Skip to main content
Human Vaccines & Immunotherapeutics logoLink to Human Vaccines & Immunotherapeutics
. 2017 Jun 12;13(12):2837–2848. doi: 10.1080/21645515.2017.1330236

Advancements in DNA vaccine vectors, non-mechanical delivery methods, and molecular adjuvants to increase immunogenicity

John J Suschak a, James A Williams b, Connie S Schmaljohn a,
PMCID: PMC5718814  PMID: 28604157

ABSTRACT

A major advantage of DNA vaccination is the ability to induce both humoral and cellular immune responses. DNA vaccines are currently used in veterinary medicine, but have not achieved widespread acceptance for use in humans due to their low immunogenicity in early clinical studies. However, recent clinical data have re-established the value of DNA vaccines, particularly in priming high-level antigen-specific antibody responses. Several approaches have been investigated for improving DNA vaccine efficacy, including advancements in DNA vaccine vector design, the inclusion of genetically engineered cytokine adjuvants, and novel non-mechanical delivery methods. These strategies have shown promise, resulting in augmented adaptive immune responses in not only mice, but also in large animal models. Here, we review advancements in each of these areas that show promise for increasing the immunogenicity of DNA vaccines.

KEYWORDS: DNA Vaccine, immunogenicity, molecular adjuvant, plasmid, vaccine delivery

Introduction

The constant emergence, and re-emergence, of known and novel pathogens challenges researchers to develop new vaccination technologies that allow for the rapid development of safe and effective vaccines. Nucleic acid (DNA and RNA) vaccines have characteristics that meet these challenges, including ease of production, scalability, consistency between lots, storage, and safety. DNA vaccine technology usually is based on bacterial plasmids that encode the polypeptide sequence of candidate antigens. The encoded antigen is expressed under a strong eukaryotic promoter, yielding high levels of transgene expression.1 Inclusion of transcriptional enhancers, such as Intron A, enhance the rate of polyadenylation and nuclear transport of messenger RNA (mRNA).2 The vaccine plasmids are generally produced in bacterial culture, purified, and then used to inoculate the host.

Modern DNA vaccine design generally relies on synthesis of the nucleic acid and possibly one-step cloning into the plasmid vector, reducing both the cost and the time to manufacture. Plasmid DNA is also extremely stable at room temperature, reducing the need for a cold chain during transportation. Vaccination with DNA plasmid removes the necessity for protein purification from infectious pathogens, improving safety. Furthermore, DNA vaccination has an excellent safety profile in the clinic, with the most common side effect being mild inflammation at the injection site.3 Importantly, DNA vaccines provide a safe, non-live vaccine approach to inducing balanced immune responses, as the in vivo production of antigen allows for presentation on both class I and class II major histocompatibility complex (MHC) molecules (Fig. 1). This elicits antigen specific antibodies,4 as well as cytotoxic T lymphocyte responses (CTL),5 something that remains elusive in most non-live vaccines. DNA vaccines have also demonstrated the ability to generate follicular T helper populations,6 which are critical for the induction of high quality antigen-specific B cell responses.7

Figure 1.

Figure 1.

Induction of antigen-specific, adaptive immunity by DNA vaccination. Optimized gene sequences are inserted into a plasmid backbone and then delivered to the host via one of several delivery methods. Vaccine plasmid enters the nucleus of host myocytes and antigen presenting cells by using host cellular machinery. The plasmid components are transcribed and protein is produced. The cell provides endogenous post-translational modifications to antigens, producing native protein conformations. Vaccine-derived endogenous peptides are presented on MHC class I molecules. Engulfment of apoptotic or necrotic cells by APC also allows for cross-presentation of cell-associated exogenous antigens. Secreted antigen is captured and processed by antigen presenting cells, and presented on MHC class II. Antigen experienced APC migrate to the draining lymph node to stimulate CD4+ and CD8+ T cell populations. In addition, shed antigen can be captured by antigen-specific high affinity immunoglobulins on the B cell surface for presentation to CD4+ T cells, driving B cell responses.

DNA vaccination has proven successful in several animal models for preventing or treating infectious diseases, allergies, cancer, and autoimmunity.8-12 The early success of small animal studies led to several human clinical trials. However, the protective immunity observed in small animals and non-human primates was not observed in human studies when DNA vaccines were administered alone by needle delivery. Like the more conventional protein-based vaccines, DNA can be delivered by a variety of routes, including intramuscular (IM), intradermal (ID), mucosal, or transdermal delivery. Because DNA plasmids must enter host cell nuclei to be transcribed into mRNA, the early failure of DNA vaccines to elicit strong responses in humans was largely due to their delivery by needle injection, which deposits the DNA in intracellular spaces, rather than within cells. Improved delivery technologies, such as intramuscular or intradermal electroporation, have been used to facilitate transport of DNA into cells, resulting in much better immunogenicity in both clinical and non-clinical studies.13-19 In one study, electroporation-enhanced DNA vaccination resulted in increased polyfunctional antigen-specific CD8+ T cells in patients receiving a HPV DNA vaccine expressing the E6 and E7 genes of HPV16 and HPV18 respectively.20 The majority of DNA vaccinated patients displayed complete regression of their cervical lesions, as well as viral clearance, following DNA delivery. Other mechanical delivery approaches use physical force such as particle bombardment (gene gun) to deliver the DNA plasmids into targeted tissues or cells, with some clinical successes.21-23 Delivery of a Hepatitis B DNA vaccine by particle bombardment resulted in sustained antibody titers in subjects who had previously failed to respond to a licensed subunit vaccine.23 Needle-free pneumatic or jet injectors have also shown promise in both animal and human clinical trials,24-27 and function by injecting a high-pressure, narrow stream of injection liquid into the epidermis or muscles of test subjects. In addition to these improved mechanical delivery methods, several other approaches are being explored to increase the immunogenicity of DNA vaccines in humans. Here we review 3 of these approaches which show promise for advancing DNA vaccines: non-mechanical delivery, inclusion of molecular adjuvants, and improvements in DNA vaccine vectors.

Non-mechanical DNA vaccine delivery

As already mentioned, the greatest impediment to DNA vaccination is low immunogenicity due to difficulties in delivering DNA plasmid into the host cell. The transportation of DNA vaccine plasmids into cellular nuclei requires the crossing of several barriers. Vaccine plasmid must cross the phospholipid cellular membrane through endocytosis or pinocytosis, escape degradation in endosomes and lysosomes, survive cytosolic nucleases, and translocate across the nuclear envelope. In contrast to physical delivery systems, chemical delivery approaches use biopharmaceuticals to increase DNA vaccine transfection efficiency.

The use of liposomes as a carrier molecule has become a popular DNA vaccine delivery method as liposomes not only enhance transfection efficiency, but also have an adjuvant effect. Liposomes are spherical vesicles composed of phospholipids and cholesterol arranged into a lipid bilayer, allowing for fusion with cellular lipid membranes.28 DNA plasmid can be either bound to the liposome surface, or encased within the hydrophobic core of the liposome. This facilitates delivery of the DNA vaccine plasmid into the cells. Importantly, lipid vesicles can be formulated as either unilamellar or multilamellar. Multilamellar vesicles allow for sustained delivery of vaccine over an extended period of time. While the use of liposomes for IM injection has resulted in some reactogenicity issues,29,30 liposome/DNA vaccine complexes have demonstrated an immunological benefit. IM injection of a liposome/influenza nucleoprotein formulation increased antibody titers 20-fold compared with vaccine alone.31,32 Boosting of antibody titers did not diminish the cytotoxic T cell response. Likewise, inclusion of a liposome formulation in a P. falciparum vaccine enhanced the IFN-γ production.33,34 An ensuing human trial involving DNA plasmids encoding the influenza H5 HA, nucleoprotein, and M2 genes reported cellular immune response rates and antibody titers comparable to that of the currently available inactivated protein-based H5 vaccines.35 Additionally, liposomes have shown promise as a candidate for delivery of DNA vaccines to mucosal tissue.36 A recent study demonstrated that vaccination with liposome encapsulated influenza A virus M1 induced both humoral and cellular immune responses that protected against respiratory infection.36 Liposomes have also been shown to be an effective delivery method for intranasal DNA vaccination, conferring protective immune responses against infection.37,38

DNA vaccine delivery can also be accomplished through the use of biodegradable polymeric micro- and nanoparticles consisting of amphiphilic molecules between 0.5–10 µm in size. Similar to loading of DNA plasmid on liposomes, plasmid molecules can be either encapsulated or adsorbed onto the surface of the nanoparticles.39-42 These particles function as a carrier system, protecting the vaccine plasmid from degradation by extracellular deoxyribonucleases. In addition to shielding plasmid DNA from nucleases, micro- and nanoparticles promote the sustained release of vaccine instead of the bolus type of delivery characteristic of larger submicrometer complexes.39,43 High molecular weight cationic polymers have proven significantly more effective than cationic liposomes in aggregating DNA vaccine plasmid. Plasmid DNA immobilized within biodegradable chitosan-coated polymeric microspheres (ranging from 20 to 500 μm) can induce both mucosal and systemic immune responses.44 Microspheres may be delivered either by the oral or intraperitoneal route, allowing for direct transfection of dendritic cells (DC), thereby increasing DC activation. The benefits of microsphere formulations have been shown in mice, non-human primates, and humans 45-49 against a wide range of diseases including hepatitis B,50 tuberculosis,51 and cancer.52 These results suggest that microparticle-based delivery systems are capable of significantly improving DNA vaccine immunogenicity, and boosting cellular and humoral immune responses.

The use of liposomes or nanoparticles appears to be safe and well tolerated in clinical studies. Microparticle-based delivery systems can increase gene expression, as well as, DNA vaccine immunogenicity. Although many of the earliest carrier formulations did not show a significant clinical benefit, more recent studies highlighted herein yielded promising clinical data. As microparticles can be prepared with significant structural diversity (size, surface charge, lipid content), they offer considerable flexibility of vaccine formulation. This allows for optimization of the vaccine based on the specific needs of the clinician.

Molecular adjuvants

Another approach that has been effective in increasing DNA vaccine immunogenicity is the use of “vaccine cocktails” containing the DNA vaccine as well as plasmids encoding adjuvanting immunomodulatory proteins. Plasmid DNA contains unmethylated deoxycytidylate-phosphate-deoxyguanylate (CpG) motifs that function as a “built in” adjuvant.53-59 Molecular adjuvant plasmids expressing cytokines, chemokines, or co-stimulatory molecules may be co-administered with the antigenic DNA vaccine plasmid. Cells transfected by molecular adjuvant plasmids secrete the adjuvant into the surrounding region, stimulating both local antigen presenting cells (APC) and cells in the draining lymph node. This results in durable, but low level, production of immune modulating cytokines that can tailor the immune response toward a more desirable outcome without the concerns of a systemic cytokine storm. While human data are limited, a wide range of inflammatory and helper T cell cytokines have been studied, in conjunction with DNA vaccination, in small animal models.60,61 In particular, we have highlighted a few of the most prominent molecular adjuvants with demonstrated ability to increase DNA vaccine immunogenicity.62 A more comprehensive list of molecular adjuvants is included in Table 1.

Table 1.

Molecular adjuvants tested in vivo.

Molecular Adjuvant Molecule Type Animal Model Adaptive Response Effect References
CD40L Co-Stimulatory Mice Cellular 161
CD80/86 Co-Stimulatory Mice, NHP Cellular 162
GM-CSF Cytokine Mice Humoral 163
ICAM-1 Co-Stimulatory Mice Cellular 164
IFN-γ Cytokine Mice, NHP Cellular 165
IL-2 Cytokine Mice Cellular, Humoral 165,166
IL-4 Cytokine Mice, NHP Humoral 166,167
IL-7 Cytokine Mice Cellular, Humoral 168
IL-8 Chemokine Mice Cellular, Humoral 169,170
IL-10 Cytokine Mice Cellular 166
IL-12 Cytokine Mice, NHP Cellular 98,171
IL-15 Cytokine Mice, NHP Cytokine 98,172
IL-18 Cytokine Mice, NHP Cytokine 166,173
MCP-1 Chemokine Mice Humoral 169
M-CSF Cytokine Mice Cellular 163
MIP-1α Chemokine Mice Humoral 169
RANTES Chemokine Mice Cellular 169, 170

Plasmid-encoded cytokines

Cytokines are a class of immunoregulatory proteins that affect the behavior of other cells, and are critical for immune cell signaling. Cytokine-encoding genes can be delivered either as a separate plasmid, or as additional genes encoded within the antigen containing plasmid. The most extensively studied molecular adjuvant is Interleukin-2 (IL-2). IL-2 plays an essential role in the immune response by promoting the differentiation of naïve T cells into effector T cells, as well as driving the generation of memory T cell pools. It is also required for the proliferation of Natural Killer (NK) cells. Inclusion of IL-2 has resulted in improved immunogenicity for HIV,63-65 influenza,66 and SARS-CoV67 anti-viral DNA vaccines. Interestingly, a therapeutic vaccine encoding for the BCR/ABL-pIRES genes of myeloid leukemia and IL-2 also demonstrated enhanced immune responses, suggesting that IL-2 molcular adjuvants have the capability of alleviating the symptoms of chronic infection.68

Similar to IL-2, IL-15 is a cytokine that induces NK and T cell proliferation. IL-15 is necessary for the generation of primary antigen-specific CD4+ and CD8+ T cell responses. It also plays a substantial role in establishment of memory CD8+ T cell populations.69-73 Results of small animal studies suggest that the adjuvant effect of IL-15 is most potent when delivered in tandem with other cytokines. For example, a synergistic effect was seen when IL-15 and IL-21 were co-delivered with a DNA vaccine against Toxoplasma gondii infection.74,75 Additionally, sequential administration of IL-6, IL-7, and IL-15 genes augmented long-term CD4+ T cell memory responses to a foot and mouth disease DNA vaccine.76 Therefore, depending on the antigen, it may be necessary to deliver IL-15 in combination with other molecular adjuvants. Notably, a study in rhesus macaques suggests that delivery of an IL-15 encoding DNA vaccine itself resulted in increased proliferation of NK and T cells, with no adverse effects.77 Another recent study demonstrated that co-vaccination of rhesus macaques with SIV pol plasmid and HIV env plasmid plus IL-15 allowed for faster control of viremia than the group not formulated with IL-15.78 Moreover, macaques vaccinated with IL-15 exhibited increased T cell proliferation compared with those receiving the antigen plasmid alone, suggesting that IL-15 has a robust effect on T cell memory responses.

IL-12 is another pro-inflammatory cytokine secreted by both dendritic cells and monocytes. IL-12 plays an integral role in shaping the innate and adaptive immune responses to infection.79-83 IL-12 signaling supports the secondary expansion of activated T helper 1 (Th1) cells,79,82,84-86 resulting in high levels of antigen-specific CD8+ T cells, and the expression of cytotoxic mediators such as interferon-γ (IFN-γ), granzyme B, and perforin.82,83 IL-12 was the first cytokine to be evaluated for use as a molecular adjuvant, and several studies have shown that inclusion of IL-12 expression plasmids within the vaccine formulation enhances Th1 immune responses.87-95 Vaccination of mice with a bicistronic plasmid expressing IL-12 and Yersinia pestis resulted in increased mucosal IgA and serum IgG, providing significantly higher levels of protection against challenge than antigen-only groups.96 Studies in rhesus macaques have shown similar increases in DNA vaccine immunogenicity. Co-vaccination with SIV gag and IL-12 allowed for dose sparing,97 as well as increased breadth of T cell responses.89,91,98,99 Additionally, multiple human clinical studies using vaccines adjuvanted with IL-12 have proven safe100 and highly immunogenic, yielding high level CD4+ and CD8+ T cell responses.87,101,102 Furthermore, inclusion of IL-12 expression plasmids can improve weakly immunogenic vaccines. A recent clinical study demonstrated that addition of IL-12 improved the immunogenicity of a Hepatitis B DNA vaccine, resulting in increased vaccine immunogenicity, as well as sustained memory T cell responses.103

The final immunomodulatory cytokine that has received considerable focus as a molecular adjuvant is granulocyte-macrophage colony stimulating factor (GM-CSF). GM-CSF recruits antigen presenting cells to the vaccination site and promotes DC maturation.104 It has been successfully used in multiple DNA vaccines.105-107 Plasmid-encoded GM-CSF, when co-delivered with a rabies virus DNA vaccine in mice, resulted in increased CD4+ T cell responses, antibody production, and protection from lethal viral challenge.108 Likewise, a bicistronic DNA vaccine encoding HIV-1 gp120 and GM-CSF recruited inflammatory cellular infiltrates and elicited a potent CD4+ T cell response.109 However, the benefit of GM-CSF molecular adjuvants remains unclear. Recent studies have shown that co-administration of GM-CSF plasmid with an antigen-encoding DNA vaccine can have deleterious effects. Co-delivery of GM-CSF suppressed the response to a DNA vaccine encoding Dengue virus type 1 and type 2, and also failed to improve the response elicited by a Hepatitis C vaccine.110 Furthermore, inclusion of plasmid GM-CSF provided minimal adjuvant effect when co-administered with a malaria DNA vaccine in rhesus macaques.111 Likewise, GM-CSF had no clear effect on T cell responses in patients receiving a melanoma DNA vaccine.112 One possible explanation for these results is that high levels of GM-CSF can expand myeloid suppressor cell populations, and suppress the generation of adaptive immune responses. Alternatively, the lack of improved immunogenicity seen in clinical trials may be due to the relative lack of GM-CSF receptors on rhesus and human APC compared with murine cells.113 While no specific adverse effects have been reported, the use of GM-CSF as an adjuvant may require some fine-tuning, particularly if GM-CSF expression levels must be considered with regards to immunosuppression.

In addition to cytokine-encoding plasmids, several other methods for increasing DNA vaccine immunogenicity exist. The increased understanding of immune signaling pathways has led to the development of adjuvant plasmids encoding adhesion molecules, chemokines, costimulatory molecules, and Toll-like receptor (TLR) ligands. These molecular adjuvants have had some success in small animal models. For example, the innate immune signaling molecule TRIF increased the antibody response generated by a swine fever virus DNA vaccine.114 Moreover, TRIF increased the protective activity of an influenza HA-encoding DNA vaccine.115 Similar results were seen in studies encoding the dsRNA receptors MDA5 and RIG-I.116,117 Additionally, antigen-fusion constructs, whereby the antigen of interest is linked to a “carrier protein,” can increase the immune visibility of the vaccine, and enhance DNA vaccine potency.118-120

A major advantage of DNA vaccination is the ability of multiple molecules such as molecular adjuvants to be inserted into the plasmid. Unlike the addition of recombinant cytokines, co-stimulatory molecules, and TLR ligands, which have a limited duration due to the short half-life of recombinant protein in vivo, molecular adjuvant-encoding plasmids will express protein for the same duration as the antigen, stimulating the immune system for a greater length of time. This can be done without fear of eliciting a cytokine storm, as generation of the adjuvanting signal will be localized to the site of vaccination. Of note, homologous recombination between plasmid-encoded cytokines and the host gene sequence does not appear to be a significant concern, as multiple studies have shown that only extrachromosomal plasmid DNA has been identified following intramuscular injection.121,122 Furthermore, many current plasmids have been-codon optimized to improve gene expression in mammalian cells. This has resulted in changes to the cytokine gene sequence, limiting the possibility for homologous recombination and/or integration. Molecular adjuvants therefore show great promise for both increasing immunogenicity and extending the longevity of the immune response.

Improvements in DNA plasmid design

Plasmid DNA vectors contain functional elements, such as the origin of replication and selection markers, that are only required during the prokaryotic growth process in E. coli. These “bacterial region” elements (Fig. 2) are no longer needed once cell culture is halted, and may have a negative effect on vaccine stability, uptake, and efficacy. Additionally, these elements can pose safety concerns, particularly if widely used antibiotic resistance markers are horizontally transmitted to host enteric bacteria populations.123,124

Figure 2.

Figure 2.

Molecular mechanisms of DNA vaccines. Transfected double stranded B DNA (dsDNA) is sensed by cytoplasmic DNA receptors such as interferon-inducible protein 16 (IFI16), DEAD (Asp-Glu-Ala-Asp) box polypeptide 41 (DDX41) and the cGAMP synthase (cGAS), each of which can activate the STING▸TBK1▸IRF3 pathway to induce type 1 interferon production.143 An additional cytoplasmic innate immune pathway activated nonspecifically by transfected dsDNA is the cytoplasmic AIM2 inflammasome.157 Other dsDNA receptors and innate immune activation pathways exist,143 including a recently identified STING/IRF7 signaling pathway required for DNA vaccine immunogenicity.158 By contrast, the endosomal innate immune receptor TLR9 recognizes specific unmethylated CpG DNA motifs in DNA vaccines. To improve innate immune activation, addition of optimized immunostimulatory CpG motifs in the vector backbone may be used to increase TLR9 activation. Immunostimulatory RNA expressed from the vector may be used to activate alternative RNA sensing innate immune receptors such as RIG-I using an additional RNA Polymerase III RNA expression cassette117 (plasmid backbone adjuvant) or incorporation of RNA recognizing TLR agonist motifs such as CpG RNA into the 3′ UTR.152 Due to limited transgene expression after DNA vaccination in large animals, vector modifications (e.g., <500 bp bacterial region Nanoplasmid™ vectors; intronic bacterial region MIP vectors) and deliveries (e.g., Electroporation) that improve transgene expression also improve adaptive immunity.62,125,159 Adapted under a Creative Commons Attribution license from Williams, 2013.160

These concerns have been addressed by development of small bacterial RNA-based antibiotic free selection markers.124,125 Noncoding RNA markers are preferable to protein markers since proteins, like antibiotic resistance markers, can be expressed in the host organism after vector transfection, or horizontally transmitted to host bacteria. Noncoding RNA markers are also very small (< 200 basepairs) which decreases the overall vector size; this is advantageous since vector transfection efficiency is inversely related to vector size,126-128 perhaps because smaller vectors are more resistant to delivery associated shear forces129 and may have improved nuclear localization since they are more motile in the cytoplasm.130 Additionally, some bacterial region protein marker genes have been shown to dramatically reduce vector expression. For example, the TN5 derived NPT-II kanamycin resistance marker (kanR) gene in the pVAX1 vector bacterial region significantly reduces transgene expression. Three groups have demonstrated that pVAX1 bacterial region mediated repression of transgene expression can be alleviated by replacement of the kanR gene with either a tRNA RNA selection marker, the RNA-OUT antisense RNA selection marker, or the endogenous pUC origin RNAI antisense RNA selection marker.131-133 Consistent with this, removal of the pVAX1 bacterial region in a minicircle vector improved humoral and cellular immune responses up to 3-fold compared with a pVAX1 vector control.134

DNA vaccine vectors with dramatically higher transgene expression have recently been developed through identification of novel bacterial region and eukaryotic region vector configurations. Pioneering work by Mark Kay's laboratory at Stanford University demonstrated that bacterial regions larger than 1 kbase silenced transgene expression in quiescent tissue such as the liver, likely due to untranscribed bacterial region mediated heterochromatin formation that spreads to the eukaryotic region and inactivates the promoter.135-137 Minicircle vectors, in which the bacterial region is removed by the action of a phage recombinase during production, alleviated this silencing.135,136,138 However, production of minicircle vectors is low yield and poorly scalable due to the required in vivo or in vitro recombination during manufacture.139 In an effort to create alternative short bacterial region vectors that could be efficiently manufactured, the Mini-Intronic Plasmid (MIP) and Nanoplasmid™ vector plasmid platforms were developed. MIP vectors incorporate a RNA-OUT selection marker-pUC origin bacterial region within a 3′ UTR intron. In this configuration the bacterial region is within the transcription unit and the downstream polyA signal is linked to the eukaryotic promoter without an intervening selection marker or replication origin. Nanoplasmid™ vectors are RNA-OUT selection marker vectors in which the large pUC bacterial replication origin is replaced by a small R6K bacterial replication origin. In this configuration, the < 500 basepair (bp) bacterial region separates the polyA signal and the eukaryotic promoter. Unlike minicircles, both MIP and Nanoplasmid™ RNA-OUT selection vectors can be efficiently manufactured in gram/liter yields without antibiotic selection.140

As expected, both vector platforms alleviate gene silencing in quiescent tissues similarly to minicircle vectors.141,142 However, unexpectedly both MIP and Nanoplasmid™ vectors dramatically improve overall gene expression up to 10-fold compared with plasmid and minicircle vectors in quiescent (liver) and non-quiescent tissues.141,142 The improved expression level after ID and IM delivery has application to improve DNA vaccination since increased expression level is correlative with improved humoral and cellular immune response.62

Another approach to improve DNA vaccines is to engineer the vector to increase innate immune activation. DNA vaccines are potent triggers of innate immunity. Various studies have determined several innate immune pathways are activated by DNA vaccination (Fig. 2). Most of the intrinsic adjuvant effect of DNA is mediated by cytoplasmic innate immune receptors that nonspecifically recognize B DNA and activate Sting or Inflammasome mediated signaling,53,143 but unmethylated CpG sequences specific for TLR9 activation may also be important for priming CD8 T cell responses.144,145 Along these lines, DNA vaccine vectors may be sequence modified to introduce immunostimulatory xxCGxx TLR9 agonists into the vector to increase innate immune activation. This approach has been used to improve DNA vaccine immunogenicity,58,59,146 but the results are variable. Some of the variability may be due to unintended inhibition of the eukaryotic promoter expression resulting from integration of CpG motifs into non-permissive sites in the vector.125 As well, certain DNA delivery methods may not transfer DNA to the endosome as effectively as other deliveries (e.g. liposomes), preventing unmethylated CpG interaction with, and activation of, TLR9. Part of the complexity is that optimal TLR9 activating xxCGxx motifs are species-specific; different xxCGxx agonist motifs differentially modulate the immune response147 and many xxCGxx motifs are immunosuppressive.

An alternative strategy is to encode immunostimulatory RNA within the plasmid to increase innate immune activation. This approach has the potential advantage that additional innate immune pathways not normally stimulated by DNA alone are activated, resulting in polyvalent activation of multiple innate immune pathways to enhance immune activation.148,149 Like TLR9 for DNA, several innate immune TLRs for RNA are endosomal.150 Activation of these receptors requires motif introduction into an expressed RNA, as well as cytoplasmic RNA shuttling into the endosome by autophagy. For example, 3′UTR incorporation of a 20 bp immunostimulatory ssRNA encoding D type CpG upstream of a 28 bp hairpin dsRNA resulted in a 4-fold increase in antigen reactive IgG titers,151 and a 2-fold increase in IFN-γ secreting CD4+ and CD8+ T cells.152 Moreover, several RNA-sensing innate immune receptors such as RIG-I, MDA5 and DDX3 are cytoplasmic.143 DNA vaccine expressed RNA can be used to target these receptors directly, without autophagy. Of these, RIG-I is of particular interest since RIG-I agonists have demonstrated adjuvant properties to improve the humoral response,153 humoral and CD4+ T cell response,154,155 and CD8+ T cell response153 to co-administered antigens.156 In addition, RIG-I is ubiquitiously expressed in most tissues (expression of TLRs typically is restricted to immune cell subtypes) and certain RIG-I agonists that can be expressed in DNA vaccines (e.g., a blunt dsRNA with a 3′ triphosphate) are structurally conserved between humans and mice. A DNA vaccine vector that co-expresses with antigen a RIG-I dsRNA agonist in a vector backbone encoded RNA Polymerase III transcription unit (Fig. 2) enhanced the humoral and CD8+ T cell response after DNA vaccination.117

DNA vaccines encoding immunostimulatory sequences that selectively improve CTL responses to encoded antigen may have niche application in vaccines for intracellular pathogens or cancer. Innovations that increase transgene expression may be used to improve the performance of immunomodulatory molecular adjuvant plasmids, in addition to traditional antigen expressing DNA vaccine plasmids. Collectively, vector design innovations that improve transgene expression level and innate immune activation are complementary to improved mechanical and non-mechanical DNA vaccine delivery platforms. Combining improved vectors with liposome or polymeric particle non-mechanical delivery, or with needle free injector device delivery, has the potential to increase immunogenicity with these well tolerated, safe, delivery platforms.

Conclusion

While DNA vaccination provides several advantages over more conventional vaccination strategies, further optimization is necessary before it becomes the predominant strategy in human patients. Despite initial setbacks, significant progress has been made in overcoming the problem of low immunogenicity in humans. A clearer understanding of the immune mechanisms governing DNA vaccine immunogenicity has illuminated several pathways that may be useful in further improving DNA vaccine efficacy. A large catalog of cytokines, chemokines, adhesion molecules, and transcription factors are in the process of being tested as molecular adjuvants, although it is likely that each will need to be carefully assessed for safety and tolerability. Likewise, continued development of vaccine delivery methods appears promising. New formulations exploiting sustained vaccine delivery methods, such as slow-releasing micropatches or multilamellar vesicles, are on the horizon. The strong appeal of needle-free injection and mucosal delivery, the ease of design, and the recent clinical successes with DNA vaccines suggests that this approach is on the precipice of redefining the field of vaccinology.

Disclosure of potential conflicts of interest

James A. Williams has an equity interest in Nature Technology Corporation. Due to this relationship with Nature Technology Corporation, the author acknowledges that there is a potential conflict of interest inherent in the publication of this manuscript and assert that an effort to reduce or eliminate that conflict has been made where possible.

Funding

John J. Suschak and Connie S. Schmaljohn would like to acknowledge funding from the Joint Science and Technology Office for Chemical and Biological Defense of the Defense Threat and Reduction Agency. The opinions, interpretations, conclusions, and recommendations are those of the authors and are not necessarily endorsed by the U.S. Army.

References

  • [1].Galvin TA, Muller J, Khan AS. Effect of different promoters on immune responses elicited by HIV-1 gag/env multigenic DNA vaccine in Macaca mulatta and Macaca nemestrina. Vaccine 2000; 18:2566-83; PMID:10775791; https://doi.org/ 10.1016/S0264-410X(99)00569-1 [DOI] [PubMed] [Google Scholar]
  • [2].Huang MT, Gorman CM. Intervening sequences increase efficiency of RNA 3′ processing and accumulation of cytoplasmic RNA. Nucleic Acids Res 1990; 18:937-47; PMID:1690394; https://doi.org/ 10.1093/nar/18.4.937 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [3].Li L, Petrovsky N. Molecular mechanisms for enhanced DNA vaccine immunogenicity. Expert Rev Vaccines 2016; 15:313-29. PMID:26707950 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [4].Goulder PJ, Rowland-Jones SL, McMichael AJ, Walker BD. Anti-HIV cellular immunity: recent advances towards vaccine design. AIDS 1999; 13 Suppl A:S121-36. PMID:10885771 [PubMed] [Google Scholar]
  • [5].Giri M, Ugen KE, Weiner DB. DNA vaccines against human immunodeficiency virus type 1 in the past decade. Clin Microbiol Rev 2004; 17:370-89; PMID:15084506; https://doi.org/ 10.1128/CMR.17.2.370-389.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [6].Wang R, Epstein J, Baraceros FM, Gorak EJ, Charoenvit Y, Carucci DJ, Hedstrom RC, Rahardjo N, Gay T, Hobart P, et al.. Induction of CD4(+) T cell-dependent CD8(+) type 1 responses in humans by a malaria DNA vaccine. Proc Natl Acad Sci U S A 2001; 98:10817-22; PMID:11526203; https://doi.org/ 10.1073/pnas.181123498 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7].Hollister K, Chen Y, Wang S, Wu H, Mondal A, Clegg N, Lu S, Dent A. The role of follicular helper T cells and the germinal center in HIV-1 gp120 DNA prime and gp120 protein boost vaccination. Hum Vaccin Immunother 2014; 10:1985-92; PMID:25424808; https://doi.org/ 10.4161/hv.28659 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [8].Wolff JA, Malone RW, Williams P, Chong W, Acsadi G, Jani A and Felgner PL. Direct gene transfer into mouse muscle in vivo. Science 1990; 247:1465-8; PMID:1690918; https://doi.org/ 10.1126/science.1690918 [DOI] [PubMed] [Google Scholar]
  • [9].Ulmer JB, Donnelly JJ, Parker SE, Rhodes GH, Felgner PL, Dwarki VJ, Gromkowski SH, Deck RR, DeWitt CM, Friedman A, et al.. Heterologous protection against influenza by injection of DNA encoding a viral protein. Science 1993; 259:1745-9; PMID:8456302; https://doi.org/ 10.1126/science.8456302 [DOI] [PubMed] [Google Scholar]
  • [10].Wang S, Kennedy JS, West K, Montefiori DC, Coley S, Lawrence J, Shen S, Green S, Rothman AL, Ennis FA, et al.. Cross-subtype antibody and cellular immune responses induced by a polyvalent DNA prime-protein boost HIV-1 vaccine in healthy human volunteers. Vaccine 2008; 26:3947-57; PMID:18724414; https://doi.org/ 10.1016/j.vaccine.2007.12.060 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [11].Donnelly JJ, Martinez D, Jansen KU, Ellis RW, Montgomery DL, Liu MA. Protection against papillomavirus with a polynucleotide vaccine. J Infect Dis 1996; 173:314-20; PMID:8568291; https://doi.org/ 10.1093/infdis/173.2.314 [DOI] [PubMed] [Google Scholar]
  • [12].Fuller DH, Haynes JR. A qualitative progression in HIV type 1 glycoprotein 120-specific cytotoxic cellular and humoral immune responses in mice receiving a DNA-based glycoprotein 120 vaccine. AIDS Res Hum Retroviruses 1994; 10:1433-41; PMID:7888198; https://doi.org/ 10.1089/aid.1994.10.1433 [DOI] [PubMed] [Google Scholar]
  • [13].Dupuy LC, Richards MJ, Ellefsen B, Chau L, Luxembourg A, Hannaman D, Livingston BD, Schmaljohn CS. A DNA vaccine for venezuelan equine encephalitis virus delivered by intramuscular electroporation elicits high levels of neutralizing antibodies in multiple animal models and provides protective immunity to mice and nonhuman primates. Clin Vaccine Immunol 2011; 18:707-16; PMID:21450977; https://doi.org/ 10.1128/CVI.00030-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [14].Dupuy LC, Richards MJ, Reed DS, Schmaljohn CS. Immunogenicity and protective efficacy of a DNA vaccine against Venezuelan equine encephalitis virus aerosol challenge in nonhuman primates. Vaccine 2010; 28:7345-50; PMID:20851089; https://doi.org/ 10.1016/j.vaccine.2010.09.005 [DOI] [PubMed] [Google Scholar]
  • [15].Grant-Klein RJ, Altamura LA, Badger CV, Bounds CE, Van Deusen NM, Kwilas SA, Vu HA, Warfield KL, Hooper JW, Hannaman D, et al.. Codon-optimized filovirus DNA vaccines delivered by intramuscular electroporation protect cynomolgus macaques from lethal Ebola and Marburg virus challenges. Hum Vaccines Immunotherapeutics 2015; 11:1991-2004; https://doi.org/ 10.1080/21645515.2015.1039757 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Grant-Klein RJ, Van Deusen NM, Badger CV, Hannaman D, Dupuy LC, Schmaljohn CS. A multiagent filovirus DNA vaccine delivered by intramuscular electroporation completely protects mice from ebola and Marburg virus challenge. Hum Vaccin Immunother 2012; 8:1703-6; PMID:22922764; https://doi.org/ 10.4161/hv.21873 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Bagarazzi ML, Yan J, Morrow MP, Shen X, Parker RL, Lee JC, Giffear M, Pankhong P, Khan AS, Broderick KE, et al.. Immunotherapy against HPV16/18 generates potent TH1 and cytotoxic cellular immune responses. Sci Transl Med 2012; 4:155ra138; PMID:23052295; https://doi.org/ 10.1126/scitranslmed.3004414 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [18].Vasan S, Hurley A, Schlesinger SJ, Hannaman D, Gardiner DF, Dugin DP, Boente-Carrera M, Vittorino R, Caskey M, Andersen J, et al.. In vivo electroporation enhances the immunogenicity of an HIV-1 DNA vaccine candidate in healthy volunteers. PLoS One 2011; 6:e19252; PMID:21603651; https://doi.org/ 10.1371/journal.pone.0019252 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [19].Low L, Mander A, McCann K, Dearnaley D, Tjelle T, Mathiesen I, Stevenson F, Ottensmeier CH. DNA vaccination with electroporation induces increased antibody responses in patients with prostate cancer. Hum Gene Ther 2009; 20:1269-78; PMID:19619001; https://doi.org/ 10.1089/hum.2009.067 [DOI] [PubMed] [Google Scholar]
  • [20].Kim TJ, Jin HT, Hur SY, Yang HG, Seo YB, Hong SR, Lee CW, Kim S, Woo JW, Park KS, et al.. Clearance of persistent HPV infection and cervical lesion by therapeutic DNA vaccine in CIN3 patients. Nat Commun 2014; 5:5317; PMID:25354725; https://doi.org/ 10.1038/ncomms6317 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [21].Tang DC, DeVit M, Johnston SA. Genetic immunization is a simple method for eliciting an immune response. Nature 1992; 356:152-4; PMID:1545867; https://doi.org/ 10.1038/356152a0 [DOI] [PubMed] [Google Scholar]
  • [22].Roy MJ, Wu MS, Barr LJ, Fuller JT, Tussey LG, Speller S, Culp J, Burkholder JK, Swain WF, Dixon RM, et al.. Induction of antigen-specific CD8+ T cells, T helper cells, and protective levels of antibody in humans by particle-mediated administration of a hepatitis B virus DNA vaccine. Vaccine 2000; 19:764-78; PMID:11115698; https://doi.org/ 10.1016/S0264-410X(00)00302-9 [DOI] [PubMed] [Google Scholar]
  • [23].Rottinghaus ST, Poland GA, Jacobson RM, Barr LJ, Roy MJ. Hepatitis B DNA vaccine induces protective antibody responses in human non-responders to conventional vaccination. Vaccine 2003; 21:4604-8; PMID:14575774; https://doi.org/ 10.1016/S0264-410X(03)00447-X [DOI] [PubMed] [Google Scholar]
  • [24].Choi AH, Smiley K, Basu M, McNeal MM, Shao M, Bean JA, Clements JD, Stout RR, Ward RL. Protection of mice against rotavirus challenge following intradermal DNA immunization by Biojector needle-free injection. Vaccine 2007; 25:3215-8; PMID:17280754; https://doi.org/ 10.1016/j.vaccine.2007.01.035 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [25].Ledgerwood JE, Hu Z, Gordon IJ, Yamshchikov G, Enama ME, Plummer S, Bailer R, Pearce MB, Tumpey TM, Koup RA, et al.. Influenza virus h5 DNA vaccination is immunogenic by intramuscular and intradermal routes in humans. Clin Vaccine Immunol 2012; 19:1792-7; PMID:22956656; https://doi.org/ 10.1128/CVI.05663-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [26].Aguiar JC, Hedstrom RC, Rogers WO, Charoenvit Y, Sacci JB Jr, Lanar DE, Majam VF, Stout RR, Hoffman SL. Enhancement of the immune response in rabbits to a malaria DNA vaccine by immunization with a needle-free jet device. Vaccine 2001; 20:275-80; PMID:11567774; https://doi.org/ 10.1016/S0264-410X(01)00273-0 [DOI] [PubMed] [Google Scholar]
  • [27].Trimble C, Lin CT, Hung CF, Pai S, Juang J, He L, Gillison M, Pardoll D, Wu L, Wu TC. Comparison of the CD8+ T cell responses and antitumor effects generated by DNA vaccine administered through gene gun, biojector, and syringe. Vaccine 2003; 21:4036-42; PMID:12922140; https://doi.org/ 10.1016/S0264-410X(03)00275-5 [DOI] [PubMed] [Google Scholar]
  • [28].Karkada M, Weir GM, Quinton T, Fuentes-Ortega A, Mansour M. A liposome-based platform, VacciMax, and its modified water-free platform DepoVax enhance efficacy of in vivo nucleic acid delivery. Vaccine 2010; 28:6176-82; PMID:20656034; https://doi.org/ 10.1016/j.vaccine.2010.07.025 [DOI] [PubMed] [Google Scholar]
  • [29].Fries LF, Gordon DM, Richards RL, Egan JE, Hollingdale MR, Gross M, Silverman C, Alving CR. Liposomal malaria vaccine in humans: a safe and potent adjuvant strategy. Proc Natl Acad Sci U S A 1992; 89:358-62; PMID:1729706; https://doi.org/ 10.1073/pnas.89.1.358 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [30].Zollinger WD, Babcock JG, Moran EE, Brandt BL, Matyas GR, Wassef NM, Alving CR. Phase I study of a Neisseria meningitidis liposomal vaccine containing purified outer membrane proteins and detoxified lipooligosaccharide. Vaccine 2012; 30:712-21; PMID:22138211; https://doi.org/ 10.1016/j.vaccine.2011.11.084 [DOI] [PubMed] [Google Scholar]
  • [31].Hartikka J, Bozoukova V, Ferrari M, Sukhu L, Enas J, Sawdey M, Wloch MK, Tonsky K, Norman J, Manthorpe M, et al.. Vaxfectin enhances the humoral immune response to plasmid DNA-encoded antigens. Vaccine 2001; 19:1911-23; PMID:11228361; https://doi.org/ 10.1016/S0264-410X(00)00445-X [DOI] [PubMed] [Google Scholar]
  • [32].Reyes L, Hartikka J, Bozoukova V, Sukhu L, Nishioka W, Singh G, Ferrari M, Enas J, Wheeler CJ, Manthorpe M, et al.. Vaxfectin enhances antigen specific antibody titers and maintains Th1 type immune responses to plasmid DNA immunization. Vaccine 2001; 19:3778-86; PMID:11395213; https://doi.org/ 10.1016/S0264-410X(01)00090-1 [DOI] [PubMed] [Google Scholar]
  • [33].Sedegah M, Rogers WO, Belmonte A, Belmonte M, Banania G, Patterson N, Ferrari M, Kaslow DC, Carucci DJ, Richie TL, et al.. Vaxfectin enhances immunogenicity and protective efficacy of P. yoelii circumsporozoite DNA vaccines. Vaccine 2006; 24:1921-7 [DOI] [PubMed] [Google Scholar]
  • [34].Sedegah M, Rogers WO, Belmonte M, Belmonte A, Banania G, Patterson NB, Rusalov D, Ferrari M, Richie TL, Doolan DL. Vaxfectin enhances both antibody and in vitro T cell responses to each component of a 5 gene Plasmodium falciparum plasmid DNA vaccine mixture administered at low doses. Vaccine 2010; 28:3055-65; PMID:19879998; https://doi.org/ 10.1016/j.vaccine.2009.10.044 [DOI] [PubMed] [Google Scholar]
  • [35].Smith LR, Wloch MK, Ye M, Reyes LR, Boutsaboualoy S, Dunne CE, Chaplin JA, Rusalov D, Rolland AP, Fisher CL, et al.. Phase 1 clinical trials of the safety and immunogenicity of adjuvanted plasmid DNA vaccines encoding influenza A virus H5 hemagglutinin. Vaccine 2010; 28:2565-72; PMID:20117262; https://doi.org/ 10.1016/j.vaccine.2010.01.029 [DOI] [PubMed] [Google Scholar]
  • [36].Liu J, Wu J, Wang B, Zeng S, Qi F, Lu C, Kimura Y, Liu B. Oral vaccination with a liposome-encapsulated influenza DNA vaccine protects mice against respiratory challenge infection. J Med Virol 2014; 86:886-94; PMID:24122866; https://doi.org/ 10.1002/jmv.23768 [DOI] [PubMed] [Google Scholar]
  • [37].Ma J, Wang H, Zheng X, Xue X, Wang B, Wu H, Zhang K, Fan S, Wang T, Li N, et al.. CpG/Poly (I:C) mixed adjuvant priming enhances the immunogenicity of a DNA vaccine against eastern equine encephalitis virus in mice. Int Immunopharmacol 2014; 19:74-80; PMID:24440303; https://doi.org/ 10.1016/j.intimp.2014.01.002 [DOI] [PubMed] [Google Scholar]
  • [38].Chen L, Zhu J, Li Y, Lu J, Gao L, Xu H, Fan M, Yang X. Enhanced nasal mucosal delivery and immunogenicity of anti-caries DNA vaccine through incorporation of anionic liposomes in chitosan/DNA complexes. PLoS One 2013; 8:e71953; PMID:23977186; https://doi.org/ 10.1371/journal.pone.0071953 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [39].Basarkar A, Devineni D, Palaniappan R, Singh J. Preparation, characterization, cytotoxicity and transfection efficiency of poly(DL-lactide-co-glycolide) and poly(DL-lactic acid) cationic nanoparticles for controlled delivery of plasmid DNA. Int J Pharm 2007; 343:247-54; PMID:17611054; https://doi.org/ 10.1016/j.ijpharm.2007.05.023 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [40].Mok H, Park TG. Direct plasmid DNA encapsulation within PLGA nanospheres by single oil-in-water emulsion method. Eur J Pharm Biopharm 2008; 68:105-11; PMID:17870446; https://doi.org/ 10.1016/j.ejpb.2007.04.022 [DOI] [PubMed] [Google Scholar]
  • [41].Hao T, McKeever U, Hedley ML. Biological potency of microsphere encapsulated plasmid DNA. J Control Release 2000; 69:249-59; PMID:11064132; https://doi.org/ 10.1016/S0168-3659(00)00304-7 [DOI] [PubMed] [Google Scholar]
  • [42].Singh M, Briones M, Ott G, O'Hagan D. Cationic microparticles: A potent delivery system for DNA vaccines. Proc Natl Acad Sci U S A 2000; 97:811-6; PMID:10639162; https://doi.org/ 10.1073/pnas.97.2.811 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [43].Aral C, Akbuga J. Preparation and in vitro transfection efficiency of chitosan microspheres containing plasmid DNA:poly(L-lysine) complexes. J Pharm Pharm Sci 2003; 6:321-6; PMID:14738712 [PubMed] [Google Scholar]
  • [44].Alexakis T, Boadi DK, Quong D, Groboillot A, O'Neill I, Poncelet D, Neufeld RJ. Microencapsulation of DNA within alginate microspheres and crosslinked chitosan membranes for in vivo application. Appl Biochem Biotechnol 1995; 50:93-106; PMID:7702366; https://doi.org/ 10.1007/BF02788043 [DOI] [PubMed] [Google Scholar]
  • [45].Herrmann JE, Chen SC, Jones DH, Tinsley-Bown A, Fynan EF, Greenberg HB, Farrar GH. Immune responses and protection obtained by oral immunization with rotavirus VP4 and VP7 DNA vaccines encapsulated in microparticles. Virology 1999; 259:148-53; PMID:10364499; https://doi.org/ 10.1006/viro.1999.9751 [DOI] [PubMed] [Google Scholar]
  • [46].Kaur R, Rauthan M, Vrati S. Immunogenicity in mice of a cationic microparticle-adsorbed plasmid DNA encoding Japanese encephalitis virus envelope protein. Vaccine 2004; 22:2776-82; PMID:15246611; https://doi.org/ 10.1016/j.vaccine.2004.01.040 [DOI] [PubMed] [Google Scholar]
  • [47].Otten GR, Schaefer M, Doe B, Liu H, Srivastava I, Megede J, Kazzaz J, Lian Y, Singh M, Ugozzoli M, et al.. Enhanced potency of plasmid DNA microparticle human immunodeficiency virus vaccines in rhesus macaques by using a priming-boosting regimen with recombinant proteins. J Virol 2005; 79:8189-200; PMID:15956564; https://doi.org/ 10.1128/JVI.79.13.8189-8200.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [48].Wang F, He XW, Jiang L, Ren D, He Y, Li DA, Sun SH. Enhanced immunogenicity of microencapsulated multiepitope DNA vaccine encoding T and B cell epitopes of foot-and-mouth disease virus in mice. Vaccine 2006; 24:2017-26; PMID:16414158; https://doi.org/ 10.1016/j.vaccine.2005.11.042 [DOI] [PubMed] [Google Scholar]
  • [49].Minigo G, Scholzen A, Tang CK, Hanley JC, Kalkanidis M, Pietersz GA, Apostolopoulos V, Plebanski M. Poly-L-lysine-coated nanoparticles: a potent delivery system to enhance DNA vaccine efficacy. Vaccine 2007; 25:1316-27; PMID:17052812; https://doi.org/ 10.1016/j.vaccine.2006.09.086 [DOI] [PubMed] [Google Scholar]
  • [50].He X, Jiang L, Wang F, Xiao Z, Li J, Liu LS, Li D, Ren D, Jin X, Li K, et al.. Augmented humoral and cellular immune responses to hepatitis B DNA vaccine adsorbed onto cationic microparticles. J Control Release 2005; 107:357-72; PMID:16099068; https://doi.org/ 10.1016/j.jconrel.2005.06.020 [DOI] [PubMed] [Google Scholar]
  • [51].Mollenkopf HJ, Dietrich G, Fensterle J, Grode L, Diehl KD, Knapp B, Singh M, O'Hagan DT, Ulmer JB, Kaufmann SH. Enhanced protective efficacy of a tuberculosis DNA vaccine by adsorption onto cationic PLG microparticles. Vaccine 2004; 22:2690-5; PMID:15309815; https://doi.org/ 10.1016/j.vaccine.2004.05.005 [DOI] [PubMed] [Google Scholar]
  • [52].Luo Y, O'Hagan D, Zhou H, Singh M, Ulmer J, Reisfeld RA, James Primus F, Xiang R. Plasmid DNA encoding human carcinoembryonic antigen (CEA) adsorbed onto cationic microparticles induces protective immunity against colon cancer in CEA-transgenic mice. Vaccine 2003; 21:1938-47; PMID:12706680; https://doi.org/ 10.1016/S0264-410X(02)00821-6 [DOI] [PubMed] [Google Scholar]
  • [53].Ishii KJ, Kawagoe T, Koyama S, Matsui K, Kumar H, Kawai T, Uematsu S, Takeuchi O, Takeshita F, Coban C, et al.. TANK-binding kinase-1 delineates innate and adaptive immune responses to DNA vaccines. Nature 2008; 451:725-9; PMID:18256672; https://doi.org/ 10.1038/nature06537 [DOI] [PubMed] [Google Scholar]
  • [54].Okabe Y, Kawane K, Akira S, Taniguchi T, Nagata S. Toll-like receptor-independent gene induction program activated by mammalian DNA escaped from apoptotic DNA degradation. J Exp Med 2005; 202:1333-9; PMID:16301743; https://doi.org/ 10.1084/jem.20051654 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [55].Stetson DB, Medzhitov R. Recognition of cytosolic DNA activates an IRF3-dependent innate immune response. Immunity 2006; 24:93-103; PMID:16413926; https://doi.org/ 10.1016/j.immuni.2005.12.003 [DOI] [PubMed] [Google Scholar]
  • [56].Spies B, Hochrein H, Vabulas M, Huster K, Busch DH, Schmitz F, Heit A, Wagner H. Vaccination with plasmid DNA activates dendritic cells via Toll-like receptor 9 (TLR9) but functions in TLR9-deficient mice. J Immunol 2003; 171:5908-12; PMID:14634101; https://doi.org/ 10.4049/jimmunol.171.11.5908 [DOI] [PubMed] [Google Scholar]
  • [57].Babiuk S, Mookherjee N, Pontarollo R, Griebel P, van Drunen Littel-van den Hurk S, Hecker R, Babiuk L. TLR9−/− and TLR9+/+ mice display similar immune responses to a DNA vaccine. Immunology 2004; 113:114-20; PMID:15312142; https://doi.org/ 10.1111/j.1365-2567.2004.01938.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [58].Kojima Y, Xin KQ, Ooki T, Hamajima K, Oikawa T, Shinoda K, Ozaki T, Hoshino Y, Jounai N, Nakazawa M, Klinman D, et al.. Adjuvant effect of multi-CpG motifs on an HIV-1 DNA vaccine. Vaccine 2002; 20:2857-65; PMID:12126895; https://doi.org/ 10.1016/S0264-410X(02)00238-4 [DOI] [PubMed] [Google Scholar]
  • [59].Coban C, Ishii KJ, Gursel M, Klinman DM, Kumar N. Effect of plasmid backbone modification by different human CpG motifs on the immunogenicity of DNA vaccine vectors. J Leukoc Biol 2005; 78:647-55; PMID:15961575; https://doi.org/ 10.1189/jlb.1104627 [DOI] [PubMed] [Google Scholar]
  • [60].Abdulhaqq SA, Weiner DB. DNA vaccines: developing new strategies to enhance immune responses. Immunol Res 2008; 42:219-32; PMID:19066740; https://doi.org/ 10.1007/s12026-008-8076-3 [DOI] [PubMed] [Google Scholar]
  • [61].Laddy DJ, Weiner DB. From plasmids to protection: a review of DNA vaccines against infectious diseases. Int Rev Immunol 2006; 25:99-123; PMID:16818367; https://doi.org/ 10.1080/08830180600785827 [DOI] [PubMed] [Google Scholar]
  • [62].Flingai S, Czerwonko M, Goodman J, Kudchodkar SB, Muthumani K, Weiner DB. Synthetic DNA vaccines: improved vaccine potency by electroporation and co-delivered genetic adjuvants. Front Immunol 2013; 4:354; PMID:24204366; https://doi.org/ 10.3389/fimmu.2013.00354 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [63].Kim JJ, Nottingham LK, Wilson DM, Bagarazzi ML, Tsai A, Morrison LD, Javadian A, Chalian AA, Agadjanyan MG, Weiner DB. Engineering DNA vaccines via co-delivery of co-stimulatory molecule genes. Vaccine 1998; 16:1828-35; PMID:9795388; https://doi.org/ 10.1016/S0264-410X(98)00177-7 [DOI] [PubMed] [Google Scholar]
  • [64].Kim JJ, Simbiri KA, Sin JI, Dang K, Oh J, Dentchev T, Lee D, Nottingham LK, Chalian AA, McCallus D, et al.. Cytokine molecular adjuvants modulate immune responses induced by DNA vaccine constructs for HIV-1 and SIV. J Interferon Cytokine Res 1999; 19:77-84; PMID:10048771; https://doi.org/ 10.1089/107999099314441 [DOI] [PubMed] [Google Scholar]
  • [65].Barouch DH, Truitt DM, Letvin NL. Expression kinetics of the interleukin-2/immunoglobulin (IL-2/Ig) plasmid cytokine adjuvant. Vaccine 2004; 22:3092-7; PMID:15297060; https://doi.org/ 10.1016/j.vaccine.2004.01.065 [DOI] [PubMed] [Google Scholar]
  • [66].Henke A, Rohland N, Zell R, Wutzler P. Co-expression of interleukin-2 by a bicistronic plasmid increases the efficacy of DNA immunization to prevent influenza virus infections. Intervirology 2006; 49:249-52; PMID:16601357; https://doi.org/ 10.1159/000092487 [DOI] [PubMed] [Google Scholar]
  • [67].Hu H, Tao L, Wang Y, Chen L, Yang J, Wang H. Enhancing immune responses against SARS-CoV nucleocapsid DNA vaccine by co-inoculating interleukin-2 expressing vector in mice. Biotechnol Lett 2009; 31:1685-93; PMID:19579009; https://doi.org/ 10.1007/s10529-009-0061-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [68].Qin Y, Tian H, Wang G, Lin C, Li Y. A BCR/ABL-hIL-2 DNA vaccine enhances the immune responses in BALB/c mice. Biomed Res Int 2013; 2013:136492; PMID:23841051; https://doi.org/ 10.1155/2013/136492 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [69].McGill J, Van Rooijen N, Legge KL. IL-15 trans-presentation by pulmonary dendritic cells promotes effector CD8 T cell survival during influenza virus infection. J Exp Med 2010; 207:521-34; PMID:20212069; https://doi.org/ 10.1084/jem.20091711 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [70].Schluns KS, Williams K, Ma A, Zheng XX, Lefrancois L. Cutting edge: requirement for IL-15 in the generation of primary and memory antigen-specific CD8 T cells. J Immunol 2002; 168:4827-31; PMID:11994430; https://doi.org/ 10.4049/jimmunol.168.10.4827 [DOI] [PubMed] [Google Scholar]
  • [71].Yajima T, Yoshihara K, Nakazato K, Kumabe S, Koyasu S, Sad S, Shen H, Kuwano H, Yoshikai Y. IL-15 regulates CD8+ T cell contraction during primary infection. J Immunol 2006; 176:507-15; PMID:16365444; https://doi.org/ 10.4049/jimmunol.176.1.507 [DOI] [PubMed] [Google Scholar]
  • [72].Combe CL, Moretto MM, Schwartzman JD, Gigley JP, Bzik DJ, Khan IA. Lack of IL-15 results in the suboptimal priming of CD4+ T cell response against an intracellular parasite. Proc Natl Acad Sci U S A 2006; 103:6635-40; PMID:16614074; https://doi.org/ 10.1073/pnas.0506180103 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [73].Ruckert R, Brandt K, Bulanova E, Mirghomizadeh F, Paus R, Bulfone-Paus S. Dendritic cell-derived IL-15 controls the induction of CD8 T cell immune responses. Eur J Immunol 2003; 33:3493-503; PMID:14635060; https://doi.org/ 10.1002/eji.200324545 [DOI] [PubMed] [Google Scholar]
  • [74].Chen J, Li ZY, Huang SY, Petersen E, Song HQ, Zhou DH, Zhu XQ. Protective efficacy of Toxoplasma gondii calcium-dependent protein kinase 1 (TgCDPK1) adjuvated with recombinant IL-15 and IL-21 against experimental toxoplasmosis in mice. BMC Infect Dis 2014; 14:487; PMID:25192845; https://doi.org/ 10.1186/1471-2334-14-487 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [75].Li ZY, Chen J, Petersen E, Zhou DH, Huang SY, Song HQ, Zhu XQ. Synergy of mIL-21 and mIL-15 in enhancing DNA vaccine efficacy against acute and chronic Toxoplasma gondii infection in mice. Vaccine 2014; 32:3058-65; PMID:24690150; https://doi.org/ 10.1016/j.vaccine.2014.03.042 [DOI] [PubMed] [Google Scholar]
  • [76].Su B, Wang J, Zhao G, Wang X, Li J, Wang B. Sequential administration of cytokine genes to enhance cellular immune responses and CD4 (+) T memory cells during DNA vaccination. Hum Vaccin Immunother 2012; 8:1659-67; PMID:23151452; https://doi.org/ 10.4161/hv.22105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [77].Bergamaschi C, Kulkarni V, Rosati M, Alicea C, Jalah R, Chen S, Bear J, Sardesai NY, Valentin A, Felber BK, et al.. Intramuscular delivery of heterodimeric IL-15 DNA in macaques produces systemic levels of bioactive cytokine inducing proliferation of NK and T cells. Gene Ther 2015; 22:76-86; PMID:25273353; https://doi.org/ 10.1038/gt.2014.84 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [78].Boyer JD, Robinson TM, Kutzler MA, Vansant G, Hokey DA, Kumar S, Parkinson R, Wu L, Sidhu MK, Pavlakis GN, et al.. Protection against simian/human immunodeficiency virus (SHIV) 89.6P in macaques after coimmunization with SHIV antigen and IL-15 plasmid. Proc Natl Acad Sci U S A 2007; 104:18648-53; PMID:18000037; https://doi.org/ 10.1073/pnas.0709198104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [79].Hsieh CS, Macatonia SE, Tripp CS, Wolf SF, O'Garra A, Murphy KM. Development of TH1 CD4+ T cells through IL-12 produced by Listeria-induced macrophages. Science 1993; 260:547-9; PMID:8097338; https://doi.org/ 10.1126/science.8097338 [DOI] [PubMed] [Google Scholar]
  • [80].Sypek JP, Chung CL, Mayor SE, Subramanyam JM, Goldman SJ, Sieburth DS, Wolf SF, Schaub RG. Resolution of cutaneous leishmaniasis: interleukin 12 initiates a protective T helper type 1 immune response. J Exp Med 1993; 177:1797-802; PMID:8098733; https://doi.org/ 10.1084/jem.177.6.1797 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [81].Macatonia SE, Hosken NA, Litton M, Vieira P, Hsieh CS, Culpepper JA, Wysocka M, Trinchieri G, Murphy KM, O'Garra A. Dendritic cells produce IL-12 and direct the development of Th1 cells from naive CD4+ T cells. J Immunol 1995; 154:5071-9; PMID:7730613 [PubMed] [Google Scholar]
  • [82].Kobayashi M, Fitz L, Ryan M, Hewick RM, Clark SC, Chan S, Loudon R, Sherman F, Perussia B, Trinchieri G. Identification and purification of natural killer cell stimulatory factor (NKSF), a cytokine with multiple biologic effects on human lymphocytes. J Exp Med 1989; 170:827-45; PMID:2504877; https://doi.org/ 10.1084/jem.170.3.827 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [83].Trinchieri G. Interleukin-12: a cytokine at the interface of inflammation and immunity. Adv Immunol 1998; 70:83-243; PMID:9755338 [DOI] [PubMed] [Google Scholar]
  • [84].Hunter CA, Chizzonite R, Remington JS. IL-1 beta is required for IL-12 to induce production of IFN-gamma by NK cells. A role for IL-1 beta in the T cell-independent mechanism of resistance against intracellular pathogens. J Immunol 1995; 155:4347-54; PMID:7594594 [PubMed] [Google Scholar]
  • [85].Chan SH, Kobayashi M, Santoli D, Perussia B, Trinchieri G. Mechanisms of IFN-gamma induction by natural killer cell stimulatory factor (NKSF/IL-12). Role of transcription and mRNA stability in the synergistic interaction between NKSF and IL-2. J Immunol 1992; 148:92-8; PMID:1345792 [PubMed] [Google Scholar]
  • [86].Manetti R, Gerosa F, Giudizi MG, Biagiotti R, Parronchi P, Piccinni MP, Sampognaro S, Maggi E, Romagnani S, Trinchieri G, et al.. Interleukin 12 induces stable priming for interferon gamma (IFN-gamma) production during differentiation of human T helper (Th) cells and transient IFN-gamma production in established Th2 cell clones. J Exp Med 1994; 179:1273-83; PMID:7908322; https://doi.org/ 10.1084/jem.179.4.1273 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [87].Kalams SA, Parker SD, Elizaga M, Metch B, Edupuganti S, Hural J, De Rosa S, Carter DK, Rybczyk K, Frank I, et al.. Safety and comparative immunogenicity of an HIV-1 DNA vaccine in combination with plasmid interleukin 12 and impact of intramuscular electroporation for delivery. J Infect Dis 2013; 208:818-29; PMID:23840043; https://doi.org/ 10.1093/infdis/jit236 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [88].Kim JJ, Ayyavoo V, Bagarazzi ML, Chattergoon M, Boyer JD, Wang B, Weiner DB. Development of a multicomponent candidate vaccine for HIV-1. Vaccine 1997; 15:879-83; PMID:9234538; https://doi.org/ 10.1016/S0264-410X(96)00260-5 [DOI] [PubMed] [Google Scholar]
  • [89].Hirao LA, Wu L, Khan AS, Hokey DA, Yan J, Dai A, Betts MR, Draghia-Akli R, Weiner DB. Combined effects of IL-12 and electroporation enhances the potency of DNA vaccination in macaques. Vaccine 2008; 26:3112-20; PMID:18430495; https://doi.org/ 10.1016/j.vaccine.2008.02.036 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [90].Boyer JD, Robinson TM, Kutzler MA, Parkinson R, Calarota SA, Sidhu MK, Muthumani K, Lewis M, Pavlakis G, Felber B, et al.. SIV DNA vaccine co-administered with IL-12 expression plasmid enhances CD8 SIV cellular immune responses in cynomolgus macaques. J Med Primatol 2005; 34:262-70; PMID:16128921; https://doi.org/ 10.1111/j.1600-0684.2005.00124.x [DOI] [PubMed] [Google Scholar]
  • [91].Chong SY, Egan MA, Kutzler MA, Megati S, Masood A, Roopchard V, Garcia-Hand D, Montefiori DC, Quiroz J, Rosati M, et al.. Comparative ability of plasmid IL-12 and IL-15 to enhance cellular and humoral immune responses elicited by a SIVgag plasmid DNA vaccine and alter disease progression following SHIV(89.6P) challenge in rhesus macaques. Vaccine 2007; 25:4967-82; PMID:17335943; https://doi.org/ 10.1016/j.vaccine.2006.11.070 [DOI] [PubMed] [Google Scholar]
  • [92].Robinson TM, Sidhu MK, Pavlakis GN, Felber BK, Silvera P, Lewis MG, Eldridge J, Weiner DB, Boyer JD. Macaques co-immunized with SIVgag/pol-HIVenv and IL-12 plasmid have increased cellular responses. J Med Primatol 2007; 36:276-84; PMID:17669216; https://doi.org/ 10.1111/j.1600-0684.2007.00245.x [DOI] [PubMed] [Google Scholar]
  • [93].Bagley KC, Schwartz JA, Andersen H, Eldridge JH, Xu R, Ota-Setlik A, Geltz JJ, Halford WP, Fouts TR. An interleukin 12 adjuvanted herpes simplex virus 2 DNA vaccine is more protective than a glycoprotein D subunit vaccine in a high-dose murine challenge model. Viral Immunol. 2017; 30:178-95; PMID:28085634; https://doi.org/ 10.1089/vim.2016.0136 [DOI] [PubMed] [Google Scholar]
  • [94].Morrow MP, Yan J, Pankhong P, Ferraro B, Lewis MG, Khan AS, Sardesai NY, Weiner DB. Unique Th1/Th2 phenotypes induced during priming and memory phases by use of interleukin-12 (IL-12) or IL-28B vaccine adjuvants in rhesus macaques. Clin Vaccine Immunol 2010; 17:1493-9; https://doi.org/ 10.1128/CVI.00181-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [95].Bhaumik S, Basu R, Sen S, Naskar K, Roy S. KMP-11 DNA immunization significantly protects against L. donovani infection but requires exogenous IL-12 as an adjuvant for comparable protection against L. major. Vaccine 2009; 27:1306-16; https://doi.org/ 10.1016/j.vaccine.2008.12.053 [DOI] [PubMed] [Google Scholar]
  • [96].Yamanaka H, Hoyt T, Yang X, Golden S, Bosio CM, Crist K, Becker T, Maddaloni M, Pascual DW. A nasal interleukin-12 DNA vaccine coexpressing Yersinia pestis F1-V fusion protein confers protection against pneumonic plague. Infect Immun 2008; 76:4564-73; PMID:18694965; https://doi.org/ 10.1128/IAI.00581-08 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [97].Schadeck EB, Sidhu M, Egan MA, Chong SY, Piacente P, Masood A, Garcia-Hand D, Cappello S, Roopchand V, Megati S, et al.. A dose sparing effect by plasmid encoded IL-12 adjuvant on a SIVgag-plasmid DNA vaccine in rhesus macaques. Vaccine 2006; 24:4677-87; PMID:16288822; https://doi.org/ 10.1016/j.vaccine.2005.10.035 [DOI] [PubMed] [Google Scholar]
  • [98].Halwani R, Boyer JD, Yassine-Diab B, Haddad EK, Robinson TM, Kumar S, Parkinson R, Wu L, Sidhu MK, Phillipson-Weiner R, et al.. Therapeutic vaccination with simian immunodeficiency virus (SIV)-DNA + IL-12 or IL-15 induces distinct CD8 memory subsets in SIV-infected macaques. J Immunol 2008; 180:7969-79; PMID:18523260; https://doi.org/ 10.4049/jimmunol.180.12.7969 [DOI] [PubMed] [Google Scholar]
  • [99].Li J, Valentin A, Kulkarni V, Rosati M, Beach RK, Alicea C, Hannaman D, Reed SG, Felber BK, Pavlakis GN. HIV/SIV DNA vaccine combined with protein in a co-immunization protocol elicits highest humoral responses to envelope in mice and macaques. Vaccine 2013; 31:3747-55; PMID:23624057; https://doi.org/ 10.1016/j.vaccine.2013.04.037 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [100].Cha E, Daud A. Plasmid IL-12 electroporation in melanoma. Hum Vaccin Immunother 2012; 8:1734-8; PMID:23151447; https://doi.org/ 10.4161/hv.22573 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [101].Kalams SA, Goulder PJ, Shea AK, Jones NG, Trocha AK, Ogg GS, Walker BD. Levels of human immunodeficiency virus type 1-specific cytotoxic T-lymphocyte effector and memory responses decline after suppression of viremia with highly active antiretroviral therapy. J Virol 1999; 73:6721-8; PMID:10400770 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [102].Kalams SA, Parker S, Jin X, Elizaga M, Metch B, Wang M, Hural J, Lubeck M, Eldridge J, Cardinali M, et al.. Safety and immunogenicity of an HIV-1 gag DNA vaccine with or without IL-12 and/or IL-15 plasmid cytokine adjuvant in healthy, HIV-1 uninfected adults. PLoS One 2012; 7:e29231; PMID:22242162; https://doi.org/ 10.1371/journal.pone.0029231 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [103].Yang SH, Lee CG, Park SH, Im SJ, Kim YM, Son JM, Wang JS, Yoon SK, Song MK, Ambrozaitis A, et al.. Correlation of antiviral T-cell responses with suppression of viral rebound in chronic hepatitis B carriers: a proof-of-concept study. Gene Ther 2006; 13:1110-7; PMID:16525482 [DOI] [PubMed] [Google Scholar]
  • [104].Haddad D, Ramprakash J, Sedegah M, Charoenvit Y, Baumgartner R, Kumar S, Hoffman SL, Weiss WR. Plasmid vaccine expressing granulocyte-macrophage colony-stimulating factor attracts infiltrates including immature dendritic cells into injected muscles. J Immunol 2000; 165:3772-81; PMID:11034382 [DOI] [PubMed] [Google Scholar]
  • [105].Weiss WR, Ishii KJ, Hedstrom RC, Sedegah M, Ichino M, Barnhart K, Klinman DM, Hoffman SL. A plasmid encoding murine granulocyte-macrophage colony-stimulating factor increases protection conferred by a malaria DNA vaccine. J Immunol 1998; 161:2325-32; PMID:9725227 [PubMed] [Google Scholar]
  • [106].Ahlers JD, Belyakov IM, Terabe M, Koka R, Donaldson DD, Thomas EK, Berzofsky JA. A push-pull approach to maximize vaccine efficacy: abrogating suppression with an IL-13 inhibitor while augmenting help with granulocyte/macrophage colony-stimulating factor and CD40L. Proc Natl Acad Sci U S A 2002; 99:13020-5; PMID:12232042 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [107].Yoon HA, Aleyas AG, George JA, Park SO, Han YW, Lee JH, Cho JG, Eo SK. Cytokine GM-CSF genetic adjuvant facilitates prophylactic DNA vaccine against pseudorabies virus through enhanced immune responses. Microbiol Immunol 2006; 50:83-92; PMID:16490926 [DOI] [PubMed] [Google Scholar]
  • [108].Xiang Z, Ertl HC. Manipulation of the immune response to a plasmid-encoded viral antigen by coinoculation with plasmids expressing cytokines. Immunity 1995; 2:129-35; PMID:7895169 [DOI] [PubMed] [Google Scholar]
  • [109].Santana VC, Almeida RR, Ribeiro SP, Ferreira LC, Kalil J, Rosa DS, Cunha-Neto E. Co-administration of plasmid-encoded granulocyte-macrophage colony-stimulating factor increases human immunodeficiency virus-1 DNA vaccine-induced polyfunctional CD4+ T-cell responses. Mem Inst Oswaldo Cruz 2015; 110:1010-6; PMID:26602876 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [110].Chen H, Gao N, Wu J, Zheng X, Li J, Fan D, An J. Variable effects of the co-administration of a GM-CSF-expressing plasmid on the immune response to flavivirus DNA vaccines in mice. Immunol Lett 2014; 162:140-8; PMID:25128840 [DOI] [PubMed] [Google Scholar]
  • [111].Kumar S, Villinger F, Oakley M, Aguiar JC, Jones TR, Hedstrom RC, Gowda K, Chute J, Stowers A, Kaslow DC, et al.. A DNA vaccine encoding the 42 kDa C-terminus of merozoite surface protein 1 of Plasmodium falciparum induces antibody, interferon-gamma and cytotoxic T cell responses in rhesus monkeys: immuno-stimulatory effects of granulocyte macrophage-colony stimulating factor. Immunol Lett 2002; 81:13-24; PMID:11841841 [DOI] [PubMed] [Google Scholar]
  • [112].Cassaday RD, Sondel PM, King DM, Macklin MD, Gan J, Warner TF, Zuleger CL, Bridges AJ, Schalch HG, Kim KM, et al.. A phase I study of immunization using particle-mediated epidermal delivery of genes for gp100 and GM-CSF into uninvolved skin of melanoma patients. Clin Cancer Res 2007; 13:540-9; PMID:17255276; https://doi.org/ 10.1158/1078-0432.CCR-06-2039 [DOI] [PubMed] [Google Scholar]
  • [113].Barouch DH, Letvin NL, Seder RA. The role of cytokine DNAs as vaccine adjuvants for optimizing cellular immune responses. Immunol Rev 2004; 202:266-74; PMID:15546399; https://doi.org/ 10.1111/j.0105-2896.2004.00200.x [DOI] [PubMed] [Google Scholar]
  • [114].Wan C, Yi L, Yang Z, Yang J, Shao H, Zhang C, Pan Z. The Toll-like receptor adaptor molecule TRIF enhances DNA vaccination against classical swine fever. Vet Immunol Immunopathol 2010; 137:47-53; PMID:20466439; https://doi.org/ 10.1016/j.vetimm.2010.04.008 [DOI] [PubMed] [Google Scholar]
  • [115].Takeshita F, Tanaka T, Matsuda T, Tozuka M, Kobiyama K, Saha S, Matsui K, Ishii KJ, Coban C, Akira S, et al.. Toll-like receptor adaptor molecules enhance DNA-raised adaptive immune responses against influenza and tumors through activation of innate immunity. J Virol 2006; 80:6218-24; PMID:16775309; https://doi.org/ 10.1128/JVI.00121-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [116].Liniger M, Summerfield A, Ruggli N. MDA5 can be exploited as efficacious genetic adjuvant for DNA vaccination against lethal H5N1 influenza virus infection in chickens. PLoS One 2012; 7:e49952; PMID:23227156; https://doi.org/ 10.1371/journal.pone.0049952 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [117].Luke JM, Simon GG, Soderholm J, Errett JS, August JT, Gale M Jr., Hodgson CP, Williams JA. Coexpressed RIG-I agonist enhances humoral immune response to influenza virus DNA vaccine. J Virol 2011; 85:1370-83; PMID:21106745; https://doi.org/ 10.1128/JVI.01250-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [118].Yang B, Yang A, Peng S, Pang X, Roden RB, Wu TC, Hung CF. Co-administration with DNA encoding papillomavirus capsid proteins enhances the antitumor effects generated by therapeutic HPV DNA vaccination. Cell Biosci 2015; 5:35; PMID:26113972; https://doi.org/ 10.1186/s13578-015-0025-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [119].Massa S, Paolini F, Spano L, Franconi R, Venuti A. Mutants of plant genes for developing cancer vaccines. Hum Vaccin 2011; 7 Suppl:147-55; PMID:21266841; https://doi.org/ 10.4161/hv.7.0.14577 [DOI] [PubMed] [Google Scholar]
  • [120].Kim JW, Hung CF, Juang J, He L, Kim TW, Armstrong DK, Pai SI, Chen PJ, Lin CT, Boyd DA, et al.. Comparison of HPV DNA vaccines employing intracellular targeting strategies. Gene Ther 2004; 11:1011-8; PMID:14985791; https://doi.org/ 10.1038/sj.gt.3302252 [DOI] [PubMed] [Google Scholar]
  • [121].Manam S, Ledwith BJ, Barnum AB, Troilo PJ, Pauley CJ, Harper LB, Griffiths TG 2nd, Niu Z, Denisova L, Follmer TT, et al.. Plasmid DNA vaccines: tissue distribution and effects of DNA sequence, adjuvants and delivery method on integration into host DNA. Intervirology 2000; 43:273-81; PMID:11251382; https://doi.org/ 10.1159/000053994 [DOI] [PubMed] [Google Scholar]
  • [122].Nichols WW, Ledwith BJ, Manam SV, Troilo PJ. Potential DNA vaccine integration into host cell genome. Ann N Y Acad Sci 1995; 772:30-9; PMID:8546411; https://doi.org/ 10.1111/j.1749-6632.1995.tb44729.x [DOI] [PubMed] [Google Scholar]
  • [123].Sorensen SJ, Bailey M, Hansen LH, Kroer N, Wuertz S. Studying plasmid horizontal transfer in situ: a critical review. Nat Rev Microbiol 2005; 3:700-10; PMID:16138098; https://doi.org/ 10.1038/nrmicro1232 [DOI] [PubMed] [Google Scholar]
  • [124].Oliveira PH, Mairhofer J. Marker-free plasmids for biotechnological applications - implications and perspectives. Trends Biotechnol 2013; 31:539-47; PMID:23830144; https://doi.org/ 10.1016/j.tibtech.2013.06.001 [DOI] [PubMed] [Google Scholar]
  • [125].Williams JA. Improving DNA vaccine performance through vector design. Curr Gene Ther 2014; 14:170-89; PMID:25142448; https://doi.org/ 10.2174/156652321403140819122538 [DOI] [PubMed] [Google Scholar]
  • [126].Kreiss P, Cameron B, Rangara R, Mailhe P, Aguerre-Charriol O, Airiau M, Scherman D, Crouzet J, Pitard B. Plasmid DNA size does not affect the physicochemical properties of lipoplexes but modulates gene transfer efficiency. Nucleic Acids Res 1999; 27:3792-8; PMID:10481017; https://doi.org/ 10.1093/nar/27.19.3792 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Yin W, Xiang P, Li Q. Investigations of the effect of DNA size in transient transfection assay using dual luciferase system. Anal Biochem 2005; 346:289-94; PMID:16213455; https://doi.org/ 10.1016/j.ab.2005.08.029 [DOI] [PubMed] [Google Scholar]
  • [128].Hornstein BD, Roman D, Arevalo-Soliz LM, Engevik MA, Zechiedrich L. Effects of circular DNA length on transfection efficiency by electroporation into HeLa Cells. PLoS One 2016; 11:e0167537; PMID:27918590; https://doi.org/ 10.1371/journal.pone.0167537 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [129].Stenler S, Wiklander OP, Badal-Tejedor M, Turunen J, Nordin JZ, Hallengard D, Wahren B, Andaloussi SE, Rutland MW, Smith CE, et al.. Micro-minicircle Gene therapy: Implications of size on fermentation, complexation, shearing resistance, and expression. Mol Ther Nucleic Acids 2014; 2:e140; PMID:24399204; https://doi.org/ 10.1038/mtna.2013.67 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [130].Lukacs GL, Haggie P, Seksek O, Lechardeur D, Freedman N, Verkman AS. Size-dependent DNA mobility in cytoplasm and nucleus. J Biol Chem 2000; 275:1625-9; PMID:10636854; https://doi.org/ 10.1074/jbc.275.3.1625 [DOI] [PubMed] [Google Scholar]
  • [131].Marie C, Vandermeulen G, Quiviger M, Richard M, Preat V, Scherman D. pFARs, plasmids free of antibiotic resistance markers, display high-level transgene expression in muscle, skin and tumour cells. J Gene Med 2010; 12:323-32; PMID:20209487; https://doi.org/ 10.1002/jgm.1441 [DOI] [PubMed] [Google Scholar]
  • [132].Carnes AE, Luke JM, Vincent JM, Anderson S, Schukar A, Hodgson CP, Williams JA. Critical design criteria for minimal antibiotic-free plasmid vectors necessary to combine robust RNA Pol II and Pol III-mediated eukaryotic expression with high bacterial production yields. J Gene Med 2010; 12:818-31; PMID:20806425; https://doi.org/ 10.1002/jgm.1499 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [133].Ribeiro S, Mairhofer J, Madeira C, Diogo MM, Lobato da Silva C, Monteiro G, Grabherr R, Cabral JM. Plasmid DNA size does affect nonviral gene delivery efficiency in stem cells. Cell Reprogram 2012; 14:130-7; PMID:22339198 [DOI] [PubMed] [Google Scholar]
  • [134].Wang Q, Jiang W, Chen Y, Liu P, Sheng C, Chen S, Zhang H, Pan C, Gao S, Huang W. In vivo electroporation of minicircle DNA as a novel method of vaccine delivery to enhance HIV-1-specific immune responses. J Virol 2013; 88:1924-34; PMID:24284319; https://doi.org/ 10.1128/JVI.02757-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [135].Lu J, Zhang F, Xu S, Fire AZ, Kay MA. The extragenic spacer length between the 5′ and 3′ ends of the transgene expression cassette affects transgene silencing from plasmid-based vectors. Mol Ther 2012; 20:2111-9; PMID:22565847; https://doi.org/ 10.1038/mt.2012.65 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [136].Gracey Maniar LE, Maniar JM, Chen ZY, Lu J, Fire AZ, Kay MA. Minicircle DNA vectors achieve sustained expression reflected by active chromatin and transcriptional level. Mol Ther 2013; 21:131-8; PMID:23183534; https://doi.org/ 10.1038/mt.2012.244 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [137].Chen ZY, He CY, Ehrhardt A, Kay MA. Minicircle DNA vectors devoid of bacterial DNA result in persistent and high-level transgene expression in vivo. Mol Ther 2003; 8:495-500; PMID:12946323; https://doi.org/ 10.1016/S1525-0016(03)00168-0 [DOI] [PubMed] [Google Scholar]
  • [138].Simcikova M, Prather KL, Prazeres DM, Monteiro GA. Towards effective non-viral gene delivery vector. Biotechnol Genet Eng Rev 2015; 31:82-107; PMID:27160661; https://doi.org/ 10.1080/02648725.2016.1178011 [DOI] [PubMed] [Google Scholar]
  • [139].Chen ZY, He CY, Kay MA. Improved production and purification of minicircle DNA vector free of plasmid bacterial sequences and capable of persistent transgene expression in vivo. Hum Gene Ther 2005; 16:126-31; PMID:15703495; https://doi.org/ 10.1089/hum.2005.16.126 [DOI] [PubMed] [Google Scholar]
  • [140].Nelson J, Rodriguez S, Finlayson N, Williams J, Carnes A. Antibiotic-free production of a herpes simplex virus 2 DNA vaccine in a high yield cGMP process. Hum Vaccin Immunother 2013; 9:2211-5; PMID:23899469; https://doi.org/ 10.4161/hv.25048 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [141].Lu J, Zhang F, Kay MA. A mini-intronic plasmid (MIP): a novel robust transgene expression vector in vivo and in vitro. Mol Ther 2013; 21:954-63; PMID:23459514; https://doi.org/ 10.1038/mt.2013.33 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [142].Williams JA. Replicative minicircle vectors with improved expression. Nature Technology Corporation, US 2014; PCT/US2013/000259, WO2014/077,866, US 2015/0275221. [Google Scholar]
  • [143].Pandey S, Kawai T, Akira S. Microbial sensing by Toll-like receptors and intracellular nucleic acid sensors. Cold Spring Harb Perspect Biol 2014; 7:a016246; PMID:25301932; https://doi.org/ 10.1101/cshperspect.a016246 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [144].Pavlenko M, Leder C, Moreno S, Levitsky V, Pisa P. Priming of CD8+ T-cell responses after DNA immunization is impaired in TLR9- and MyD88-deficient mice. Vaccine 2007; 25:6341-7; PMID:17628235; https://doi.org/ 10.1016/j.vaccine.2007.06.016 [DOI] [PubMed] [Google Scholar]
  • [145].Rottembourg D, Filippi CM, Bresson D, Ehrhardt K, Estes EA, Oldham JE, von Herrath MG. Essential role for TLR9 in prime but not prime-boost plasmid DNA vaccination to activate dendritic cells and protect from lethal viral infection. J Immunol 2010; 184:7100-7; PMID:20483769; https://doi.org/ 10.4049/jimmunol.0803935 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [146].Ohlschlager P, Spies E, Alvarez G, Quetting M, Groettrup M. The combination of TLR-9 adjuvantation and electroporation-mediated delivery enhances in vivo antitumor responses after vaccination with HPV-16 E7 encoding DNA. Int J Cancer 2011; 128:473-81; PMID:20309939; https://doi.org/ 10.1002/ijc.25344 [DOI] [PubMed] [Google Scholar]
  • [147].Yu YZ, Li N, Ma Y, Wang S, Yu WY, Sun ZW. Three types of human CpG motifs differentially modulate and augment immunogenicity of nonviral and viral replicon DNA vaccines as built-in adjuvants. Eur J Immunol 2013; 43:228-39; PMID:23037552; https://doi.org/ 10.1002/eji.201242690 [DOI] [PubMed] [Google Scholar]
  • [148].Arsenault RJ, Kogut MH, He H. Combined CpG and poly I:C stimulation of monocytes results in unique signaling activation not observed with the individual ligands. Cell Signal 2013; 25:2246-54; PMID:23876795; https://doi.org/ 10.1016/j.cellsig.2013.07.014 [DOI] [PubMed] [Google Scholar]
  • [149].Gutjahr A, Tiraby G, Perouzel E, Verrier B, Paul S. Triggering intracellular receptors for vaccine adjuvantation. Trends Immunol 2016; 37:573-87; PMID:27474233; https://doi.org/ 10.1016/j.it.2016.07.001 [DOI] [PubMed] [Google Scholar]
  • [150].Wu J, Chen ZJ. Innate immune sensing and signaling of cytosolic nucleic acids. Annu Rev Immunol 2014; 32:461-88; PMID:24655297; https://doi.org/ 10.1146/annurev-immunol-032713-120156 [DOI] [PubMed] [Google Scholar]
  • [151].Sugiyama T, Gursel M, Takeshita F, Coban C, Conover J, Kaisho T, Akira S, Klinman DM, Ishii KJ. CpG RNA: identification of novel single-stranded RNA that stimulates human CD14+CD11c+ monocytes. J Immunol 2005; 174:2273-9; PMID:15699162; https://doi.org/ 10.4049/jimmunol.174.4.2273 [DOI] [PubMed] [Google Scholar]
  • [152].Wu J, Ma H, Qu Q, Zhou WJ, Luo YP, Thangaraj H, Lowrie DB, Fan XY. Incorporation of immunostimulatory motifs in the transcribed region of a plasmid DNA vaccine enhances Th1 immune responses and therapeutic effect against Mycobacterium tuberculosis in mice. Vaccine 2011; 29:7624-30; PMID:21856352; https://doi.org/ 10.1016/j.vaccine.2011.08.020 [DOI] [PubMed] [Google Scholar]
  • [153].Hochheiser K, Klein M, Gottschalk C, Hoss F, Scheu S, Coch C, Hartmann G, Kurts C. Cutting edge: The RIG-I ligand 3pRNA potently improves CTL cross-priming and facilitates antiviral vaccination. J Immunol 2016; 196:2439-43; PMID:26819202; https://doi.org/ 10.4049/jimmunol.1501958 [DOI] [PubMed] [Google Scholar]
  • [154].Kulkarni RR, Rasheed MA, Bhaumik SK, Ranjan P, Cao W, Davis C, Marisetti K, Thomas S, Gangappa S, Sambhara S, et al.. Activation of the RIG-I pathway during influenza vaccination enhances the germinal center reaction, promotes T follicular helper cell induction, and provides a dose-sparing effect and protective immunity. J Virol 2014; 88:13990-4001; PMID:25253340; https://doi.org/ 10.1128/JVI.02273-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [155].Beljanski V, Chiang C, Kirchenbaum GA, Olagnier D, Bloom CE, Wong T, Haddad EK, Trautmann L, Ross TM, Hiscott J. Enhanced influenza virus-like particle vaccination with a structurally optimized RIG-I agonist as adjuvant. J Virol 2015; 89:10612-24; PMID:26269188; https://doi.org/ 10.1128/JVI.01526-15 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [156].Zevini A, Olagnier D, Hiscott J. Crosstalk between cytoplasmic RIG-I and STING sensing pathways. Trends Immunol. 2017; 38:194-205; PMID:28073693; https://doi.org/ 10.1016/j.it.2016.12.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [157].Suschak JJ, Wang S, Fitzgerald KA, Lu S. Identification of aim2 as a sensor for DNA vaccines. J Immunol 2015; 194:630-6; PMID:25488991; https://doi.org/ 10.4049/jimmunol.1402530 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [158].Suschak JJ, Wang S, Fitzgerald KA, Lu S. A cGAS-Independent STING/IRF7 Pathway Mediates the Immunogenicity of DNA Vaccines. J Immunol 2016; 196:310-6; PMID:26590319; https://doi.org/ 10.4049/jimmunol.1501836 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [159].Borggren M, Nielsen J, Bragstad K, Karlsson I, Krog JS, Williams JA, Fomsgaard A. Vector optimization and needle-free intradermal application of a broadly protective polyvalent influenza A DNA vaccine for pigs and humans. Hum Vaccin Immunother 2015; 11:1983-90; PMID:25746201; https://doi.org/ 10.1080/21645515.2015.1011987 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [160].Williams JA. Vector design for improved DNA vaccine efficacy, safety and production. Vaccines 2013; 1:225-249; PMID:26344110; https://doi.org/ 10.3390/vaccines1030225 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [161].Sin JI, Kim JJ, Zhang D, Weiner DB. Modulation of cellular responses by plasmid CD40L: CD40L plasmid vectors enhance antigen-specific helper T cell type 1 CD4+ T cell-mediated protective immunity against herpes simplex virus type 2 in vivo. Hum Gene Ther 2001; 12:1091-102; PMID:11399230; https://doi.org/ 10.1089/104303401750214302 [DOI] [PubMed] [Google Scholar]
  • [162].Agadjanyan MG, Chattergoon MA, Holterman MJ, Monzavi-Karbassi B, Kim JJ, Dentchev T, Wilson D, Ayyavoo V, Montaner LJ, Kieber-Emmons T, et al.. Costimulatory molecule immune enhancement in a plasmid vaccine model is regulated in part through the Ig constant-like domain of CD80/86. J Immunol 2003; 171:4311-9; PMID:14530356; https://doi.org/ 10.4049/jimmunol.171.8.4311 [DOI] [PubMed] [Google Scholar]
  • [163].Kim JJ, Yang JS, Lee DJ, Wilson DM, Nottingham LK, Morrison L, Tsai A, Oh J, Dang K, Dentchev T, et al.. Macrophage colony-stimulating factor can modulate immune responses and attract dendritic cells in vivo. Hum Gene Ther 2000; 11:305-21; PMID:10680844; https://doi.org/ 10.1089/10430340050016049 [DOI] [PubMed] [Google Scholar]
  • [164].Kim JJ, Tsai A, Nottingham LK, Morrison L, Cunning DM, Oh J, Lee DJ, Dang K, Dentchev T, Chalian AA, et al.. Intracellular adhesion molecule-1 modulates beta-chemokines and directly costimulates T cells in vivo. J Clin Invest 1999; 103:869-77; PMID:10079108; https://doi.org/ 10.1172/JCI6024 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [165].Kim JJ, Yang JS, Montaner L, Lee DJ, Chalian AA, Weiner DB. Coimmunization with IFN-gamma or IL-2, but not IL-13 or IL-4 cDNA can enhance Th1-type DNA vaccine-induced immune responses in vivo. J Interferon Cytokine Res 2000; 20:311-9; PMID:10762079; https://doi.org/ 10.1089/107999000312450 [DOI] [PubMed] [Google Scholar]
  • [166].Kim JJ, Trivedi NN, Nottingham LK, Morrison L, Tsai A, Hu Y, Mahalingam S, Dang K, Ahn L, Doyle NK, et al.. Modulation of amplitude and direction of in vivo immune responses by co-administration of cytokine gene expression cassettes with DNA immunogens. Eur J Immunol 1998; 28:1089-103; PMID:9541605; https://doi.org/ 10.1002/(SICI)1521-4141(199803)28:03%3c1089::AID-IMMU1089%3e3.0.CO;2-L [DOI] [PubMed] [Google Scholar]
  • [167].Boyer JD, Nath B, Schumann K, Curley E, Manson K, Kim J, Weiner DB. IL-4 increases Simian immunodeficiency virus replication despite enhanced SIV immune responses in infected rhesus macaques. Int J Parasitol 2002; 32:543-50; PMID:11943227; https://doi.org/ 10.1016/S0020-7519(01)00355-1 [DOI] [PubMed] [Google Scholar]
  • [168].Sin JI, Kim J, Pachuk C, Weiner DB. Interleukin 7 can enhance antigen-specific cytotoxic-T-lymphocyte and/or Th2-type immune responses in vivo. Clin Diagn Lab Immunol 2000; 7:751-8; PMID:10973449 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [169].Kim JJ, Nottingham LK, Sin JI, Tsai A, Morrison L, Oh J, Dang K, Hu Y, Kazahaya K, Bennett M, et al.. CD8 positive T cells influence antigen-specific immune responses through the expression of chemokines. J Clin Invest 1998; 102:1112-24; PMID:9739045; https://doi.org/ 10.1172/JCI3986 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [170].Sin J, Kim JJ, Pachuk C, Satishchandran C, Weiner DB. DNA vaccines encoding interleukin-8 and RANTES enhance antigen-specific Th1-type CD4(+) T-cell-mediated protective immunity against herpes simplex virus type 2 in vivo. J Virol 2000; 74:11173-80; PMID:11070014; https://doi.org/ 10.1128/JVI.74.23.11173-11180.2000 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [171].Chattergoon MA, Saulino V, Shames JP, Stein J, Montaner LJ, Weiner DB. Co-immunization with plasmid IL-12 generates a strong T-cell memory response in mice. Vaccine 2004; 22:1744-50; PMID:15068858; https://doi.org/ 10.1016/j.vaccine.2004.01.036 [DOI] [PubMed] [Google Scholar]
  • [172].Kutzler MA, Robinson TM, Chattergoon MA, Choo DK, Choo AY, Choe PY, Ramanathan MP, Parkinson R, Kudchodkar S, Tamura Y, et al.. Coimmunization with an optimized IL-15 plasmid results in enhanced function and longevity of CD8 T cells that are partially independent of CD4 T cell help. J Immunol 2005; 175:112-23; PMID:15972637; https://doi.org/ 10.4049/jimmunol.175.1.112 [DOI] [PubMed] [Google Scholar]
  • [173].Kim JJ, Nottingham LK, Tsai A, Lee DJ, Maguire HC, Oh J, Dentchev T, Manson KH, Wyand MS, Agadjanyan MG, et al.. Antigen-specific humoral and cellular immune responses can be modulated in rhesus macaques through the use of IFN-gamma, IL-12, or IL-18 gene adjuvants. J Med Primatol 1999; 28:214-23; PMID:10593488; https://doi.org/ 10.1111/j.1600-0684.1999.tb00272.x [DOI] [PubMed] [Google Scholar]

Articles from Human Vaccines & Immunotherapeutics are provided here courtesy of Taylor & Francis

RESOURCES