.; King, S.R. 7SL RNA, but Not the 54-Kd Signal Recognition Particle Protein, Is an Abundant Component of Both Infectious HIV-1 and Minimal Virus-like Particles. RNA 2006, 12, 542–546. [Google Scholar] [CrossRef] [PubMed]
  202. Huang, Y.; Mak, J.; Cao, Q.; Li, Z.; Wainberg, M.A.; Kleiman, L. Incorporation of Excess Wild-Type and Mutant TRNA(3Lys) into Human Immunodeficiency Virus Type 1. J. Virol. 1994, 68, 7676–7683. [Google Scholar] [CrossRef] [PubMed]
  203. Kleiman, L.; Jones, C.P.; Musier-Forsyth, K. Formation of the TRNA Lys Packaging Complex in HIV-1. FEBS Lett. 2010, 584, 359–365. [Google Scholar] [CrossRef] [PubMed]
  204. Franke, E.K.; Yuan, H.E.; Luban, J. Specific Incorporation of Cyclophilin A into HIV-1 Virions. Nature 1994, 372, 359–362. [Google Scholar] [CrossRef] [PubMed]
  205. Ott, D.E.; Coren, L.V.; Kane, B.P.; Busch, L.K.; Johnson, D.G.; Sowder, R.C.; Chertova, E.N.; Arthur, L.O.; Henderson, L.E. Cytoskeletal Proteins inside Human Immunodeficiency Virus Type 1 Virions. J. Virol. 1996, 70, 7734–7743. [Google Scholar] [CrossRef] [PubMed]
  206. Müller, B.; Tessmer, U.; Schubert, U.; Kräusslich, H.G. Human Immunodeficiency Virus Type 1 Vpr Protein Is Incorporated into the Virion in Significantly Smaller Amounts than Gag and Is Phosphorylated in Infected Cells. J. Virol. 2000, 74, 9727–9731. [Google Scholar] [CrossRef] [PubMed]
  207. Linde, M.E.; Colquhoun, D.R.; Ubaida Mohien, C.; Kole, T.; Aquino, V.; Cotter, R.; Edwards, N.; Hildreth, J.E.K.; Graham, D.R. The Conserved Set of Host Proteins Incorporated into HIV-1 Virions Suggests a Common Egress Pathway in Multiple Cell Types. J. Proteome Res. 2013, 12, 2045–2054. [Google Scholar] [CrossRef] [PubMed]
  208. Kaushik, R.; Ratner, L. Role of Human Immunodeficiency Virus Type 1 Matrix Phosphorylation in an Early Postentry Step of Virus Replication. J. Virol. 2004, 78, 2319–2326. [Google Scholar] [CrossRef] [PubMed]
  209. Burnie, J.; Guzzo, C. The Incorporation of Host Proteins into the External HIV-1 Envelope. Viruses 2019, 11, 85. [Google Scholar] [CrossRef] [PubMed]
  210. Guzzo, C.; Ichikawa, D.; Park, C.; Phillips, D.; Liu, Q.; Zhang, P.; Kwon, A.; Miao, H.; Lu, J.; Rehm, C.; et al. Virion Incorporation of Integrin A4β7 Facilitates HIV-1 Infection and Intestinal Homing. Sci. Immunol. 2017, 2, eaam7341. [Google Scholar] [CrossRef] [PubMed]
  211. Chen, B. Molecular Mechanism of HIV-1 Entry. Trends Microbiol. 2019, 27, 878–891. [Google Scholar] [CrossRef] [PubMed]
  212. Melikyan, G.B. HIV Entry: A Game of Hide-and-Fuse? Curr. Opin. Virol. 2014, 4, 1–7. [Google Scholar] [CrossRef] [PubMed]
  213. Miyauchi, K.; Kim, Y.; Latinovic, O.; Morozov, V.; Melikyan, G.B. HIV Enters Cells via Endocytosis and Dynamin-Dependent Fusion with Endosomes. Cell 2009, 137, 433–444. [Google Scholar] [CrossRef] [PubMed]
  214. Engelman, A.N.; Singh, P.K. Cellular and Molecular Mechanisms of HIV-1 Integration Targeting. Cell. Mol. Life Sci. 2018, 75, 2491–2507. [Google Scholar] [CrossRef] [PubMed]
  215. Cronin, J.; Zhang, X.-Y.; Reiser, J. Altering the Tropism of Lentiviral Vectors through Pseudotyping. Curr. Gene Ther. 2005, 5, 387–398. [Google Scholar] [CrossRef] [PubMed]
  216. Duvergé, A.; Negroni, M. Pseudotyping Lentiviral Vectors: When the Clothes Make the Virus. Viruses 2020, 12, 1311. [Google Scholar] [CrossRef] [PubMed]
  217. Shinohara, T.; Kanatsu-Shinohara, M. Transgenesis and Genome Editing of Mouse Spermatogonial Stem Cells by Lentivirus Pseudotyped with Sendai Virus F Protein. Stem Cell Rep. 2020, 14, 447–461. [Google Scholar] [CrossRef] [PubMed]
  218. Jargalsaikhan, B.-E.; Muto, M.; Been, Y.; Matsumoto, S.; Okamura, E.; Takahashi, T.; Narimichi, Y.; Kurebayashi, Y.; Takeuchi, H.; Shinohara, T.; et al. The Dual-Pseudotyped Lentiviral Vector with VSV-G and Sendai Virus HN Enhances Infection Efficiency through the Synergistic Effect of the Envelope Proteins. Viruses 2024, 16, 827. [Google Scholar] [CrossRef] [PubMed]
  219. Dautzenberg, I.J.C.; Rabelink, M.J.W.E.; Hoeben, R.C. The Stability of Envelope-Pseudotyped Lentiviral Vectors. Gene Ther. 2021, 28, 89–104. [Google Scholar] [CrossRef] [PubMed]
  220. Haid, S.; Grethe, C.; Bankwitz, D.; Grunwald, T.; Pietschmann, T. Identification of a Human Respiratory Syncytial Virus Cell Entry Inhibitor by Using a Novel Lentiviral Pseudotype System. J. Virol. 2015, 90, 3065–3073. [Google Scholar] [CrossRef] [PubMed]
  221. Frecha, C.; Costa, C.; Nègre, D.; Gauthier, E.; Russell, S.J.; Cosset, F.-L.; Verhoeyen, E. Stable Transduction of Quiescent T Cells without Induction of Cycle Progression by a Novel Lentiviral Vector Pseudotyped with Measles Virus Glycoproteins. Blood 2008, 112, 4843–4852. [Google Scholar] [CrossRef] [PubMed]
  222. Chai, N.; Chang, H.E.; Nicolas, E.; Gudima, S.; Chang, J.; Taylor, J. Assembly of Hepatitis B Virus Envelope Proteins onto a Lentivirus Pseudotype That Infects Primary Human Hepatocytes. J. Virol. 2007, 81, 10897–10904. [Google Scholar] [CrossRef] [PubMed]
  223. Hatziioannou, T.; Russell, S.J.; Cosset, F.L. Incorporation of Simian Virus 5 Fusion Protein into Murine Leukemia Virus Particles and Its Effect on the Co-Incorporation of Retroviral Envelope Glycoproteins. Virology 2000, 267, 49–57. [Google Scholar] [CrossRef] [PubMed]
  224. Finkelshtein, D.; Werman, A.; Novick, D.; Barak, S.; Rubinstein, M. LDL Receptor and Its Family Members Serve as the Cellular Receptors for Vesicular Stomatitis Virus. Proc. Natl. Acad. Sci. USA 2013, 110, 7306–7311. [Google Scholar] [CrossRef] [PubMed]
  225. Bour, S.; Geleziunas, R.; Wainberg, M.A. The Human Immunodeficiency Virus Type 1 (HIV-1) CD4 Receptor and Its Central Role in Promotion of HIV-1 Infection. Microbiol. Rev. 1995, 59, 63–93. [Google Scholar] [CrossRef] [PubMed]
  226. Stitz, J.; Buchholz, C.J.; Engelstädter, M.; Uckert, W.; Bloemer, U.; Schmitt, I.; Cichutek, K. Lentiviral Vectors Pseudotyped with Envelope Glycoproteins Derived from Gibbon Ape Leukemia Virus and Murine Leukemia Virus 10A1. Virology 2000, 273, 16–20. [Google Scholar] [CrossRef] [PubMed]
  227. Schambach, A.; Galla, M.; Modlich, U.; Will, E.; Chandra, S.; Reeves, L.; Colbert, M.; Williams, D.A.; von Kalle, C.; Baum, C. Lentiviral Vectors Pseudotyped with Murine Ecotropic Envelope: Increased Biosafety and Convenience in Preclinical Research. Exp. Hematol. 2006, 34, 588–592. [Google Scholar] [CrossRef] [PubMed]
  228. Tomás, H.A.; Mestre, D.A.; Rodrigues, A.F.; Guerreiro, M.R.; Carrondo, M.J.T.; Coroadinha, A.S. Improved GaLV-TR Glycoproteins to Pseudotype Lentiviral Vectors: Impact of Viral Protease Activity in the Production of LV Pseudotypes. Mol. Ther. Methods Clin. Dev. 2019, 15, 1–8. [Google Scholar] [CrossRef] [PubMed]
  229. Watson, D.J.; Kobinger, G.P.; Passini, M.A.; Wilson, J.M.; Wolfe, J.H. Targeted Transduction Patterns in the Mouse Brain by Lentivirus Vectors Pseudotyped with VSV, Ebola, Mokola, LCMV, or MuLV Envelope Proteins. Mol. Ther. 2002, 5, 528–537. [Google Scholar] [CrossRef] [PubMed]
  230. Mirow, M.; Schwarze, L.I.; Fehse, B.; Riecken, K. Efficient Pseudotyping of Different Retroviral Vectors Using a Novel, Codon-Optimized Gene for Chimeric GALV Envelope. Viruses 2021, 13, 1471. [Google Scholar] [CrossRef] [PubMed]
  231. Sandrin, V.; Boson, B.; Salmon, P.; Gay, W.; Nègre, D.; Le Grand, R.; Trono, D.; Cosset, F.-L. Lentiviral Vectors Pseudotyped with a Modified RD114 Envelope Glycoprotein Show Increased Stability in Sera and Augmented Transduction of Primary Lymphocytes and CD34+ Cells Derived from Human and Nonhuman Primates. Blood 2002, 100, 823–832. [Google Scholar] [CrossRef] [PubMed]
  232. Matsuyama, M.; Ohashi, Y.; Tsubota, T.; Yaguchi, M.; Kato, S.; Kobayashi, K.; Miyashita, Y. Avian Sarcoma Leukosis Virus Receptor-Envelope System for Simultaneous Dissection of Multiple Neural Circuits in Mammalian Brain. Proc. Natl. Acad. Sci. USA 2015, 112, E2947–E2956. [Google Scholar] [CrossRef] [PubMed]
  233. Liu, S.-L.; Halbert, C.L.; Miller, A.D. Jaagsiekte Sheep Retrovirus Envelope Efficiently Pseudotypes Human Immunodeficiency Virus Type 1-Based Lentiviral Vectors. J. Virol. 2004, 78, 2642–2647. [Google Scholar] [CrossRef] [PubMed]
  234. Sun, Z.; Li, Y.; Li, J.; Cai, S.; Zeng, Y.; Lindemann, D.; Pollok, K.E.; Hanenberg, H.; Clapp, D.W. A Modified Foamy Viral Envelope Enhances Gene Transfer Efficiency and Reduces Toxicity of Lentiviral FANCA Vectors in FANCA-/-HSCs. Blood 2009, 114, 696. [Google Scholar] [CrossRef]
  235. Colamartino, A.B.L.L.; Lemieux, W.; Bifsha, P.; Nicoletti, S.; Chakravarti, N.; Sanz, J.; Roméro, H.; Selleri, S.; Béland, K.; Guiot, M.; et al. Efficient and Robust NK-Cell Transduction with Baboon Envelope Pseudotyped Lentivector. Front. Immunol. 2019, 10, 2873. [Google Scholar] [CrossRef] [PubMed]
  236. Renner, A.; Stahringer, A.; Ruppel, K.E.; Fricke, S.; Koehl, U.; Schmiedel, D. Development of KoRV-Pseudotyped Lentiviral Vectors for Efficient Gene Transfer into Freshly Isolated Immune Cells. Gene Ther. 2024, 31, 378–390. [Google Scholar] [CrossRef] [PubMed]
  237. Girard-Gagnepain, A.; Amirache, F.; Costa, C.; Lévy, C.; Frecha, C.; Fusil, F.; Nègre, D.; Lavillette, D.; Cosset, F.-L.; Verhoeyen, E. Baboon Envelope Pseudotyped LVs Outperform VSV-G-LVs for Gene Transfer into Early-Cytokine-Stimulated and Resting HSCs. Blood 2014, 124, 1221–1231. [Google Scholar] [CrossRef] [PubMed]
  238. Drakopoulou, E.; Georgomanoli, M.; Lederer, C.W.; Kleanthous, M.; Costa, C.; Bernadin, O.; Cosset, F.-L.; Voskaridou, E.; Verhoeyen, E.; Papanikolaou, E.; et al. A Novel BaEVRless-Pseudotyped γ-Globin Lentiviral Vector Drives High and Stable Fetal Hemoglobin Expression and Improves Thalassemic Erythropoiesis In Vitro. Hum. Gene Ther. 2019, 30, 601–617. [Google Scholar] [CrossRef] [PubMed]
  239. Klatt, D.; Sereni, L.; Liu, B.; Genovese, P.; Schambach, A.; Verhoeyen, E.; Williams, D.A.; Brendel, C. Engineered Packaging Cell Line for the Enhanced Production of Baboon-Enveloped Retroviral Vectors. Mol. Ther. Nucleic Acids 2024, 35, 102389. [Google Scholar] [CrossRef] [PubMed]
  240. DePolo, N.J.; Reed, J.D.; Sheridan, P.L.; Townsend, K.; Sauter, S.L.; Jolly, D.J.; Dubensky, T.W. VSV-G Pseudotyped Lentiviral Vector Particles Produced in Human Cells Are Inactivated by Human Serum. Mol. Ther. 2000, 2, 218–222. [Google Scholar] [CrossRef] [PubMed]
  241. Hislop, J.N.; Islam, T.A.; Eleftheriadou, I.; Carpentier, D.C.J.; Trabalza, A.; Parkinson, M.; Schiavo, G.; Mazarakis, N.D. Rabies Virus Envelope Glycoprotein Targets Lentiviral Vectors to the Axonal Retrograde Pathway in Motor Neurons. J. Biol. Chem. 2014, 289, 16148–16163. [Google Scholar] [CrossRef] [PubMed]
  242. Cannon, J.R.; Sew, T.; Montero, L.; Burton, E.A.; Greenamyre, J.T. Pseudotype-Dependent Lentiviral Transduction of Astrocytes or Neurons in the Rat Substantia Nigra. Exp. Neurol. 2011, 228, 41–52. [Google Scholar] [CrossRef] [PubMed]
  243. Trobridge, G.D.; Wu, R.A.; Hansen, M.; Ironside, C.; Watts, K.L.; Olsen, P.; Beard, B.C.; Kiem, H.-P. Cocal-Pseudotyped Lentiviral Vectors Resist Inactivation by Human Serum and Efficiently Transduce Primate Hematopoietic Repopulating Cells. Mol. Ther. 2010, 18, 725–733. [Google Scholar] [CrossRef] [PubMed]
  244. Hu, S.; Mohan Kumar, D.; Sax, C.; Schuler, C.; Akkina, R. Pseudotyping of Lentiviral Vector with Novel Vesiculovirus Envelope Glycoproteins Derived from Chandipura and Piry Viruses. Virology 2016, 488, 162–168. [Google Scholar] [CrossRef] [PubMed]
  245. Kumar, M.; Bradow, B.P.; Zimmerberg, J. Large-Scale Production of Pseudotyped Lentiviral Vectors Using Baculovirus GP64. Hum. Gene Ther. 2003, 14, 67–77. [Google Scholar] [CrossRef] [PubMed]
  246. Schauber, C.A.; Tuerk, M.J.; Pacheco, C.D.; Escarpe, P.A.; Veres, G. Lentiviral Vectors Pseudotyped with Baculovirus Gp64 Efficiently Transduce Mouse Cells In Vivo and Show Tropism Restriction against Hematopoietic Cell Types In Vitro. Gene Ther. 2004, 11, 266–275. [Google Scholar] [CrossRef] [PubMed]
  247. Milani, M.; Canepari, C.; Assanelli, S.; Merlin, S.; Borroni, E.; Starinieri, F.; Biffi, M.; Russo, F.; Fabiano, A.; Zambroni, D.; et al. GP64-Pseudotyped Lentiviral Vectors Target Liver Endothelial Cells and Correct Hemophilia A Mice. EMBO Mol. Med. 2024, 16, 1427–1450. [Google Scholar] [CrossRef] [PubMed]
  248. Kataoka, C.; Kaname, Y.; Taguwa, S.; Abe, T.; Fukuhara, T.; Tani, H.; Moriishi, K.; Matsuura, Y. Baculovirus GP64-Mediated Entry into Mammalian Cells. J. Virol. 2012, 86, 2610–2620. [Google Scholar] [CrossRef] [PubMed]
  249. Amirache, F.; Lévy, C.; Costa, C.; Mangeot, P.E.; Torbett, B.E.; Wang, C.X.; Nègre, D.; Cosset, F.L.; Verhoeyen, E. Mystery Solved: VSV-G-LVs Do Not Allow Efficient Gene Transfer into Unstimulated T Cells, B Cells, and HSCs Because They Lack the LDL Receptor. Blood 2014, 123, 1422–1424. [Google Scholar] [CrossRef] [PubMed]
  250. Humbert, J.-M.; Frecha, C.; Bouafia, F.A.; N’Guyen, T.H.; Boni, S.; Cosset, F.-L.; Verhoeyen, E.; Halary, F. Measles Virus Glycoprotein-Pseudotyped Lentiviral Vectors Are Highly Superior to Vesicular Stomatitis Virus G Pseudotypes for Genetic Modification of Monocyte-Derived Dendritic Cells. J. Virol. 2012, 86, 5192–5203. [Google Scholar] [CrossRef] [PubMed]
  251. Lévy, C.; Amirache, F.; Costa, C.; Frecha, C.; Muller, C.P.; Kweder, H.; Buckland, R.; Cosset, F.-L.; Verhoeyen, E. Lentiviral Vectors Displaying Modified Measles Virus Gp Overcome Pre-Existing Immunity in in Vivo-like Transduction of Human T and B Cells. Mol. Ther. 2012, 20, 1699–1712. [Google Scholar] [CrossRef] [PubMed]
  252. Lévy, C.; Amirache, F.; Girard-Gagnepain, A.; Frecha, C.; Roman-Rodríguez, F.J.; Bernadin, O.; Costa, C.; Nègre, D.; Gutierrez-Guerrero, A.; Vranckx, L.S.; et al. Measles Virus Envelope Pseudotyped Lentiviral Vectors Transduce Quiescent Human HSCs at an Efficiency without Precedent. Blood Adv. 2017, 1, 2088–2104. [Google Scholar] [CrossRef] [PubMed]
  253. Jung, C.; Grzybowski, B.N.; Tong, S.; Cheng, L.; Compans, R.W.; LeDoux, J.M. Lentiviral Vectors Pseudotyped with Envelope Glycoproteins Derived from Human Parainfluenza Virus Type 3. Biotechnol. Prog. 2004, 20, 1810–1816. [Google Scholar] [CrossRef] [PubMed]
  254. Jung, C.; Le Doux, J.M. Lentiviruses Inefficiently Incorporate Human Parainfluenza Type 3 Envelope Proteins. Biotechnol. Bioeng. 2008, 99, 1016–1027. [Google Scholar] [CrossRef] [PubMed]
  255. Khetawat, D.; Broder, C.C. A Functional Henipavirus Envelope Glycoprotein Pseudotyped Lentivirus Assay System. Virol. J. 2010, 7, 312. [Google Scholar] [CrossRef] [PubMed]
  256. Palomares, K.; Vigant, F.; Van Handel, B.; Pernet, O.; Chikere, K.; Hong, P.; Sherman, S.P.; Patterson, M.; An, D.S.; Lowry, W.E.; et al. Nipah Virus Envelope-Pseudotyped Lentiviruses Efficiently Target EphrinB2-Positive Stem Cell Populations In Vitro and Bypass the Liver Sink When Administered In Vivo. J. Virol. 2013, 87, 2094–2108. [Google Scholar] [CrossRef] [PubMed]
  257. Abdelrahman, D.N.; Abdullahi, F.L.; Abdu-Raheem, F.; Abicher, L.T.; Adelaiye, H.; Badjie, A.; Bah, A.; Bista, K.P.; Bont, L.J.; Boom, T.T.; et al. Respiratory Syncytial Virus Infection among Children Younger than 2 Years Admitted to a Paediatric Intensive Care Unit with Extended Severe Acute Respiratory Infection in Ten Gavi-Eligible Countries: The RSV GOLD—ICU Network Study. Lancet Glob. Health 2024, 12, e1611–e1619. [Google Scholar] [CrossRef] [PubMed]
  258. Biuso, F.; Palladino, L.; Manenti, A.; Stanzani, V.; Lapini, G.; Gao, J.; Couzens, L.; Eichelberger, M.C.; Montomoli, E. Use of Lentiviral Pseudotypes as an Alternative to Reassortant or Triton X-100-treated Wild-type Influenza Viruses in the Neuraminidase Inhibition Enzyme-linked Lectin Assay. Influenza Other Respir. Viruses 2019, 13, 504–516. [Google Scholar] [CrossRef] [PubMed]
  259. Ferrara, F.; Del Rosario, J.M.M.; da Costa, K.A.S.; Kinsley, R.; Scott, S.; Fereidouni, S.; Thompson, C.; Kellam, P.; Gilbert, S.; Carnell, G.; et al. Development of Lentiviral Vectors Pseudotyped With Influenza B Hemagglutinins: Application in Vaccine Immunogenicity, MAb Potency, and Sero-Surveillance Studies. Front. Immunol. 2021, 12, 661379. [Google Scholar] [CrossRef] [PubMed]
  260. Jacob, S.T.; Crozier, I.; Fischer, W.A., II; Hewlett, A.; Kraft, C.S.; de La Vega, M.-A.; Soka, M.J.; Wahl, V.; Griffiths, A.; Bollinger, L.; et al. Ebola Virus Disease. Nat. Rev. Dis. Prim. 2020, 6, 13. [Google Scholar] [CrossRef] [PubMed]
  261. Shifflett, K.; Marzi, A. Marburg Virus Pathogenesis—Differences and Similarities in Humans and Animal Models. Virol. J. 2019, 16, 165. [Google Scholar] [CrossRef] [PubMed]
  262. Cutts, T.; Leung, A.; Banadyga, L.; Krishnan, J. Inactivation Validation of Ebola, Marburg, and Lassa Viruses in AVL and Ethanol-Treated Viral Cultures. Viruses 2024, 16, 1354. [Google Scholar] [CrossRef] [PubMed]
  263. Fe Medina, M.; Kobinger, G.P.; Rux, J.; Gasmi, M.; Looney, D.J.; Bates, P.; Wilson, J.M. Lentiviral Vectors Pseudotyped with Minimal Filovirus Envelopes Increased Gene Transfer in Murine Lung. Mol. Ther. 2003, 8, 777–789. [Google Scholar] [CrossRef] [PubMed]
  264. Sinn, P.L.; Hickey, M.A.; Staber, P.D.; Dylla, D.E.; Jeffers, S.A.; Davidson, B.L.; Sanders, D.A.; McCray, P.B. Lentivirus Vectors Pseudotyped with Filoviral Envelope Glycoproteins Transduce Airway Epithelia from the Apical Surface Independently of Folate Receptor Alpha. J. Virol. 2003, 77, 5902–5910. [Google Scholar] [CrossRef] [PubMed]
  265. Cao, Z.; Jin, H.; Wong, G.; Zhang, Y.; Jiao, C.; Feng, N.; Wu, F.; Xu, S.; Chi, H.; Zhao, Y.; et al. The Application of a Safe Neutralization Assay for Ebola Virus Using Lentivirus-Based Pseudotyped Virus. Virol. Sin. 2021, 36, 1648–1651. [Google Scholar] [CrossRef] [PubMed]
  266. Almasaud, A.; Alharbi, N.K.; Hashem, A.M. Generation of MERS-CoV Pseudotyped Viral Particles for the Evaluation of Neutralizing Antibodies in Mammalian Sera. Methods Mol. Biol. 2020, 2099, 117–126. [Google Scholar] [CrossRef] [PubMed]
  267. Sampson, A.T.; Heeney, J.; Cantoni, D.; Ferrari, M.; Sans, M.S.; George, C.; Di Genova, C.; Mayora Neto, M.; Einhauser, S.; Asbach, B.; et al. Coronavirus Pseudotypes for All Circulating Human Coronaviruses for Quantification of Cross-Neutralizing Antibody Responses. Viruses 2021, 13, 1579. [Google Scholar] [CrossRef] [PubMed]
  268. Ryu, W.-S. Other Positive-Strand RNA Viruses. In Molecular Virology of Human Pathogenic Viruses; Elsevier: Amsterdam, The Netherlands, 2017. [Google Scholar]
  269. Khongwichit, S.; Chansaenroj, J.; Chirathaworn, C.; Poovorawan, Y. Chikungunya Virus Infection: Molecular Biology, Clinical Characteristics, and Epidemiology in Asian Countries. J. Biomed. Sci. 2021, 28, 84. [Google Scholar] [CrossRef] [PubMed]
  270. Gylfe, Å.; Ribers, Å.; Forsman, O.; Bucht, G.; Alenius, G.-M.; Wållberg-Jonsson, S.; Ahlm, C.; Evander, M. Mosquitoborne Sindbis Virus Infection and Long-Term Illness. Emerg. Infect. Dis. 2018, 24, 1141–1142. [Google Scholar] [CrossRef] [PubMed]
  271. Mathiot, C.C.; Grimaud, G.; Garry, P.; Bouquety, J.C.; Mada, A.; Daguisy, A.M.; Georges, A.J. An Outbreak of Human Semliki Forest Virus Infections in Central African Republic. Am. J. Trop. Med. Hyg. 1990, 42, 386–393. [Google Scholar] [CrossRef] [PubMed]
  272. Flies, E.J.; Lau, C.L.; Carver, S.; Weinstein, P. Another Emerging Mosquito-Borne Disease? Endemic Ross River Virus Transmission in the Absence of Marsupial Reservoirs. Bioscience 2018, 68, 288–293. [Google Scholar] [CrossRef]
  273. Kahl, C.A.; Marsh, J.; Fyffe, J.; Sanders, D.A.; Cornetta, K. Human Immunodeficiency Virus Type 1-Derived Lentivirus Vectors Pseudotyped with Envelope Glycoproteins Derived from Ross River Virus and Semliki Forest Virus. J. Virol. 2004, 78, 1421–1430. [Google Scholar] [CrossRef] [PubMed]
  274. Kahl, C.A.; Pollok, K.; Haneline, L.S.; Cornetta, K. Lentiviral Vectors Pseudotyped with Glycoproteins from Ross River and Vesicular Stomatitis Viruses: Variable Transduction Related to Cell Type and Culture Conditions. Mol. Ther. 2005, 11, 470–482. [Google Scholar] [CrossRef] [PubMed]
  275. Morizono, K.; Ku, A.; Xie, Y.; Harui, A.; Kung, S.K.P.; Roth, M.D.; Lee, B.; Chen, I.S.Y. Redirecting Lentiviral Vectors Pseudotyped with Sindbis Virus-Derived Envelope Proteins to DC-SIGN by Modification of N-Linked Glycans of Envelope Proteins. J. Virol. 2010, 84, 6923–6934. [Google Scholar] [CrossRef] [PubMed]
  276. Kishishita, N.; Takeda, N.; Nuegoonpipat, A.; Anantapreecha, S. A Safe and Convenient Chikungunya-Pseudotyped Lentiviral Vector for Epidemiological Studies on Chikungunya Virus in Thailand. Int. J. Infect. Dis. 2012, 16, e262. [Google Scholar] [CrossRef]
  277. Kishishita, N.; Takeda, N.; Anuegoonpipat, A.; Anantapreecha, S. Development of a Pseudotyped-Lentiviral-Vector-Based Neutralization Assay for Chikungunya Virus Infection. J. Clin. Microbiol. 2013, 51, 1389–1395. [Google Scholar] [CrossRef] [PubMed]
  278. Eckels, K.H.; Putnak, R. Formalin-Inactivated Whole Virus and Recombinant Subunit Flavivirus Vaccines. Adv. Virus Res. 2003, 61, 395–418. [Google Scholar] [PubMed]
  279. Lazear, H.M.; Diamond, M.S. Zika Virus: New Clinical Syndromes and Its Emergence in the Western Hemisphere. J. Virol. 2016, 90, 4864–4875. [Google Scholar] [CrossRef] [PubMed]
  280. Manns, M.P.; Buti, M.; Gane, E.; Pawlotsky, J.-M.; Razavi, H.; Terrault, N.; Younossi, Z. Hepatitis C Virus Infection. Nat. Rev. Dis. Prim. 2017, 3, 17006. [Google Scholar] [CrossRef] [PubMed]
  281. Liu, J.; Mao, Y.; Li, Q.; Qiu, Z.; Li, J.J.; Li, X.; Liang, W.; Xu, M.; Li, A.; Cai, X.; et al. Efficient Gene Transfer to Kidney Using a Lentiviral Vector Pseudotyped with Zika Virus Envelope Glycoprotein. Hum. Gene Ther. 2022, 33, 1269–1278. [Google Scholar] [CrossRef] [PubMed]
  282. Bartosch, B.; Dubuisson, J.; Cosset, F.-L. Infectious Hepatitis C Virus Pseudo-Particles Containing Functional E1–E2 Envelope Protein Complexes. J. Exp. Med. 2003, 197, 633–642. [Google Scholar] [CrossRef] [PubMed]
  283. Liu, H.; Wu, R.; Yuan, L.; Tian, G.; Huang, X.; Wen, Y.; Ma, X.; Huang, Y.; Yan, Q.; Zhao, Q.; et al. Introducing a Cleavable Signal Peptide Enhances the Packaging Efficiency of Lentiviral Vectors Pseudotyped with Japanese Encephalitis Virus Envelope Proteins. Virus Res. 2017, 229, 9–16. [Google Scholar] [CrossRef] [PubMed]
  284. Nair, N.; Osterhaus, A.D.M.E.; Rimmelzwaan, G.F.; Prajeeth, C.K. Rift Valley Fever Virus—Infection, Pathogenesis and Host Immune Responses. Pathogens 2023, 12, 1174. [Google Scholar] [CrossRef] [PubMed]
  285. Hawman, D.W.; Feldmann, H. Crimean-Congo Haemorrhagic Fever Virus. Nat. Rev. Microbiol. 2023, 21, 463–477. [Google Scholar] [CrossRef] [PubMed]
  286. Martínez, V.P.; Di Paola, N.; Alonso, D.O.; Pérez-Sautu, U.; Bellomo, C.M.; Iglesias, A.A.; Coelho, R.M.; López, B.; Periolo, N.; Larson, P.A.; et al. “Super-Spreaders” and Person-to-Person Transmission of Andes Virus in Argentina. N. Engl. J. Med. 2020, 383, 2230–2241. [Google Scholar] [CrossRef] [PubMed]
  287. Cifuentes-Muñoz, N.; Darlix, J.-L.; Tischler, N.D. Development of a Lentiviral Vector System to Study the Role of the Andes Virus Glycoproteins. Virus Res. 2010, 153, 29–35. [Google Scholar] [CrossRef] [PubMed]
  288. Li, Y.; Zhao, Y.; Wang, C.; Zheng, X.; Wang, H.; Gai, W.; Jin, H.; Yan, F.; Qiu, B.; Gao, Y.; et al. Packaging of Rift Valley Fever Virus Pseudoviruses and Establishment of a Neutralization Assay Method. J. Vet. Sci. 2018, 19, 200–206. [Google Scholar] [CrossRef] [PubMed]
  289. Vasmehjani, A.A.; Salehi-Vaziri, M.; Azadmanesh, K.; Nejati, A.; Pouriayevali, M.H.; Gouya, M.M.; Parsaeian, M.; Shahmahmoodi, S. Efficient Production of a Lentiviral System for Displaying Crimean-Congo Hemorrhagic Fever Virus Glycoproteins Reveals a Broad Range of Cellular Susceptibility and Neutralization Ability. Arch. Virol. 2020, 165, 1109–1120. [Google Scholar] [CrossRef] [PubMed]
  290. Hastie, K.M.; Melnik, L.I.; Cross, R.W.; Klitting, R.M.; Andersen, K.G.; Saphire, E.O.; Garry, R.F. The Arenaviridae Family: Knowledge Gaps, Animal Models, Countermeasures, and Prototype Pathogens. J. Infect. Dis. 2023, 228, S359–S375. [Google Scholar] [CrossRef] [PubMed]
  291. Larson, R.A.; Dai, D.; Hosack, V.T.; Tan, Y.; Bolken, T.C.; Hruby, D.E.; Amberg, S.M. Identification of a Broad-Spectrum Arenavirus Entry Inhibitor. J. Virol. 2008, 82, 10768–10775. [Google Scholar] [CrossRef] [PubMed]
  292. Schountz, T. Unraveling the Mystery of Tacaribe Virus. mSphere 2024, 9, e00605-24. [Google Scholar] [CrossRef] [PubMed]
  293. Shimojima, M.; Ströher, U.; Ebihara, H.; Feldmann, H.; Kawaoka, Y. Identification of Cell Surface Molecules Involved in Dystroglycan-Independent Lassa Virus Cell Entry. J. Virol. 2012, 86, 2067–2078. [Google Scholar] [CrossRef] [PubMed]
  294. Beyer, W.R.; Westphal, M.; Ostertag, W.; von Laer, D. Oncoretrovirus and Lentivirus Vectors Pseudotyped with Lymphocytic Choriomeningitis Virus Glycoprotein: Generation, Concentration, and Broad Host Range. J. Virol. 2002, 76, 1488–1495. [Google Scholar] [CrossRef] [PubMed]
  295. Kumar, N.; Wang, J.; Lan, S.; Danzy, S.; McLay Schelde, L.; Seladi-Schulman, J.; Ly, H.; Liang, Y. Characterization of Virulence-Associated Determinants in the Envelope Glycoprotein of Pichinde Virus. Virology 2012, 433, 97–103. [Google Scholar] [CrossRef] [PubMed]
  296. Shao, J.; Liu, X.; Ly, H.; Liang, Y. Characterization of the Glycoprotein Stable Signal Peptide in Mediating Pichinde Virus Replication and Virulence. J. Virol. 2016, 90, 10390–10397. [Google Scholar] [CrossRef] [PubMed]
  297. Dreja, H.; Piechaczyk, M. The Effects of N-Terminal Insertion into VSV-G of an ScFv Peptide. Virol. J. 2006, 3, 69. [Google Scholar] [CrossRef] [PubMed]
  298. Goyvaerts, C.; De Groeve, K.; Dingemans, J.; Van Lint, S.; Robays, L.; Heirman, C.; Reiser, J.; Zhang, X.-Y.; Thielemans, K.; De Baetselier, P.; et al. Development of the Nanobody Display Technology to Target Lentiviral Vectors to Antigen-Presenting Cells. Gene Ther. 2012, 19, 1133–1140. [Google Scholar] [CrossRef] [PubMed]
  299. Ahani, R.; Roohvand, F.; Cohan, R.A.; Etemadzadeh, M.H.; Mohajel, N.; Behdani, M.; Shahosseini, Z.; Madani, N.; Azadmanesh, K. Sindbis Virus-Pseudotyped Lentiviral Vectors Carrying VEGFR2-Specific Nanobody for Potential Transductional Targeting of Tumor Vasculature. Mol. Biotechnol. 2016, 58, 738–747. [Google Scholar] [CrossRef] [PubMed]
  300. Luong, Q.T.; Morizono, K.; Chen, I.S.; De Vos, S. Construction of Lentiviruses Pseudotyped with Sindbis E2-Single Chain Antibody (SCA) Fusions or Membrane-Anchored SCAs for Targeted Therapies in Lymphoma. Blood 2009, 114, 3571. [Google Scholar] [CrossRef]
  301. Anastasov, N.; Höfig, I.; Mall, S.; Krackhardt, A.M.; Thirion, C. Optimized Lentiviral Transduction Protocols by Use of a Poloxamer Enhancer, Spinoculation, and ScFv-Antibody Fusions to VSV-G. Methods Mol. Biol. 2016, 1448, 49–61. [Google Scholar] [CrossRef] [PubMed]
  302. Höfig, I.; Barth, S.; Salomon, M.; Jagusch, V.; Atkinson, M.J.; Anastasov, N.; Thirion, C. Systematic Improvement of Lentivirus Transduction Protocols by Antibody Fragments Fused to VSV-G as Envelope Glycoprotein. Biomaterials 2014, 35, 4204–4212. [Google Scholar] [CrossRef] [PubMed]
  303. Berckmueller, K.; Thomas, J.; Taha, E.A.; Choo, S.; Madhu, R.; Kanestrom, G.; Rupert, P.B.; Strong, R.; Kiem, H.-P.; Radtke, S. CD90-Targeted Lentiviral Vectors for HSC Gene Therapy. Mol. Ther. 2023, 31, 2901–2913. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of HIV-1 genome and three generations of lentiviral vectors. (A) The HIV-1 genome is flanked by an LTR that consists of U3 (long red), R (gray), and U5 (short red) regions, encoding structural (gag, gag-pol, and env), regulatory (rev and tat), and accessory (vif, vpr, vpu, and nef) protein-coding genes. The packaging signal (ψ) is required to package vRNA into the viral particle. The Gag precursor contains the core viral proteins—including the matrix (MA), capsid (CA), nucleocapsid (NC), and p6 proteins—while the Gag-Pol precursor additionally contains the protease (PR), reverse transcriptase (RT), and integrase (IN) proteins. (B) First-generation vector system. The packaging plasmid includes all HIV-1 genes in a single vector except for env and vpu. The separated envelope plasmid provides envelope protein. The transfer plasmid contains a transgene expression cassette flanked by the HIV-1 LTRs. (C) Second-generation vector system. The packaging plasmid contains all HIV-1 genes in a single vector without accessory protein sequences. The separated envelope plasmid provides envelope protein. The transfer plasmid contains a transgene expression cassette flanked by the HIV-1 LTRs. (D) Third-generation vector system. The packaging plasmid contains only the gag-pol sequence with RRE element. The separated envelope plasmid provides envelope protein. The transfer plasmid includes CMV promoter-driven truncated 5′-LTR, required cis-acting elements, transgene expression cassette, WPRE element and partially deleted self-inactivating ΔU3-LTR (SIN-LTR) with polyadenylation (pA), and separated Rev expression plasmid vector.
Figure 1. Schematic illustration of HIV-1 genome and three generations of lentiviral vectors. (A) The HIV-1 genome is flanked by an LTR that consists of U3 (long red), R (gray), and U5 (short red) regions, encoding structural (gag, gag-pol, and env), regulatory (rev and tat), and accessory (vif, vpr, vpu, and nef) protein-coding genes. The packaging signal (ψ) is required to package vRNA into the viral particle. The Gag precursor contains the core viral proteins—including the matrix (MA), capsid (CA), nucleocapsid (NC), and p6 proteins—while the Gag-Pol precursor additionally contains the protease (PR), reverse transcriptase (RT), and integrase (IN) proteins. (B) First-generation vector system. The packaging plasmid includes all HIV-1 genes in a single vector except for env and vpu. The separated envelope plasmid provides envelope protein. The transfer plasmid contains a transgene expression cassette flanked by the HIV-1 LTRs. (C) Second-generation vector system. The packaging plasmid contains all HIV-1 genes in a single vector without accessory protein sequences. The separated envelope plasmid provides envelope protein. The transfer plasmid contains a transgene expression cassette flanked by the HIV-1 LTRs. (D) Third-generation vector system. The packaging plasmid contains only the gag-pol sequence with RRE element. The separated envelope plasmid provides envelope protein. The transfer plasmid includes CMV promoter-driven truncated 5′-LTR, required cis-acting elements, transgene expression cassette, WPRE element and partially deleted self-inactivating ΔU3-LTR (SIN-LTR) with polyadenylation (pA), and separated Rev expression plasmid vector.
Viruses 17 01036 g001
Figure 2. Clinical applications of lentiviral vectors in major human diseases. The applications cover various diseases—including neurological, ocular, hematological, metabolic, cancer, immunological, skin, viral infection, and lung diseases—as well as long-term follow-up for the lentiviral vector gene therapy, as detailed in Supplementary Table S1.
Figure 2. Clinical applications of lentiviral vectors in major human diseases. The applications cover various diseases—including neurological, ocular, hematological, metabolic, cancer, immunological, skin, viral infection, and lung diseases—as well as long-term follow-up for the lentiviral vector gene therapy, as detailed in Supplementary Table S1.
Viruses 17 01036 g002
Figure 3. Developmental stages of the gene therapy era and approved products. The first-stage gene therapy products originated from the first-generation delivery platforms and oligonucleotides. The second-stage products are based on second-generation or more advanced delivery platforms and oligonucleotides. The third-stage product includes precise genome editing tools.
Figure 3. Developmental stages of the gene therapy era and approved products. The first-stage gene therapy products originated from the first-generation delivery platforms and oligonucleotides. The second-stage products are based on second-generation or more advanced delivery platforms and oligonucleotides. The third-stage product includes precise genome editing tools.
Viruses 17 01036 g003
Figure 4. Self-inactivating (SIN) vectors and current concerns. (A) Non-SIN vectors flanked by full LTRs in the integrated provirus, each LTR consisting of U3 (long red), R (gray), and U5 (short red) regions. The packaging signal (ψ). The U3 region has an enhancer-promoter sequence. The transcription starting site is the U3 and R junction in 5′-LTR, indicating potential transcriptional leakage on the 3′-LTR. The enhancer-promoter activity in the LTRs and possible splicing donor (SD) acceptor events are prone to activating nearby genes at the integrated sites. (B) The deletion of the enhancer-promoter sequence in 3′-U3 (ΔU3) allows for SIN-LTR. The SIN vectors are flanked by SIN-LTRs in the integrated provirus, each SIN-LTR consisting of ΔU3 (Δlong red), R (gray), and U5 (short red) regions. Transgene expression is driven by a suitable internal promoter. The loss of enhancer-promoter function in SIN-LTR minimizes the negative impact of the integrated provirus on nearby genes. (C) Inclusion of the Myeloproliferative sarcoma virus enhancer, Negative control region deleted (MNDU3) internal promoter to express ABCD1 cDNA in SIN lentiviral vector, leading to activation of nearby genes at the integrated site.
Figure 4. Self-inactivating (SIN) vectors and current concerns. (A) Non-SIN vectors flanked by full LTRs in the integrated provirus, each LTR consisting of U3 (long red), R (gray), and U5 (short red) regions. The packaging signal (ψ). The U3 region has an enhancer-promoter sequence. The transcription starting site is the U3 and R junction in 5′-LTR, indicating potential transcriptional leakage on the 3′-LTR. The enhancer-promoter activity in the LTRs and possible splicing donor (SD) acceptor events are prone to activating nearby genes at the integrated sites. (B) The deletion of the enhancer-promoter sequence in 3′-U3 (ΔU3) allows for SIN-LTR. The SIN vectors are flanked by SIN-LTRs in the integrated provirus, each SIN-LTR consisting of ΔU3 (Δlong red), R (gray), and U5 (short red) regions. Transgene expression is driven by a suitable internal promoter. The loss of enhancer-promoter function in SIN-LTR minimizes the negative impact of the integrated provirus on nearby genes. (C) Inclusion of the Myeloproliferative sarcoma virus enhancer, Negative control region deleted (MNDU3) internal promoter to express ABCD1 cDNA in SIN lentiviral vector, leading to activation of nearby genes at the integrated site.
Viruses 17 01036 g004
Figure 5. Schematic illustration of third-generation and next-generation lentiviral vectors. (A) Third-generation vector system with 4 plasmid vectors and integrated provirus. (B) Lenti-X fourth-generation vector system with 6 plasmid vectors and integrated provirus. (C) LTR1 fourth-generation vector system with 4 plasmid vectors and integrated provirus. (D) Tetravecta fourth-generation vector system with 3 plasmid vectors and integrated provirus.
Figure 5. Schematic illustration of third-generation and next-generation lentiviral vectors. (A) Third-generation vector system with 4 plasmid vectors and integrated provirus. (B) Lenti-X fourth-generation vector system with 6 plasmid vectors and integrated provirus. (C) LTR1 fourth-generation vector system with 4 plasmid vectors and integrated provirus. (D) Tetravecta fourth-generation vector system with 3 plasmid vectors and integrated provirus.
Viruses 17 01036 g005
Figure 6. Reverse transcription of lentiviral vectors. (A) Reverse transcription with two-strand DNA transfer. Viral genomic (+) ssRNA of the third-generation system is flanked with R elements in both 5′-LTR and 3′-LTR. The required cis-acting elements are located after the 5′-LTR. (I) The binding of tRNA to the primer binding site (PBS) initiates (−) strand DNA synthesis via RT. The RNA-DNA hybrid is degraded by the RNase H activity of RT during the reverse transcription. (II) The (−) strand DNA extension is continued after the (−) strand DNA transfer. The complementary (R) sequence is involved in the first jump. (II-III) All template RNA digestion, including tRNA, except cPPT and PPT, allows for initiation of the PPT-primed (+) strand DNA synthesis. (IV) The annealing of PBS sites transfers the (+) strand DNA, facilitating bidirectional DNA synthesis to continue and complete. The provirus flanked with ΔU3-LTR, containing required cis-acting elements, has risk of potential leakage of packageable vRNA. (B) Reverse transcription with single-strand DNA transfer. Viral genomic (+) ssRNA of the LTR1 fourth-generation system starts with PBS and has a single full ΔU3-LTR followed by PBS, the packaging signal (ψ), and RRE. (II) The binding of tRNA to the PBS after the full ΔU3-LTR initiates (−) strand DNA synthesis and extension by RT, without (I) initiation and (−) strand DNA transfer. The RNA–DNA hybrid is degraded by the RNase H activity of RT during reverse transcription. (II–III) All template RNA digestion, including tRNA, except cPPT and PPT, allows for initiation of the PPT-primed (+) strand DNA synthesis, resulting in removal of (ψ) and RRE. (IV) The annealing of PBS sites transfers the (+) strand DNA, enabling bidirectional DNA synthesis to continue and complete. It cannot be packaged in the event of potential leakage of LTR-1 proviral vRNA, due to the elimination of (ψ).
Figure 6. Reverse transcription of lentiviral vectors. (A) Reverse transcription with two-strand DNA transfer. Viral genomic (+) ssRNA of the third-generation system is flanked with R elements in both 5′-LTR and 3′-LTR. The required cis-acting elements are located after the 5′-LTR. (I) The binding of tRNA to the primer binding site (PBS) initiates (−) strand DNA synthesis via RT. The RNA-DNA hybrid is degraded by the RNase H activity of RT during the reverse transcription. (II) The (−) strand DNA extension is continued after the (−) strand DNA transfer. The complementary (R) sequence is involved in the first jump. (II-III) All template RNA digestion, including tRNA, except cPPT and PPT, allows for initiation of the PPT-primed (+) strand DNA synthesis. (IV) The annealing of PBS sites transfers the (+) strand DNA, facilitating bidirectional DNA synthesis to continue and complete. The provirus flanked with ΔU3-LTR, containing required cis-acting elements, has risk of potential leakage of packageable vRNA. (B) Reverse transcription with single-strand DNA transfer. Viral genomic (+) ssRNA of the LTR1 fourth-generation system starts with PBS and has a single full ΔU3-LTR followed by PBS, the packaging signal (ψ), and RRE. (II) The binding of tRNA to the PBS after the full ΔU3-LTR initiates (−) strand DNA synthesis and extension by RT, without (I) initiation and (−) strand DNA transfer. The RNA–DNA hybrid is degraded by the RNase H activity of RT during reverse transcription. (II–III) All template RNA digestion, including tRNA, except cPPT and PPT, allows for initiation of the PPT-primed (+) strand DNA synthesis, resulting in removal of (ψ) and RRE. (IV) The annealing of PBS sites transfers the (+) strand DNA, enabling bidirectional DNA synthesis to continue and complete. It cannot be packaged in the event of potential leakage of LTR-1 proviral vRNA, due to the elimination of (ψ).
Viruses 17 01036 g006
Figure 7. Lentiviral particle structure and its pseudotyping. (A) The structure of the mature HIV-1 virion, resembling a lentiviral vector particle. The two copies of viral genomic ssRNA are encapsulated by the capsid (CA) along with the nucleocapsid (NC), reverse transcriptase (RT), integrase (IN), and tRNA lys, which are surrounded by a lipid membrane and matrix protein (MA) with incorporated envelope and receptor proteins. Protease (PR), viral, and host cell proteins are also assembled in the particle. (B) The orders, families, or subfamilies of viruses that encode envelope glycoproteins which are employed to pseudotype lentiviral vectors, as well as to engineer envelope proteins.
Figure 7. Lentiviral particle structure and its pseudotyping. (A) The structure of the mature HIV-1 virion, resembling a lentiviral vector particle. The two copies of viral genomic ssRNA are encapsulated by the capsid (CA) along with the nucleocapsid (NC), reverse transcriptase (RT), integrase (IN), and tRNA lys, which are surrounded by a lipid membrane and matrix protein (MA) with incorporated envelope and receptor proteins. Protease (PR), viral, and host cell proteins are also assembled in the particle. (B) The orders, families, or subfamilies of viruses that encode envelope glycoproteins which are employed to pseudotype lentiviral vectors, as well as to engineer envelope proteins.
Viruses 17 01036 g007
Table 1. Examples of phenotypically unmixed and mixed pseudotyping of lentiviral vectors.
Table 1. Examples of phenotypically unmixed and mixed pseudotyping of lentiviral vectors.
PhenotypeGlycoproteinsOriginPseudotyping and
Lentiviral Vectors
Refs.
UnmixedVSV-G (V)Vesicular Stomatitis VirusSingleV-LV[224]
Fusion (F)
Hemagglutinin-neuraminidase (HN)
Sendai virusDualF/HN-SIV, F/HN-LV[148,149]
Hemagglutinin (H)
Fusion (F)
Measles virusDualH/F-LV[221]
Small hydrophobic protein (SH)
Glycoprotein G (G)
Fusion (F)
Respiratory syncytial virusTripleSH/G/F-LV[220]
Large surface protein (L)
Medium surface protein (M)
Small surface protein (S)
Hepatitis B virusTripleL/M/S-LV[222]
MixedVSV-G (V)
Fusion (F)
Vesicular Stomatitis Virus Sendai virusDualV/F-LV[217,218]
VSV-G (V)
Hemagglutinin-neuraminidase (HN)
Vesicular Stomatitis Virus Sendai virusDualV/HN-LV[217,218]
VSV-G (V)
Fusion (F)
Hemagglutinin-neuraminidase (HN)
Vesicular Stomatitis Virus Sendai virusTripleV/F/HN-LV[217,218]
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

Jargalsaikhan, B.-E.; Muto, M.; Ema, M. The Era of Gene Therapy: The Advancement of Lentiviral Vectors and Their Pseudotyping. Viruses 2025, 17, 1036. https://doi.org/10.3390/v17081036

AMA Style

Jargalsaikhan B-E, Muto M, Ema M. The Era of Gene Therapy: The Advancement of Lentiviral Vectors and Their Pseudotyping. Viruses. 2025; 17(8):1036. https://doi.org/10.3390/v17081036

Chicago/Turabian Style

Jargalsaikhan, Bat-Erdene, Masanaga Muto, and Masatsugu Ema. 2025. "The Era of Gene Therapy: The Advancement of Lentiviral Vectors and Their Pseudotyping" Viruses 17, no. 8: 1036. https://doi.org/10.3390/v17081036

APA Style

Jargalsaikhan, B.-E., Muto, M., & Ema, M. (2025). The Era of Gene Therapy: The Advancement of Lentiviral Vectors and Their Pseudotyping. Viruses, 17(8), 1036. https://doi.org/10.3390/v17081036

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