Next Article in Journal
Development of Blueberry-Derived Extracellular Nanovesicles for Immunomodulatory Therapy
Previous Article in Journal
An Overview of Stimuli-Responsive Intelligent Antibacterial Nanomaterials
Previous Article in Special Issue
Virus-like Particle Vaccine Expressing the Respiratory Syncytial Virus Pre-Fusion and G Proteins Confers Protection against RSV Challenge Infection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Non-Invasive Vaccines: Challenges in Formulation and Vaccine Adjuvants

Department of Chemical and Materials Engineering, University of Alberta, Edmonton, AB T6G 1H9, Canada
*
Author to whom correspondence should be addressed.
Pharmaceutics 2023, 15(8), 2114; https://doi.org/10.3390/pharmaceutics15082114
Submission received: 13 July 2023 / Revised: 7 August 2023 / Accepted: 8 August 2023 / Published: 9 August 2023

Abstract

:
Given the limitations of conventional invasive vaccines, such as the requirement for a cold chain system and trained personnel, needle-based injuries, and limited immunogenicity, non-invasive vaccines have gained significant attention. Although numerous approaches for formulating and administrating non-invasive vaccines have emerged, each of them faces its own challenges associated with vaccine bioavailability, toxicity, and other issues. To overcome such limitations, researchers have created novel supplementary materials and delivery systems. The goal of this review article is to provide vaccine formulation researchers with the most up-to-date information on vaccine formulation and the immunological mechanisms available, to identify the technical challenges associated with the commercialization of non-invasive vaccines, and to guide future research and development efforts.

Graphical Abstract

1. Introduction

Vaccine administration methods can be categorized into invasive, minimally invasive, and non-invasive administration, and conventional vaccine administration primarily relies on invasive administration (e.g., intramuscular, subcutaneous, and intradermal) to maximize vaccine bioavailability through site-specific administration. However, such methods of administration have inherent limitations that compromise their utility (Figure 1). Firstly, liquid formulations used in these methods necessitate a cold chain system, making them susceptible to accidental freezing [1,2]. Certain temperature ranges can drastically reduce the shelf life of vaccines, with examples including the Pfizer-BioNTech (<6 months at −60 to −90 °C and <2 h at 8–30 °C), ModernaTX (<30 days at 2–8 °C and <12 h at 8–25 °C), and Oxford-AstraZeneca (<6 months at 2–8 °C and <2 h at 8–25 °C) SARS-CoV-2 vaccines [3]. mRNA vaccines are especially easily affected by changes in liquid formulation (i.e., pH, buffers, temperature, amount of oxygen, etc.), which can lead to the degradation of the vaccines through hydrolysis [4].
Second, the use of needle-based delivery of liquid vaccines also presents technical and safety concerns. The variability in adipose tissue thickness among individuals necessitates adjusting the needle length based on body weight and sex. For instance, a needle length of 1 inch is recommended for individuals weighing less than 70 kg, 1–1.25 inches for men weighing 70–118 kg and women weighing 70–90 kg, and 1.5 inches for men weighing over 118 kg and women weighing over 90 kg [2,5]. Achieving accurate administration requires specialized personnel and proper setups, posing challenges for vaccination in isolated communities and developing countries with limited resources and personnel. Furthermore, needle-based administration carries the risk of needlestick injuries and the transmission of bloodborne pathogens to the injection site (e.g., hepatitis B, hepatitis C, and human immunodeficiency virus (HIV)) [6].
Third, invasive vaccines primarily elicit systemic immune responses in the lower respiratory tract, unlike natural infections that stimulate robust, both systemic and mucosal immunities [7,8]. While systemic immune responses can provide disease-attenuating or disease-preventing immunity through the production of immunoglobulin G (IgG) [7], the limited mucosal immunity in the lower respiratory tract can hinder the elimination of pathogens in the early stages, potentially exacerbating disease severity. On the other hand, minimally invasive and non-invasive vaccines can achieve mucosal vaccination that initiates both robust systemic and mucosal immunities, thereby efficiently preventing pathogens from developing severe disease [9,10].
Invasive vaccines have demonstrated high stability and safety when administered correctly, leading to their predominant use in present-day vaccine delivery. However, due to the aforementioned limitations, there is a growing demand for the development of alternative vaccines, and minimally invasive and non-invasive vaccines have been actively researched as promising candidates. The scope of this manuscript primarily revolves around non-invasive vaccines, which have undergone more comprehensive studies and research compared to minimally invasive vaccines. This review aims to shed light on the advantages and challenges of non-invasive vaccine approaches, along with conventional invasive counterparts, while providing valuable insights into their formulation strategies.

2. Types of Vaccines and Their Formulation

Since Louis Pasteur’s groundbreaking conceptualization of vaccination, significant advancements have been made in the development of vaccines in terms of their stability and effectiveness [11]. To date, seven types of vaccines have been developed: live attenuated, inactivated, replicating viral vector, non-replicating viral vector, DNA, RNA, and subunit vaccines. Each vaccine type leverages distinct immunological mechanisms to trigger adaptive immunity upon administration [12]. Replicable vaccines, including live attenuated vaccines, replicating viral vector vaccines, and replicating nucleic acid vaccines, enter host cells and replicate virulent molecules, which are subsequently taken up by dendritic cells (DCs). In contrast, non-replicable vaccines, such as non-replicating viral vector vaccines, inactivated vaccines, subunit vaccines, and non-replicating nucleic acid vaccines, are directly phagocytosed by DCs [12,13]. Replicable vaccines are characterized by prolonged immunogenicity compared to non-replicable vaccines due to their ability to replicate. Whole-particle-based vaccines, such as live attenuated and inactivated vaccines, exhibit the highest immunogenicity compared to other types of vaccines due to the presence of immunostimulant molecules within the particles, effectively inducing immunity similar to that of natural infections [14,15,16,17,18].
Following the phagocytosis, DCs carrying vaccine particles migrate to the draining lymph node, where they present peptides derived from the vaccine antigens on the major histocompatibility complex (MHC) class II or MHC class I molecules. MHC II activation stimulates CD4+ T cells (T helper cells), including T helper (Th) 1 cells and Th2 cells [19]. Th1 cells play a role in cell-mediated immunity by secreting interferon-gamma (IFN-γ), a proinflammatory cytokine that eliminates external pathogenic particles [20,21]. Th1 cells also activate CD8+ T cells (cytotoxic T cells) by interacting with MHC I [22]. This interaction leads to the production of granzymes, which induce apoptosis, and the pore-forming protein perforin, which physically damages the target cell membrane, allowing granzymes to enter the cell [23]. In contrast, Th2 cells initiate humoral immunity by producing interleukin-4 (IL-4), IL-5, and IL-13, which promote the proliferation and maturation of B cells. Additionally, Th2 cells produce anti-inflammatory IL-10, which modulates the inflammatory responses of Th1 cells (Figure 2) [20,21].

2.1. Live Attenuated Vaccines

Live attenuated vaccines (LAVs) consist of live bacteria or virus particles that have been weakened but retain the ability to initiate immune responses. The production of live attenuated vaccines involves either natural or engineered mutations. In natural production, bacteria or viruses are subjected to multiple passages through different species in a process known as serial passage exposure experiments (SPEs) [24,25]. For example, weakened Theileria annulata loses its virulence against its original host, cattle, after undergoing 50–100 passages of in vitro culturing [26]. This loss of virulence is attributed to alterations or selections in gene expression triggered by environmental changes. In the case of viruses, genetic modifications can occur even more rapidly. For example, nucleopolyhedrosis viruses (NPVs) have been shown to lose virulence against three out of their six original hosts after eight passages in Pseudoplusia includens [27].
Engineered mutation involves the use of transposon mutagenesis, which facilitates the translocation of repetitive DNA sequences to the host cell nucleus, thereby introducing random mutations in the host gene [28]. Another method, gamma rays, is commonly employed to induce DNA strand breaks and oxidative stress in viruses, leading to single base substitutions. However, such mutagenesis techniques often raise safety concerns, including the potential re-emergence of virulence upon repeated exposures to the same host or the restoration of mutated gene sites [24]. Historical examples of virulence re-emergence include the poliovirus vaccine in 1964, the yellow fever vaccine in 2008, and the rotavirus vaccine in 2009 [29]. As a result, research has been focused on developing methods to prevent pathogens from regaining virulence.
One strategy to minimize the re-emergence of virulence is to selectively choose variants with high genetic fidelity for vaccine development [30]. Another approach is multi-region replacement, often combined with gene targeting methods, which involves increasing the number of mutated regions to decrease the probability of mutations reverting to their original state. For example, codon deoptimization replaces conventional codons with synonymous codons, resulting in the same level of immune responses with reduced virulence and re-emergence [30]. Studies have shown that after achieving 97% codon replacement in the contiguous capsid region of polioviruses, the transmission rate in host cells decreased by more than 10-fold [31]. Undesirable mutations can be minimized using small interfering RNA (siRNA) and microRNA (miRNA), which target specific genes for degradation and repress the translation of respective proteins. Zinc finger (ZF) domains, which are protein motifs with finger-like protrusions that bind to DNA, can also be utilized to target specific genetic sites. ZF domains are typically engineered to express multiple types of domains, such as restriction enzymes, to suppress gene transcription or replication origin-competing domains to reduce virus replication [30].

2.2. Inactivated Vaccines

Inactivated vaccines offer a high level of safety compared to live attenuated vaccines, as they are free of concerns regarding virulence reversion and transmission [29]. However, these vaccines tend to have relatively low immunogenicity compared to live attenuated vaccines [29]. Therefore, the design of inactivated vaccines must consider two key factors: complete pathogen inactivation and epitope conservation. Meeting the aforementioned requirements requires a comprehensive study of various virus behaviors, including aggregation, protein crosslinking, protein denaturation, and degradation [32].
Delrue (2012) has described the mechanism and chemical reactions involved in different inactivation treatments, such as formaldehyde, glutaraldehyde, 2,2′-dethiodipyridine, β-propiolactone (BPL), binary ethylene imine, pH, temperature, gamma irradiation, and ultraviolet light [32]. Among these treatments, formaldehyde and β-propiolactone (BPL) have been most commonly used in vaccine development for many years [29]. Formaldehyde primarily acts by deforming the adenine residues of genes and amino acids through monohydroxymethylation (NH-CH2OH), which creates a methylene spacer (-CH2-) that bridges two independent molecules. On the other hand, BPL primarily alkylates the guanine residues of genes, thereby disrupting the reading of sequences [32]. Unlike formaldehyde, BPL treatments preserve the intact epitopes of antigens, leading to greater immunogenicity [29]. Currently, three inactivated vaccines for SARS-CoV-2 have received FDA approval [33]. All three vaccines (Coronavac—Sinovac, Covaxin (BBV152)—Bharat Biotech International, and COVID-19 vaccine (Vero cell)—Sinopharm) employ BPL for the inactivation of the whole SARS-CoV-2 virus [34,35,36]. However, inactivated vaccines face limitations, as they may exhibit poor clearance of unnecessary viral molecules from the human body, which could interact with and harm host cells. Additionally, the process of culturing whole viruses is considered time-consuming compared to gene-based vaccines that can be rapidly amplified [19]. Consequently, there has been a significant shift in the industry towards simpler vaccine forms, such as recombinant DNA and protein subunits [37].

2.3. Replicating and Non-Replicating Viral Vector Vaccines

Replicating and non-replicating viral vector vaccines have been developed to minimize the insertion of pathogenic molecules into the human body. These vaccines utilize different types of vectors to deliver the desired viral genes. Replicating viral vector vaccines employ various vectors such as adenovirus, measles virus, poxviruses, and vesicular stomatitis virus, while non-replicating viral vector vaccines use adenovirus, adeno-associated virus, alphavirus, herpesvirus, and poxviruses [17]. The selection of these vectors is based on their genetic safety, high immunogenicity, and low likelihood of pre-existing immunity in hosts [38].
The main distinction between replicating and non-replicating viral vectors lies in their ability to replicate genes within hosts, and non-replicating viral vectors are engineered to lack gene-replicating sites [39]. As an example, non-replicating adenovirus vectors (such as adenovirus serotype 5, 26, 35) have undergone modifications where the replicating site, Early region 1 (E1), is removed and replaced with an expression cassette that contains a highly active promoter responsible for expressing the inserted foreign gene [40,41]. Similarly, vesicular stomatitis virus (VSV) vectors lack the glycoprotein (G) responsible for attaching to host cells, rendering them unable to infect host cells [38]. Integrase-defective lentiviral vectors (IDLVs) serve as another example of non-replicating viral vectors that lack genes inducing pathogenicity, including replicating genes such as Tat, Rev, Nef, Vif, Vpr, Vpu, Vpx, dUTPase, and open reading frames (ORFs) [42]. However, non-replicating viral vectors require relatively high doses or frequent revaccination due to their inability to replicate [38]. Furthermore, these vectors have the potential to induce mutagenesis in wild-type viruses [43]. In contrast, replicating viral vectors can generate long-lasting immunity even with smaller doses due to their replicability. The replicating viral vectors elicit immune responses similar to natural infections, triggering the production of cytokines and other stimulatory factors that can act as adjuvants [38]. As a result, extensive research has been dedicated to developing replicating viral vector vaccines against diverse viruses (e.g., influenza, HIV, hepatitis B virus) [44]. However, replicating viral vector vaccines raise immune-related safety concerns, particularly in immunocompromised patients. Additionally, the presence of gene-replicating sites in these vectors limits their capacity to accommodate the insertion of small-sized genes (e.g., 7–9 kb for non-replicating adenovirus vectors and 3–4 kb for replicating adenovirus vectors) [17].

2.4. DNA Vaccines

The production of DNA vaccines mostly involves genetic modifications of plasmids derived from Escherichia coli. Subsequently, the modified plasmids are cultivated under fermentation conditions in an anaerobic condition using glycerol, yeast extract, and MgSO4 as carbon, nitrogen, and trace metal sources, respectively. The replicated plasmids can be used directly as genetic vectors after undergoing lysis and purification [45,46]. Conventionally, non-live and non-replicating plasmids have been considered prototypical platforms for DNA vaccines; however, due to their low immunogenicity, replicating DNA vaccines have been considered a potential alternative. In general, replicating DNA vaccines are manufactured by combining eukaryotic promoters (e.g., human cytomegalovirus (CMV) immediate promoters and enhancers) and replicon sequences of alphaviruses [14,47,48]. Furthermore, incorporating additional gene inserts, such as adjuvant coding sequences, has been attempted to enhance the immunogenicity of DNA vaccines [49]. For example, incorporating unmethylated CpG motifs into a DNA vaccine has been shown to function as a type of pathogen-associated molecular pattern (PAMP), which can be utilized to boost vaccine-initiated immune responses [50].
Currently, two types of DNA vaccines have been developed: circular vaccines (i.e., minicircle, minivector, miniknot, etc.) and linear vaccines, including immunological-defined gene expression (MIDGE), Micro-Linear vector (MiLV), and DoggyboneTM DNA, which are covalently closed (Figure 3). In the case of minicircle (mc) DNA, recombinases are used to separate a plasmid into an mc DNA molecule (containing the genes of interest) and a miniplasmid (containing the bacterial plasmid backbone). This is followed by degradation of the miniplasmid using restriction enzymes [45,51]. Studies have shown that mc DNA exhibits a significantly higher transfection rate (68%) compared to the parental plasmid (34%) due to its smaller size (3.8 kbp) in contrast to the parental plasmid (8.2 kbp) [52]. Minivector DNA is a single-stranded and supercoiled structure of less than 2 kbp in size and less than 4 nm in diameter, and is produced through treatments with recombinases (site-specific cleavage) and topoisomerase IV (separation of the daughter DNA) [53,54]. The size difference between the miniplasmid and minivector contributes to the ease of the separation process, contributing to the high purity of the minivector DNA [53]. A miniknot is an mc DNA strand treated with topoisomerase II, which enhances its physical strength to prevent linearization during processing. However, the use of miniknots is still in a hypothetical stage.
The MIDGE (2.9 kbp) and MiLV (1.7 kb) vaccines are produced in a similar manner, but they utilize different parental plasmids. This process entails the cleavage of sites containing the genes of interest using a restriction enzyme, followed by the ligation of the sticky ends of the cleaved sites with hairpin oligodeoxynucleotides to stabilize the structure [43,55,56]. The DoggyboneTM (2.6 kbp) vaccine is also created using plasmids. In this process, plasmids are initially denatured by NaOH, resulting in the production of single-stranded DNA that acts as a template for rolling circle amplification. The amplified single-stranded DNA is then polymerized to form a double-stranded DNA concatemer. This concatemer is subsequently treated with Te1N protelomerase, which recognizes the telomeric ends (Tel-L and Tel-R) and cleaves the linear DNA while covalently closing it. This step removes the bacterial plasmid backbone from the sites containing the genes of interest [57,58]. However, the yield of mc DNA (0.1–8.84 mg/L) is generally much lower compared to conventional plasmids (2.2 g/L of fermentation), and other engineered plasmids are also expected to have lower yields due to physical loss and size reduction during processing [51]. Additionally, although DNA vaccines have exhibited satisfactory immunogenicity in small animals such as rodents, there is a need for improvement when it comes to large animals and humans [59].

2.5. mRNA Vaccine

During the initial stages of nucleic acid-based vaccine development, DNA was considered a preferable option to RNA due to its stability and potential for mass production. However, DNA vaccines have shown limited potency in humans and potential risks of integration of inserted DNA into the host genome [15]. In contrast, mRNA vaccines provide certain advantages over DNA vaccines as there is no concern about integration into the host genome. Additionally, since mRNA vaccines are not required to cross the nuclear membrane to initiate the translation process, they tend to have higher immunogenicity than DNA vaccines [60]. Consequently, numerous trials have been conducted to develop mRNA vaccines, encompassing both non-replicating and replicating mRNA variants.
Conventionally, to achieve mass production of mRNA vaccines, mRNA is initially extracted from pathogen particles and converted into complementary DNA (cDNA) through a process called reverse transcription. The cDNA is subsequently inserted into a plasmid and linearized using restriction enzymes to create a 3′ end sequence to achieve similar structure to the original mRNA, followed by transcription of the cDNA in host cells [61]. Nowadays, in vitro transcription is also widely used in mRNA mass production, with a simpler procedure that consists of the insertion of the target gene into DNA templates followed by RNA amplification [62]. The transcribed mRNA contains specific components. These components include a 5′ cap, which regulates translation along with the poly(A) tail, a 5′ untranslated region that regulates mRNA translation, an open reading frame that represents the gene of interest, a 3′ untranslated region that regulates mRNA translation, and a poly(A) tail that enhances the stability of the mRNA [63,64,65].
Replicating RNA vaccines are primarily developed from alphaviruses. For this, the nsP1-4 genes from alphaviruses are extracted and incorporated into conventional non-replicating RNA. Once inside the host cells, these genes are translated into nsP1-4 proteins, which then assemble into an RNA-dependent RNA polymerase (RdRp) complex. This RdRp complex initiates the replication process for RNA, allowing for the amplification of the target gene [66,67].
In terms of immunogenicity, replicating mRNA vaccines are preferred over non-replicating mRNA vaccines. This preference arises from the fact that replicating mRNA vaccines can replicate and maintain replication sites, thereby potentially enhancing their immunogenicity [66].
mRNA vaccines necessitate the use of delivery vehicles due to their electronegativity and large molecular weight (105–106 Da). The negative charge of mRNA, stemming from its phosphate residues, causes repulsion with the anionic phospholipids in the lipid bilayer membrane of host cells [68,69]. Furthermore, mRNA is three to four times larger than other molecules that can easily enter host cells, such as small interfering RNAs (14 kDa) and antisense oligonucleotides (4–10 kDa) [70].
To address this requirement, lipid nanoparticles and polymeric nanoparticles have been extensively investigated as potential delivery vehicles for mRNA. Specifically, cationic lipids have commonly been used for lipid nanoparticles due to their capacity to readily interact with negatively charged mRNA. However, their use has been linked to liver damage and inflammation resulting from interactions with molecules in the body [68]. To tackle this issue, ionizable lipids have been developed. These lipids carry a positive charge under acidic conditions but become neutral at physiological pH [71]. This ionizable property allows them to associate with mRNA under acidic conditions while reducing toxicity in host cells at neutral pH. To date, five different types of ionizable lipids have been developed: unsaturated ionizable lipids, multi-tail ionizable lipids, ionizable polymer–lipids, biodegradable ionizable lipids, and branch-tail ionizable lipids [71].
Lipid nanoparticles are typically formulated with various types of lipids, including cholesterol, polyethylene glycol (PEG) lipid, and phospholipids, in varying composition ratios. Each component has a distinct purpose. For instance, cholesterol enhances the integrity and rigidity of the particles, while PEG lipid reduces particle aggregation and enhances interaction with ligands on target cells [72]. The properties of phospholipids are determined by lipid saturation. For example, 1,2-distearoyl-sn-glycero-3-phosphocholine (DSPC), which consists of saturated lipids, forms a stable lamellar shape that maintains excellent stability over time. In contrast, unsaturated lipids like 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) and dioleoylphosphatidylcholine (DOPE) undergo a transition from a lamellar shape to an inverted conical shape at room temperature due to their lower phase transition temperature, enhancing the efficiency of cargo release from the vehicle [73].
In addition to lipids, polymeric materials have undergone extensive research for efficient delivery of nucleic acids. Cationic polymers such as poly(lysine), poly(ethylene imine) (PEI), and poly(amidoamine) dendrimers exhibit high transfection efficiency but can be cytotoxic as lipid nanoparticles. Therefore, degradable materials like water-soluble lipopolymer (WSLP) are often combined with cationic polymers to reduce toxicity and prevent accumulation [74,75]. Chitosan, known for its biocompatibility, biodegradability, and low cost, is another promising candidate for mRNA delivery. However, its efficacy is not as pronounced as PEI due to its insolubility at physiological pH [76]. To achieve high transfection efficiency and low toxicity, biodegradable polyesters containing cationic side chains have been synthesized, including poly(L-lactide-co-L-lysine), poly(serine ester), poly(4-hydroxy-L-proline ester), and poly [R-(4-aminobutyl)-L-glycolic acid] [74].

2.6. Subunit Vaccine

Subunit vaccines have been developed to address safety concerns related to whole-particle-based vaccines. These vaccines can be categorized as protein, polysaccharide, conjugate, virus-like particle (VLP), and toxoid vaccines [29,77]. The production of protein-based subunit vaccines involves the harvest and purification of antigens from bacteria using recombinant DNA encoding the target protein. This is often followed by coupling with carriers (i.e., liposome, polymeric nanoparticles, etc.) to enhance the antigen’s stability against environmental changes in the body, such as pH variations and enzymatic activity [77].
Polysaccharide vaccines are derived from bacterial polysaccharide capsules and primarily induce B cell immune responses [78]. However, these vaccines are unable to elicit T cell immune responses as polysaccharides are considered T-independent antigens [79]. To enhance immunogenicity and promote long-term immunity, polysaccharide vaccines are often combined with immunogenic carrier proteins and form conjugate vaccines [80]. Currently, there are five licensed immunogenic carrier proteins: genetically modified cross-reacting material (CRM), tetanus (T) toxoid, diphtheria (D) toxoid, outer membrane protein complex (OMPC), and Haemophilus influenzae protein D (HiD) [81]. CRM is a toxin isolated from Corynebacterium diphtheriae and is often modified through mutations to reduce toxicity [82]. T and D toxoids are detoxified toxins obtained from Clostridium tetani and Corynebacterium diphtheriae, respectively, through formaldehyde treatment [83,84]. OMPC and HiD are non-toxic proteins derived from the outer membrane of Neisseria meningitidis serogroup B and Haemophilus influenzae, respectively [81,85].
VLPs are composed of self-assembling virus capsid proteins derived from various systems, such as bacteria, yeast, insect cells, plant cells, mammalian cells, and cell-free cultures. These extracted proteins can either self-assemble into single-protein VLPs, combine with other proteins to form multi-protein VLPs, or incorporate antigenic materials to enhance immune responses (chimeric VLPs). Since VLPs mimic the structure of pathogens and contain antigens themselves, they can stimulate immune responses, inducing both humoral and cell-mediated immunity, similar to natural infections [86,87].
The production of toxoid vaccines involves harvesting and purifying exotoxins, followed by their inactivation using heat, formaldehyde, or iodoacetamide [88,89]. Subsequently, protein subunits comprising major domains are used instead of full-length proteins to avoid potential immunostimulatory effects caused by activated CD4+ T cells and macrophages [90]. Toxoid vaccines are effective against pathogens that utilize toxins as their primary infection mechanism, as they stimulate the production of anti-toxoid antibodies that can neutralize pathogenic toxins [90].
Different subunit vaccines can be employed to effectively control infections depending on the type of pathogen. For instance, in the case of SARS-CoV and MERS-CoV, subunit vaccines have targeted the spike (S) proteins, which play a crucial role in infection by binding to host receptors [91,92,93]. As another example, the majority of research for influenza virus subunit vaccines has been performed with the HA protein, due to its high immunogenicity and conservability. As a result, several HA protein vaccines, including Fluvirin (1998), Agriflu (2009), and Flucelvax (2012), have been developed and FDA-approved [94]. Apart from HA proteins, another class of membrane proteins known as the extracellular domain of matrix protein 2 (M2e) has been explored for subunit vaccine development against influenza viruses [95]. M2e proteins are highly conserved and abundant in the cell membrane across all influenza A virus subtypes, providing cross-protective immunity when included in vaccines. However, M2e proteins have low immunogenicity and often require additional highly immunogenic carriers or adjuvants, posing challenges to the development of M2e vaccines beyond their current research stage [95]. In contrast, HA proteins are known to be more immunogenic than M2e proteins, making them an even more appealing candidate for subunit vaccines [96].

2.7. Adjuvants

Adjuvants are commonly used alongside vaccines to enhance their performance, including rapid response, long-lasting memory, and reduction in dose [97]. While whole-particle-based vaccines usually do not require adjuvants due to the presence of inherent adjuvant molecules, such as lipopolysaccharide, flagellin, and cytosine-phosphate-guanine, adjuvants are primarily used with subunit vaccines, as well as some DNA and mRNA vaccines [98]. Delivery vehicles with adjuvanticity have garnered attention as a promising type of adjuvant due to their versatility and ability to be combined with various molecules, such as other adjuvants, anchor proteins, and stabilizers. These delivery vehicles not only facilitate the transportation of the vaccine contents to targeted cells with controlled drug release but also enhance immune responses through a number of mechanisms.
Different types of delivery vehicles with adjuvanticity are currently under development, as shown in Table 1. Oil-in-water (O/W) emulsions are extensively studied adjuvants, characterized by a simple composition of tween and/or span, as well as a straightforward production process [99]. However, O/W emulsions have limitations, such as low thermal stability and relatively large particle sizes (<50–500 µm in diameter), which restrict their usage [100,101]. Conversely, inorganic micro/nanoparticles offer high thermal stability and can be flexibly shaped into many different forms. The drug release of inorganic micro/nanoparticles can be regulated by adjusting the amount of reactive moieties involved in degradation processes, such as hydrolysis (silica nanoparticles) and pH sensitivity (calcium phosphate nanoparticles) [102]. In comparison to other types of inorganic micro/nanoparticles, gold nanoparticles are known to be inert, enabling them to form stable functionalizations or have multi-layer coatings to enhance the bioavailability and immune responses of vaccines [103]. Nevertheless, long-term safety concerns persist with these vehicles due to their high stability [104].
In comparison to inorganic nanoparticles, protein-based nanoparticles (PNPs), commonly derived from animal and plant sources, have garnered attention due to their superior biodegradability, biocompatibility, and safety profiles. PNPs have demonstrated several other advantages, such as a convenient and affordable manufacturing process (attributed to the abundant presence of proteins in nature) and the possibility of modification for controlled delivery and specific targeting [150,151,152]. Some of the most commonly explored protein candidates for PNPs include human serum albumin, gelatin, and zein, while different PNPs exhibit different stability and immunogenicity profiles [123,153]. However, PNPs are also known for relative disadvantages, including challenges in controlling size and vaccine dose [153].
Lipid-based micro/nanoparticles are currently used as delivery vehicles for various types of commercial vaccines. These vehicles can encapsulate both hydrophilic and lipophilic particles due to their bilayer lipid structure [154]. They demonstrate high delivery efficiency, primarily because of their positive charge, which promotes their interaction with negatively charged antigen-presenting cells (APCs) [109]. However, it is important to note that this positive charge of lipid-based vehicles can also have side effects, such as destabilization of cell membranes by interacting with negatively charged proteins on the cell membrane [155].
Among the synthetic polymers used for vaccine delivery, poly(lactic-co-glycolic acid) (PLGA) and poly(lactic acid) (PLA) have found widespread application in vaccine delivery systems [156]. PLA exhibits a slower degradation rate compared to poly(glycolic acid) (PGA) due to the presence of a methyl group in its backbone, imparting hydrophobicity and reducing hydrolysis reactions [157]. Additionally, PLGA, the combined form of PLA and PGA, allows controlled drug release by adjusting the proportion of each polymer [158]. However, these synthetic polymers are limited by acidic byproducts, which can cause inflammation. In contrast, biopolymers are naturally occurring materials with relatively low toxicity. For example, acetalated dextran has been extensively used as a biopolymer in vaccine delivery systems due to its biocompatibility and the ease of modulating degradation time by adjusting the proportions of acyclic acetals (lower stability) and cyclic acetals (higher stability) to modulate degradation time [159].

2.8. Future Prospective

Whole-particle-based vaccines offer distinct advantages, such as strong immunogenicity and the ability to induce immune responses similar to natural infections. However, their production process is considered time-consuming and energy-intensive, making them cost-ineffective. Additionally, these vaccines can be immunostimulatory as they involve the injection of external materials. To address these challenges, gene- or protein-particle-based vaccines have emerged as alternative options. However, these vaccines often lack sufficient immunogenicity and transfection efficiency [86]. As a result, they are often encapsulated or coupled with self-adjuvating delivery vehicles. Nonetheless, these vehicles face limitations in terms of toxicity, stability, and low immunogenicity.
To overcome these limitations, new technologies have been developed, including various forms of nanomaterials such as hydrogel nanoparticles, carbon nanotubes, and dendrimers [86,160,161]. Moreover, various nanomaterials have demonstrated high efficiency in drug delivery, showing potential for inducing both humoral and cell-mediated immune responses and overcoming the limitations of conventional delivery vehicles [162,163,164,165]. For instance, hydrogel nanoparticles and carbon nanotubes can bind to bioactive molecules for long periods, resulting in sustained immunity [161]. As another example, cationic dendrimers, which form complexes with nucleic acids, do not destabilize cell membranes like lipid-based vehicles. Furthermore, due to their small sizes (1–10 nm), they can be encapsulated in other types of vehicles for additional functionalization [155].
To ensure the successful development of novel self-adjuvating carriers, several requirements need to be met, including stability, non-toxicity, biodegradability, immunogenicity, and controlled drug release. Commercializing these innovative technologies requires conducting pre-clinical and clinical research to evaluate their safety profiles. Overall, these emerging advancements in vaccine research are expected to bring about a positive transformation in vaccine development, offering safer and more effective treatment options for disease outbreaks.

3. Administration of Non-Invasive Vaccines

Non-invasive vaccines have gained significant recognition due to the drawbacks associated with conventional needle- and liquid-based vaccines, such as the need for cold chain storage, challenges of self-administration, issues with thermal and long-term stability, risks of needle-mediated infections, and needlestick injuries. Among the major non-invasive vaccine systems, oral, intranasal, and transcutaneous administration have been extensively investigated. These non-invasive approaches offer solutions to various limitations of invasive vaccines. However, the effectiveness of non-invasive vaccines is hindered by their limited bioavailability, as biological barriers prevent their efficient absorption into the body. To overcome this issue and enhance the bioavailability of non-invasive vaccines, diverse formulations tailored to specific administration routes have been developed.

3.1. Oral Administration

Oral vaccination has demonstrated its effectiveness in stimulating both mucosal and systemic immune responses by inducing the production of IgA and IgG antibodies [166]. The immune responses are mainly initiated in the intestine, where the majority of vaccine uptake and drug absorption occur [167,168]. The intestine consists of several layers, including the mucus layer, water layer, epithelial layer, basement membrane/Peyer’s patches, and lymph nodes [169]. The epithelial layer is composed of enterocytes, goblet cells, and microfold cells (M cells), which are joined together by tight junctions that prevent the movement of molecules through paracellular and transcellular routes [169,170,171]. Enterocytes, the most abundant cells in the intestinal epithelium (constituting up to 80% of the local cell population), transport antibodies through transcytosis using the neonatal Fc receptor (FcRn) and form antibody–antigen complexes [172,173]. Goblet cells produce mucin, the main component of the mucus layer, which provides physical protection against the uptake of pathogenic particles [174]. M cells play a crucial role in initiating immune responses by phagocytosing large particles and endocytosing small particles through specialized receptors like pattern recognition receptors (PRRs) [170,171]. Pathogen particles phagocytosed by M cells are subsequently taken up by dendritic cells (DC cells) in Peyer’s patches, leading to the activation of cellular and humoral immunity through the differentiation of T cells and B cells (Figure 4) [175].
Despite the advantageous characteristics of oral immunization, such as stability, independence from cold chain requirements, and ease of administration and storage, there are technical challenges in controlling the behavior of orally administered vaccines. Factors such as the varying pH levels of gastrointestinal fluids (e.g., stomach: 0.8–5, duodenum: ~7, jejunum: ≥7, ileum: ≥7, and colon: 7–8), osmolality, surface tension, viscosity, temperature, volume, hydrodynamics, composition, gastric emptying rate and force, intestinal transit time, and flow rate can impact the performance of orally administered vaccines [167,176]. Additionally, the presence of mucosal barriers and biomolecules, such as digestive enzymes, limits the bioavailability and immunogenicity of oral vaccines [77]. This low efficiency necessitates higher vaccine doses, which can eventually lead to immune tolerance in the host. Moreover, the primary approach for oral vaccines involves the use of heat-killed or attenuated pathogens, which makes it difficult to effectively activate the immune system due to poor penetration of the mucus layer. There is also a risk of attenuated pathogens regaining their toxicity, as exemplified by the oral polio vaccine (OPV) outbreak that occurred in several countries in 2000, including Egypt, Haiti, the Philippines, and the Dominican Republic. Extensive research has been conducted to address these limitations and optimize the administration of oral vaccines [177].

3.1.1. pH Sensitivity

Enhancing the bioavailability of oral vaccines requires them to withstand the low-pH environment of the stomach. To improve vaccine stability, researchers have developed anionic coatings that remain stable under acidic conditions. These coatings consist of anionic polymers containing carboxylic acid groups, which remain unionized in the stomach’s low pH but become ionized in the higher pH of the intestine [178]. One example is Eudragit®, anionic polymers composed of poly(methacrylic acid-co-acrylates) with different chemical compositions [179]. The pH sensitivity of these polymers can be adjusted by varying the number of active carboxylic groups; higher levels of carboxylic groups result in increased sensitivity to low pH. For instance, Eudragit L100 (EL100) contains 48.3% active carboxylic groups and dissolves above pH 6, while Eudragit S100 (ES100), with 29.2% active groups, dissolves above pH 7 [179]. pH sensitivity can also be modified by using ethyl acrylate instead of methyl methacrylate in the polymer formulation, as seen with Eudragit L100-55 (EL100-55), which dissolves at lower pH (>pH 5.5) [180]. Cellulose-based materials, such as hydroxypropyl methylcellulose phthalate (HPMC-P), cellulose acetate phthalate (CAP), hydroxypropyl methylcellulose acetate succinate (HPMC-AS), and cellulose acetate trimellitate (CAT), have also been explored [181]. Among these, HPMC-P and CAP are most commonly used for enteric coatings and their pH-responsive behavior is based on the carboxylic acid residues that remain unionized in the stomach but become ionized in the intestine [182]. However, HPMC-P coating processes involving organic solvents raise concerns about toxicity and safety. Moreover, HPMC-P and CAP have limited thermal stability, restricting their application in capsule production [183]. Lastly, alginate, derived from brown algae, such as Laminaria hyperborea, Laminaria digitata, Laminaria japonica, Ascophyllum nodosum, and Macrocystis pyrifera, has been extensively used to produce coating polymers due to its biocompatibility, low toxicity, and affordability. Alginate’s anionic property also enables it to resist gastric fluid, making it a promising candidate for coating oral vaccines [184,185].

3.1.2. Mucoadhesive Interaction

Increasing the bioavailability of oral vaccines requires enhancing mucoadhesion, which prolongs the retention time in the intestine [186,187,188,189]. To improve mucoadhesion, it is important to understand the behavior of different mucoadhesive materials. The mucoadhesive materials typically have hydrophilic functional groups (e.g., hydroxyl and carboxyl groups), a large molecular weight, high surface-to-volume ratio, and low cross-linking density to maximize interactions with intestinal mucus [190]. Among various natural and synthetic mucoadhesive enhancers, chitosan and carbomer have been widely studied [191].
Chitosan, a natural cationic polymer, interacts with negatively charged sialic groups in mucin through electrostatic attraction. Based on the mucoadhesive effects of chitosan, different chitosan-based conjugates (i.e., chitosan–cysteine, chitosan–glutathione, chitosan–thioglycolic acid, etc.) have been developed for vaccine delivery [187,189]. However, chitosan has limitations such as high solubility in the stomach. To address this issue, chitosan is often conjugated with alginate for coating purposes. The carboxyl groups of alginate interact with the amino groups of chitosan, resulting in strong electrostatic attraction. This alginate–chitosan complex shrinks and forms a gel in the low gastric pH, releasing drugs in the neutral pH of the intestine, enabling targeted drug delivery to the colon [192]. A similar study has also been performed with BSA to develop the alginate–chitosan complex as a vaccine delivery carrier, which has demonstrated the efficiency of the complex at a low pH [193].
Carbomer, a carboxyvinyl polymer containing large amounts of carboxylic groups (56–68%), forms hydrogen bonds with sialic acid and sulfate groups present on mucin glycoproteins’ oligosaccharide chains, thereby increasing mucoadhesive effects [194,195]. Carbopol® 971P, 974P, and 934P are mucoadhesive polymers developed based on carbomer, and their mucoadhesive effects are further improved by combining them with cysteine, which interacts with mucus glycoproteins [196].
Nowadays, nanofibers have also gained attention due to their favorable properties such as high surface-to-volume ratio, high porosity, solubility enhancement, controlled cargo release, and the ability to target various sites (e.g., buccal, vaginal, oral, sublingual, transdermal, gastric, intestinal, colonic, and ocular mucosa). However, across different types of polymers, mucoadhesive effects have intrinsic limitations, including the relatively short adherence time (less than 4–5 h) for cells to absorb vaccines and the challenge of vaccine penetration through the mucus layer [197,198].

3.1.3. Intestinal Permeability

To improve the bioavailability of oral vaccines, it is important to enhance their intestinal permeability due to the presence of the mucus layer that acts as a physical barrier [197,199]. This can be achieved through the use of small-sized particles, non-reactive coatings, or permeation enhancers (Figure 5). Nanoparticles have been extensively used for this purpose because of their ability to easily couple with functional groups and penetrate the mucus layer. It has been reported that anionic nanoparticles themselves have shown a relaxation effect on tight junctions, thereby enhancing the permeability, possibly due to interactions between the nanoparticles and integrin proteins on epithelial cells and the activation of myosin light chain kinase (MLCK) [200]. Furthermore, there have been studies demonstrating the size effect of particles on the penetration of the mucus layer, showing that large particles (1190 nm) experience a 30-fold reduction in particle penetration rate than small particles (510 nm) due to steric obstruction and capture by the mucus mesh (mucus mesh spacing ≥500 nm) [201,202].
Non-reactive nanoparticles can be created by functionalizing them with poly(ethylene glycol) (PEG) and Pluronic F127 to enhance their permeability. In particular, the development of low-molecular-weight PEG coating was inspired by the penetration of viruses through human mucus, reducing hydrophobic and electrostatic interactions between nanoparticles and the mucus layer through a hydrophilic and neutral-charged coating [203]. PEG has been successfully coupled with various types of nanoparticles (e.g., PLGA, poly sebacic acid (PSA), polyethylenimine (PEI), poly-l-lysine (PLL)) and has demonstrated efficient penetration [204]. Pluronic F127, a triblock copolymer surfactant with alternating hydrophilic and hydrophobic structures, exhibits a unique arrangement that minimizes its interaction with the mucus layer. Furthermore, its neutral charge allows it to avoid interactions with negatively charged mucins [205]. It has been reported that Pluronic F127-coated particles show an 80-fold increase in penetration rate compared to uncoated particles [186].
Chemical or enzymatic treatments, such as disulfide-breaking agents or mucolytic enzymes, have been also performed to enhance the intestinal permeability. Such permeation enhancers are typically incorporated inside or immobilized on nanoparticles to selectively affect certain regions of the mucus layer to maintain the functions of mucus layer against pathogens in non-treated areas [186]. N-acetyl-l-cysteine (NAC) is an example of a disulfide-breaking agent that breaks the disulfide bonds of cysteine groups in mucin using free sulfur groups [206,207]. Cleaving these bonds reduces the elasticity and viscosity of mucus, thereby increasing drug permeability through the epithelial layer [204]. Additionally, it is speculated that NAC may enhance the uptake of antigens by loosening the mucus layer at Peyer’s patches, where antigen uptake primarily occurs [174]. Other mucolytic enzymes, such as trypsin, papain, bromelain, and recombinant human DNase (rhDNase), can also contribute to reducing the viscosity of mucus by cleaving the cross-links in mucus glycoproteins [204,206].

3.2. Intranasal Administration

Intranasal vaccination offers several advantages, including easy accessibility, the potential for herd immunization, and the ability to initiate both mucosal and systemic immune responses, making it a favorable administration route for vaccines [208]. Similar to the Peyer’s patches in the intestinal tract, the nasal-associated lymphoid tissues (NALTs) play a significant role in inducing immunity through intranasal vaccination in rodents. NALTs are regions rich in immune cells (such as T cells, B cells, dendritic cells, macrophages, and M cells) where active immune responses occur [209]. The key difference between the Peyer’s patches and NALTs is that the formation of NALT is triggered after exposure to pathogens, whereas Peyer’s patches develop before birth. Similar to the NALTs of rodents, Waldeyer’s rings play a major role in human immune responses upon intranasal administration. Waldeyer’s rings, including the pharyngeal tonsil, tubal tonsils, palatine tonsils, and lingual tonsils, firstly transfer pathogens to DC cells, initiating mucosal immunity through the differentiation and proliferation of T cells [209,210]. Like oral vaccination, intranasal vaccination faces challenges from the mucus layer, which can result in mucociliary clearance of vaccine particles and hinder the vaccination process. Therefore, numerous methods have been investigated to enhance the immune response during intranasal vaccination, such as improving mucoadhesive interaction and intestinal permeability, as well as incorporating adjuvants.

3.2.1. Liquid Formulation

Currently, intranasal vaccines are predominantly administered in liquid form due to their effective cell penetration. Liquid formulations allow for even distribution of vaccine components, such as adjuvants and stabilizers [211]. Developing a liquid formulation involves considering the properties of the liquid phase (e.g., osmolarity, pH, viscosity, and surface tension), the delivery devices, and the inclusion of stabilizers. In terms of osmolarity, isotonic solutions (~290 osmol/L) are commonly used to achieve optimal osmolarity, as hypotonic solutions are poorly absorbed and hypertonic solutions can cause cell shrinkage upon diffusion [211]. The pH of liquid formulations should be maintained between 4.5 and 6.5 to avoid irritation, considering that the nasal mucus layer has a pH of 5.5–6.5; it has also been observed that lysozymes in the intranasal system are inactivated at neutral pH, which can increase the risk of microbial infection [212,213]. Increasing the viscosity of nasal mucus is another approach to enhance the immunogenicity of vaccines by prolonging their retention time in the nasal cavity and increasing the dosing volume of the nasal spray over conventional dosing volume (50–140 µL) [214,215,216]. However, it must be noted that high mucosal viscosity may impede effective diffusion and spreading of the vaccines. Surface tension is another critical factor affecting the behavior of liquid vaccines. Most intranasal vaccines have lower surface tension (30.3–44.9 dynes/cm) than the nasal mucus layer (<56 dynes/cm). [217] The low surface tension of vaccines primarily facilitates better bioavailability by maximizing their spread within the nasal cavity [218].
To enhance the delivery efficiency of intranasal formulations, various devices have been developed, ranging from manual dropping-based devices to multifunctional devices with filters and automatic systems that prevent particle deposition in the lungs and ensure reproducibility [219]. Examples of devices used for liquid formulations include AccuSpray (FluMist), Classic Mexican Nebulizer (Measles), Aerovax and OptiMist (influenza), ViaNase (Alzheimer’s), and MAD and CPSI Cartridge Pump System (adenovirus-based vaccine) [209]. Among the different intranasal vaccination methods and devices, AccuSpray (FluMist) is the only one approved by the FDA for human use in the United States [220].
Due to the low stability of liquid formulations, stabilizers are often incorporated. For instance, FluMist contains monosodium glutamate, hydrolyzed porcine gelatin, and sucrose [221]. Other commonly used stabilizers in intranasal vaccines include lactose, sorbitol, porcine gelatin, arginine, and tricine [222]. These stabilizers can also be used during lyophilization to reduce the glass transition temperature and decrease the lyophilization time [223]. Moreover, sugars such as lactose and sucrose do not form sharp ice crystals, unlike water, providing physical protection to vaccines at low temperatures [224].

3.2.2. Powder Formulation

Powder formulations can resolve limitations of liquid formulations, such as short residence time in the nasal cavity, low stability in terms of chemical and microbiological factors, short shelf-life, and the requirement for cold chain systems [209,225]. Dry powder vaccines are primarily manufactured using two methods: spray-drying and freeze-drying. Spray-drying starts with a liquid vaccine being sprayed into a drying chamber, followed by powder separation in a gas stream. On the other hand, freeze-drying involves freezing a liquid vaccine using liquid nitrogen and then subjecting it to lyophilization [226]. In terms of formulation process, dry powder vaccines share many similarities with conventional injectable vaccines, as well as antigens and adjuvants. However, one difference is that dry powder vaccines often incorporate bulking agents, stabilizers, and mucoadhesive materials to enhance nasal deposition [227]. Bulking agents play an important role as a vaccine carrier, facilitating the attainment of appropriate particle sizes to improve vaccine efficiency. For instance, lactose and mannitol, both FDA-approved inactive ingredients for nasal or respiratory use, are frequently chosen as carriers [228]. For a stabilizer of vaccines, mild formaldehyde has also been extensively used as it interacts with the residues of various antigens, such as the N-terminus of amino acids and side chains (i.e., arginine, cysteine, histidine, etc.) [227,229].
Dry powder vaccines are typically administered using medical devices like the μcoTM system and Opt-powder (influenza), Aktiv-Dry PuffHaler® and BD solvent® (measles), and Combitips Plus Syringe (meningococcal) [209]. While there are several advantageous aspects of powder formulations over liquid counterparts, limitations still exist due to limited research on powder formulation delivery systems and the behavior of dry powder vaccines upon contact with the nasal cavity [211,225]. This is evident considering that there are no FDA-approved dry powder vaccines currently available.

3.3. Transcutaneous Administration

Transcutaneous immunization (TCI) is a topical vaccination strategy that primarily stimulates systemic immunity, also inducing a moderate level of mucosal immunity with the assistance of an adjuvant [230]. TCI specifically targets epidermal cells, located in the middle layer of the skin, which also includes keratinocytes (KCs) and Langerhans cells (LCs). KCs are the predominant cells in the epidermis and play a role in recognizing pathogens through Toll-like receptors (TLRs), triggering innate and adaptive immune responses. LCs, a type of dendritic cells, capture a significant portion of TCI particles and recognize them through the major histocompatibility complex (MHC), activating adaptive immune responses [231]. However, for TCI to be effective, it is necessary to penetrate the stratum corneum, the outer layer of the skin consisting of 10–20 layers of dead skin cells that act as a barrier against external substances [232]. Consequently, research has focused on exploring physical and chemical methods to enhance the penetration of the stratum corneum in order to achieve successful TCI.

3.3.1. Physical Delivery Systems

To enhance the penetration of vaccines through the stratum corneum, various methods have been employed, including ultrasound, electroporation, jet injectors, laser-based technologies, and microneedles [233]. Ultrasound can be utilized to create cavities in the stratum corneum, increasing vaccine permeability. This process involves the pressure gradient generated by ultrasound, enlarging air pockets in the fibrous tissue to allow the diffusion of vaccines through the cavities [234,235]. Electroporation, on the other hand, involves applying short high-voltage pulses to the cell membrane, temporarily damaging it and creating transient hydrophilic pores that enhance membrane permeability [236]. However, both ultrasound and electroporation require large machines and trained personnel to operate them. As an alternative, jet injectors, which are compact and easy to handle, have been used. These injectors use high-pressure liquid to deliver vaccines through the stratum corneum [237,238]. Despite the increasing accessibility, jet injectors have drawbacks including pain, bleeding, and edema [232].
Currently, diverse laser-based technologies (e.g., laser ablation, fractional laser technology, and ablative fractional laser technology) have also been developed to enhance stratum corneum penetration for vaccine delivery. Laser ablation is a conventional technology used to remove the entire surface of the skin, whereas fractional laser creates micro-columns in the skin, leaving the surrounding areas intact to minimize skin damage and accelerate the healing process [239]. Ablative fractional laser (AFL) is a method that combines the laser ablation and fractional laser technologies, minimizing skin damage and enhancing vaccine permeability by ablating the stratum corneum and forming skin microchannels [240].
Among the available transcutaneous immunization (TCI) methods, microneedles have gained acceptance due to their ease of use, accessibility, and independence from patches. Microneedles are minimally invasive and pain-free, involving small needles that penetrate the stratum corneum [241]. There are five categories of microneedles based on their structures: solid, coated, hollow, dissolving, and hydrogel-forming microneedles [242,243,244]. Solid microneedles are used to create microchannels in the skin, allowing for vaccine penetration [245]. To date, a variety of materials have been considered candidates for the production of solid microneedles (e.g., silicon, polysilicon, silicon dioxide, silicon nitride, PGA) [246]. While solid microneedles offer easy manufacturing and high dose delivery, they are hindered by a slower rate of diffusion when compared to other types of microneedles [247,248].
Coated microneedles contain vaccines applied to the needle, enabling diffusion into the epidermis upon injection. However, one limitation is that the coating may cause coated particle loss during manufacturing and handling [241,249]. As another option, hollow microneedles resemble conventional needle injection-based vaccination, with vaccines contained in the bore of the needle. These microneedles also provide the highest level of particle delivery but can experience blockage by tissues upon contact, although repositioning the bore to the side can address this issue [241,244].
Dissolving microneedles are made of water-soluble materials that penetrate and dissolve in the skin simultaneously. For these microneedles, vaccines and the microneedle matrix undergo a mixing process upon manufacturing. Among various candidates for the matrix of dissolving needles (e.g., sodium carboxymethylcellulose, poly(vinylalcohol), poly(vinylpyrrolidone), methylvinylether-co-maleic anhydride), sodium hyaluronate is most commonly used due to its biodegradability and mechanical strength [250,251]. Remarkably, dissolving microneedles have shown high diffusion efficiency within a patch wearing time of just several minutes, while it could be up to several hours for solid microneedles [247].
Lastly, hydrogel-forming microneedles have emerged as a promising option for their biocompatibility and controllable degradability [252]. These microneedles are equipped with hydrogel, capitalizing on its three-dimensional structure that allows for a significant capacity to carry vaccines; however, there is a possibility that vaccines can become trapped within the patch and they require a longer diffusion time of up to several hours [247,253,254]. The mechanism of hydrogel-forming microneedles relies on the absorption of interstitial skin fluid, which causes the hydrogel to swell and burst, triggering the release of the vaccines from the microneedle [254,255]. The vaccine-release behavior of the hydrogel is directly influenced by the swelling property of the matrix polymer and can be modulated by adjusting the amount of crosslinkers; increasing the percentage of crosslinkers results in a reduced swelling rate in the skin [252,254]. Last but not least, achieving optimal performance with hydrogel-forming microneedles requires maintaining their hardness in a dry state, while ensuring rapid swelling in the skin [255].
The efficiency of microneedle-based vaccines can be enhanced by using different types of nanoparticles with various strengths. For instance, polymeric nanoparticles offer controlled release and targeted delivery, while lipid-based nanoparticles provide enhanced solubility. Furthermore, inorganic nanoparticles are known to allow controlled release and have consistent pore structures, whereas protein-based nanoparticles enable controlled release and targeted delivery [256,257]. Currently, lipid-based carriers and polymeric carriers (i.e., poly(vinyl pyrrolidone) (PVP), Poly(vinyl alcohol) (PVA), PLGA, polypropylene sulfide, chitosan, etc.) have been the primary focus of research for vaccine delivery. These carriers have demonstrated enhanced immune responses without requiring the addition of adjuvants [242].

3.3.2. Chemical Enhancers

Various chemical substances, including water, dimethylsulphoxide (DMSO), acids, essential oils, and surfactants, have been investigated as potential enhancers for transcutaneous delivery [258]. These enhancers act to improve the permeability of vaccines through the skin by targeting different components of the stratum corneum: hydrophilic space (intercellular transport), hydrophilic lipid heads (transcellular transport), and hydrophobic lipid chains (transcellular transport) [259]. The hydrophilic space can accommodate solvents (e.g., propylene glycol, ethanol, pyrrolidines, dimethyl sulfoxide) through chemical treatment, allowing the delivery of vaccines and drugs in the stratum corneum, followed by their transport to the epidermis by a chemical gradient [259,260].
Water can interact with the hydrophilic heads of the bilipid membrane, increasing the fluidity and permeability of the stratum corneum, and this favorable property has led to the development of hydro patches used for vaccine penetration through the skin [232,261]. In a similar manner, DMSO ((CH3)2SO) can also effectively enhance the permeability of the stratum corneum; the oxygen atom present in the sulfoxide group of DMSO interacts with the hydrophilic heads of the bilipid membrane, causing a disruption of the hydrogen bonds between lipid heads. This process leads to a transformation of the gel-like structure in the ceramide bilayer of the stratum corneum into a liquid–crystalline structure [262]. On the other hand, lipophilic agents represented by azone, terpenes, and oleic acids disrupt the structure of lipid tails in the stratum corneum, expanding the membrane and facilitating vaccine diffusion [263]. However, these hydrophilic and lipophilic chemical enhancers have limitations in terms of efficiency and safety. Firstly, as the particle delivery into the epidermis is primarily mediated by chemical potential, the effectiveness of these agents diminishes as the concentration gradient decreases. This necessitates the application of mechanical force to enhance the efficiency of delivery. Furthermore, chemical molecules have the potential to affect tissue integrity and cause various forms of irritation, including local inflammation, erythema, swelling, dermatitis, and other related conditions [258,264].

3.4. Adjuvants

Currently, the majority of FDA-approved non-invasive vaccines utilize whole pathogen particles, such as Adenovirus Type 4 and Type 7 vaccines, VAXCHORA, ROTARIX, RotaTeq, Vivotif, and FluMist [220,265,266,267,268,269]. These vaccines typically do not require additional adjuvants, as the parts of the vaccines (e.g., lipopolysaccharide, flagellin, and cytosine-phosphate-guanine) have inherent adjuvanticity [98]. Regardless, there have been attempts to incorporate adjuvants into delivery systems to overcome the low bioavailability of non-invasive vaccinations (Table 2). Among the available adjuvants, aluminum-based adjuvants, including aluminum hydroxide and aluminum phosphate, have been the most widely used for over 80 years with invasive vaccines [270]. These adjuvants enhance immune responses by promoting antigen uptake by dendritic cells [271]. However, aluminum-based adjuvants have demonstrated lower immunogenicity for non-invasive vaccines compared to other adjuvants, likely due to their large particle size, which interferes with diffusion [272,273,274]. Consequently, efforts have been made to identify suitable adjuvants for non-invasive vaccines.
One alternative that has been studied is toxin-based adjuvants, such as cholera toxin (CT) and heat-labile enterotoxin (LT). However, their high toxicity raises concerns regarding their use. To mitigate this, mutations at specific amino acid sites of CT and LT have been introduced, resulting in improved efficiency and safety (e.g., double-mutant LT (dmLT) and multiple-mutated CT (mmCT)) [296]. These adjuvants are reported to specifically enhance the interleukin-17 (IL-17) response, thereby activating antigen-specific T helper cell responses [285,296].
In addition to toxin-based adjuvants, adjuvants based on other biological molecules (e.g., such as polysaccharides, lipids, proteins, and genes) have also been investigated. Polysaccharide-based adjuvants are primarily used to enhance the activity of immune cells, such as macrophages, and induce cytokine-mediated immune responses (e.g., IL-1 and -2 (mannatide), IL-5 and -6 (Advax™), IL-4 and -13 (chitin)) [297,298,299,300,301,302]. Nucleotide-based adjuvants (e.g., synthetic double-stranded RNA polyriboinosinic acid-polyribocytidylic acid [poly (I:C)], cytosine-phosphate-guanine oligodeoxynucleotide (CpG ODN), and cyclic guanosine monophosphate-adenosine monophosphate (2′3′-cGAMP or cGAMP)) are, on the other hand, target-specific molecules. For instance, poly (1:C), CpG ODN, and cGAMP specifically interact with Toll-like receptor (TLR) 3, TLR9, and stimulator of interferon genes (STING) receptors, respectively, activating inflammatory cytokines and type 1 interferon, eventually stimulating both innate and adaptive immune responses [301,302,303,304,305,306].
Lipid-based adjuvants (e.g., liposomes, lipid nanoparticles, emulsions, and immune-stimulating complexes (ISCOMs)) are often employed as vaccine carriers with adjuvanticity, capable of incorporating other additives, as well as facilitating interactions between vaccines and dendritic cells (DCs) [307]. Protein-based adjuvants (e.g., flagellin, poly(γ-glutamic acid), and proteosome), meanwhile, function as ligands for receptors (TLR5, TLR4, and TLR2, respectively) [308,309,310]. Their interactions activate helper T cells and cytotoxic T cells, thus stimulating adaptive immunity [311].
Despite the advantages offered by adjuvants, only a few have received FDA approval, including aluminum salts, adjuvant System 04 (AS04), oil-in-water emulsion, CpG 1018, and Quillaja saponaria 21 (QS-21) [312]. This delayed progress in developing vaccine adjuvants is primarily attributed to their reactogenicity and potential interactions with vaccines [313]. Consequently, in order to advance the development of vaccine adjuvants, extensive research into their interaction mechanisms with vaccines and the human body is essential, along with thorough preclinical and clinical studies aimed at establishing the safety profiles of emerging adjuvants.

3.5. Future Perspectives

Researchers have highlighted the numerous advantages of non-invasive vaccines compared to their invasive counterparts. One significant advantage is the potential to increase vaccination rates by eliminating needle-related fears and phobias that affect a substantial portion of children, adolescents (20–50%), and young adults (20–30%) [314]. Additionally, non-invasive vaccines are cost-effective by eliminating the need for cold chain systems, specialized facilities, and experienced personnel. Furthermore, disease-specific immunization can be achieved by developing and commercializing non-invasive vaccines. As shown in Table 2, Table 3 and Table 4, each administration method targets different diseases: intestine diseases (oral vaccines), respiratory diseases (intranasal vaccines), and outmost skin diseases (transcutaneous). Such an approach concentrates and maximizes immune responses at targeted sites. Additionally, in this sense, combination treatments of different non-invasive vaccines hold the potential for multi-site targeting. However, there is still a scarcity of approved non-invasive vaccines, as well as pre-clinical and clinical trials focused on non-invasive vaccines [24,29,77,315]. Therefore, further research endeavors should be directed towards addressing the challenges of non-invasive vaccines, optimizing their formulations and administration methods, in order to expand their usage and pave the way for effective global immunization against future disease outbreaks.

4. Conclusions

Despite the evident benefits of non-invasive vaccines, their commercialization has encountered physiological barriers that impede the efficient delivery of vaccines to immunological cells, such as antigen-presenting cells (APCs). Each non-invasive vaccine faces specific challenges that restrict their bioavailability. For example, oral administration is hindered by gastrointestinal fluids and mucosal layers, intranasal administration is hindered by mucosal layers, and transcutaneous administration is hindered by the stratum corneum. Therefore, to enhance the immunogenicity of non-invasive vaccines, the formulation of the vaccine should carefully consider limitations associated with each route of administration. Additionally, various supplementary materials can be employed in vaccine formulation to enhance bioavailability, such as pH-sensitive polymers, mucus-penetrating polymers, and microneedles for addressing gastrointestinal fluids, mucosal layers, and the stratum corneum, respectively.
To optimize the immunogenicity of vaccines and minimize their immune-toxicity, careful selection of vaccine types and adjuvants is also essential, as different combinations of these components can lead to varying immune responses and clinical outcomes. For instance, for non-invasive administration with low immunogenicity, adjuvants are often combined with whole-particle-based vaccines to boost their effectiveness. On the other hand, in the case of invasive administration, whole-particle-based vaccines are typically not paired with adjuvants due to their inherent high immunogenicity, whereas adjuvants are mostly used with subunit vaccines due to their lower immunogenicity.
Adjuvants serve not only as immune enhancers but also as carriers for vaccines. These carriers can be further functionalized with other types of adjuvants to bolster immune responses or coupled with anchor proteins and stabilizers for improved performance. Nevertheless, careful consideration must be given to the selection of adjuvants to be paired with vaccines, and their toxicity should be thoroughly assessed through pre-clinical and clinical tests prior to clinical application. Only with such precautions in practice will the development of effective, safe non-invasive vaccines be achieved.

Author Contributions

S.H., P.L. and H.-J.C. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by the Natural Sciences and Engineering Research Council of Canada (NSERC) Discovery Grant [RGPIN-2018-04314]. The funders had no role in the study design, the data collection and analysis, the decision to publish, or the preparation of the manuscript.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lloyd, J.; Lydon, P.; Ouhichi, R.; Zaffran, M. Reducing the loss of vaccines from accidental freezing in the cold chain: The experience of continuous temperature monitoring in Tunisia. Vaccine 2015, 33, 902–907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Criscuolo, E.; Caputo, V.; Diotti, R.A.; Sautto, G.A.; Kirchenbaum, G.A.; Clementi, N. Alternative methods of vaccine delivery: An overview of edible and intradermal vaccines. J. Immunol. Res. 2019, 2019, 8303648. [Google Scholar] [CrossRef] [Green Version]
  3. World Health Organization. COVID-19 Vaccination: Supply and Logistics Guidance. Available online: https://www.who.int/publications/i/item/who-2019-ncov-vaccine-deployment-logistics-2021-1 (accessed on 13 January 2023).
  4. Crommelin, D.J.A.; Anchordoquy, T.J.; Volkin, D.B.; Jiskoot, W.; Mastrobattista, E. Addressing the cold reality of mRNA vaccine stability. J. Pharm. Sci. 2021, 110, 997–1001. [Google Scholar] [CrossRef]
  5. Centers for Disease Control and Prevention. Vaccine Administration. Available online: https://www.cdc.gov/vaccines/hcp/acip-recs/general-recs/administration.html (accessed on 5 February 2023).
  6. Deisenhammer, S.; Radon, K.; Nowak, D.; Reichert, J. Needlestick injuries during medical training. J. Hosp. Infect. 2006, 63, 263–267. [Google Scholar] [CrossRef]
  7. Krammer, F. SARS-CoV-2 vaccines in development. Nature 2020, 586, 516–527. [Google Scholar] [CrossRef] [PubMed]
  8. Park, J.H.; Lee, H.K. Delivery routes for COVID-19 vaccines. Vaccines 2021, 9, 524. [Google Scholar] [CrossRef] [PubMed]
  9. An, X.; Martinez-Paniagua, M.; Rezvan, A.; Sefat, S.R.; Fathi, M.; Singh, S.; Biswas, S.; Pourpak, M.; Yee, C.; Liu, X. Single-dose intranasal vaccination elicits systemic and mucosal immunity against SARS-CoV-2. iScience 2021, 24, 103037. [Google Scholar] [CrossRef]
  10. Ma, Y.; Tao, W.; Krebs, S.J.; Sutton, W.F.; Haigwood, N.L.; Gill, H.S. Vaccine delivery to the oral cavity using coated microneedles induces systemic and mucosal immunity. Pharm. Res. 2014, 31, 2393–2403. [Google Scholar] [CrossRef] [Green Version]
  11. Smith, K.A. Louis Pasteur, the father of immunology? Front. Immunol. 2012, 3, 68. [Google Scholar] [CrossRef] [Green Version]
  12. Trovato, M.; Sartorius, R.; D’Apice, L.; Manco, R.; De Berardinis, P. Viral emerging diseases: Challenges in developing vaccination strategies. Front. Immunol. 2020, 11, 2130. [Google Scholar] [CrossRef]
  13. Bellamkonda, N.; Lambe, U.P.; Sawant, S.; Nandi, S.S.; Chakraborty, C.; Shukla, D. Immune Response to SARS-CoV-2 Vaccines. Biomedicines 2022, 10, 1464. [Google Scholar] [CrossRef]
  14. Kutzler, M.A.; Weiner, D.B. DNA vaccines: Ready for prime time? Nat. Rev. Genet. 2008, 9, 776–788. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, C.; Maruggi, G.; Shan, H.; Li, J. Advances in mRNA vaccines for infectious diseases. Front. Immunol. 2019, 10, 594. [Google Scholar] [CrossRef] [Green Version]
  16. V’kovski, P.; Kratzel, A.; Steiner, S.; Stalder, H.; Thiel, V. Coronavirus biology and replication: Implications for SARS-CoV-2. Nat. Rev. Microbiol. 2021, 19, 155–170. [Google Scholar] [CrossRef]
  17. Robert-Guroff, M. Replicating and non-replicating viral vectors for vaccine development. Curr. Opin. Biotechnol. 2007, 18, 546–556. [Google Scholar] [CrossRef]
  18. Liu, M.A. DNA vaccines: A review. J. Intern. Med. 2003, 253, 402–410. [Google Scholar] [CrossRef] [Green Version]
  19. Pollard, A.J.; Bijker, E.M. A guide to vaccinology: From basic principles to new developments. Nat. Rev. Immunol. 2021, 21, 83–100. [Google Scholar] [CrossRef] [PubMed]
  20. Berger, A. Th1 and Th2 responses: What are they? BMJ 2000, 321, 424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Yanagawa, Y.; Iwabuchi, K.; Onoé, K. Co-operative action of interleukin-10 and interferon-γ to regulate dendritic cell functions. Immunology 2009, 127, 345–353. [Google Scholar] [CrossRef] [PubMed]
  22. Nakamura, T.; Shirouzu, T.; Nakata, K.; Yoshimura, N.; Ushigome, H. The role of major histocompatibility complex in organ transplantation-donor specific anti-major histocompatibility complex antibodies analysis goes to the next stage. Int. J. Mol. Sci. 2019, 20, 4544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Ritz, S.A. The immune system in health and disease. In Clinical Immunology; Elsevier: Amsterdam, The Netherlands, 2004; Volume 6, pp. 393–416. [Google Scholar]
  24. Ebert, D. Experimental evolution of parasites. Science 1998, 282, 1432–1436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Hajra, D.; Datey, A.; Chakravortty, D. Attenuation Methods for Live Vaccines. Methods Mol. Biol. 2021, 2183, 331–356. [Google Scholar]
  26. Sutherland, I.A.; Shiels, B.R.; Jackson, L.; Brown, D.J.; Brown, C.G.D.; Preston, P.M. Theileria annulata: Altered Gene Expression and Clonal Selection during Continuousin VitroCulture. Exp. Parasitol. 1996, 83, 125–133. [Google Scholar] [CrossRef] [PubMed]
  27. Pavan, O.H.; Boucias, D.G.; Pendland, J.C. The effects of serial passage of a nucleopolyhedrosis virus through an alternate host system. Entomophaga 1981, 26, 99–108. [Google Scholar] [CrossRef]
  28. Muñoz-López, M.; García-Pérez, J.L. DNA transposons: Nature and applications in genomics. Curr. Genom. 2010, 11, 115–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Sanders, B.; Koldijk, M.; Schuitemaker, H. Inactivated viral vaccines. In Vaccine Analysis: Strategies, Principles, and Control; Springer: Berlin/Heidelberg, Germany, 2015. [Google Scholar]
  30. Lauring, A.S.; Jones, J.O.; Andino, R. Rationalizing the development of live attenuated virus vaccines. Nat. Biotechnol. 2010, 28, 573–579. [Google Scholar] [CrossRef] [PubMed]
  31. Burns, C.C.; Shaw, J.; Campagnoli, R.; Jorba, J.; Vincent, A.; Quay, J.; Kew, O. Modulation of poliovirus replicative fitness in HeLa cells by deoptimization of synonymous codon usage in the capsid region. J. Virol. 2006, 80, 3259–3272. [Google Scholar] [CrossRef] [Green Version]
  32. Delrue, I.; Verzele, D.; Madder, A.; Nauwynck, H.J. Inactivated virus vaccines from chemistry to prophylaxis: Merits, risks and challenges. Expert Rev. Vaccines 2012, 11, 695–719. [Google Scholar] [CrossRef] [Green Version]
  33. Food and Drug Administration. Emergency Use Authorization. Available online: https://www.fda.gov/emergency-preparedness-and-response/mcm-legal-regulatory-and-policy-framework/emergency-use-authorization (accessed on 25 April 2023).
  34. Sharma, R.; Tiwari, S.; Dixit, A. Covaxin: An overview of its immunogenicity and safety trials in India. Bioinformation 2021, 17, 840. [Google Scholar]
  35. Tanriover, M.D.; Doğanay, H.L.; Akova, M.; Güner, H.R.; Azap, A.; Akhan, S.; Köse, Ş.; Erdinç, F.Ş.; Akalın, E.H.; Tabak, Ö.F. Efficacy and safety of an inactivated whole-virion SARS-CoV-2 vaccine (CoronaVac): Interim results of a double-blind, randomised, placebo-controlled, phase 3 trial in Turkey. Lancet 2021, 398, 213–222. [Google Scholar] [CrossRef] [PubMed]
  36. Tregoning, J.S.; Brown, E.S.; Cheeseman, H.M.; Flight, K.E.; Higham, S.L.; Lemm, N.M.; Pierce, B.F.; Stirling, D.C.; Wang, Z.; Pollock, K.M. Vaccines for COVID-19. Clin. Exp. Immunol. 2020, 202, 162–192. [Google Scholar] [CrossRef]
  37. Sánchez-Ramón, S.; Conejero, L.; Netea, M.G.; Sancho, D.; Palomares, Ó.; Subiza, J.L. Trained immunity-based vaccines: A new paradigm for the development of broad-spectrum anti-infectious formulations. Front. Immunol. 2018, 9, 2936. [Google Scholar] [CrossRef] [PubMed]
  38. Travieso, T.; Li, J.; Mahesh, S.; Mello, J.D.F.R.E.; Blasi, M. The use of viral vectors in vaccine development. NPJ Vaccines 2022, 7, 75. [Google Scholar] [CrossRef] [PubMed]
  39. Chavda, V.P.; Bezbaruah, R.; Athalye, M.; Parikh, P.K.; Chhipa, A.S.; Patel, S.; Apostolopoulos, V. Replicating viral vector-based vaccines for COVID-19: Potential avenue in vaccination arena. Viruses 2022, 14, 759. [Google Scholar] [CrossRef]
  40. Sm Wold, W.; Toth, K. Adenovirus vectors for gene therapy, vaccination and cancer gene therapy. Curr. Gene Ther. 2013, 13, 421–433. [Google Scholar] [CrossRef] [Green Version]
  41. Geisbert, T.W.; Bailey, M.; Hensley, L.; Asiedu, C.; Geisbert, J.; Stanley, D.; Honko, A.; Johnson, J.; Mulangu, S.; Pau, M.G. Recombinant adenovirus serotype 26 (Ad26) and Ad35 vaccine vectors bypass immunity to Ad5 and protect nonhuman primates against ebolavirus challenge. J. Virol. 2011, 85, 4222–4233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Negri, D.R.M.; Michelini, Z.; Bona, R.; Blasi, M.; Filati, P.; Leone, P.; Rossi, A.; Franco, M.; Cara, A. Integrase-defective lentiviral-vector-based vaccine: A new vector for induction of T cell immunity. Expert Opin. Biol. Ther. 2011, 11, 739–750. [Google Scholar] [CrossRef] [PubMed]
  43. Schakowski, F.; Gorschlüter, M.; Junghans, C.; Schroff, M.; Buttgereit, P.; Ziske, C.; Schöttker, B.; König-Merediz, S.A.; Sauerbruch, T.; Wittig, B. A novel minimal-size vector (MIDGE) improves transgene expression in colon carcinoma cells and avoids transfection of undesired DNA. Mol. Ther. 2001, 3, 793–800. [Google Scholar] [CrossRef]
  44. Lundstrom, K. Self-replicating RNA viruses for vaccine development against infectious diseases and cancer. Vaccines 2021, 9, 1187. [Google Scholar] [CrossRef] [PubMed]
  45. Shafaati, M.; Saidijam, M.; Soleimani, M.; Hazrati, F.; Mirzaei, R.; Amirheidari, B.; Tanzadehpanah, H.; Karampoor, S.; Kazemi, S.; Yavari, B. A brief review on DNA vaccines in the era of COVID-19. Future Virol. 2022, 17, 49–66. [Google Scholar] [CrossRef] [PubMed]
  46. Carnes, A.E.; Hodgson, C.P.; Williams, J.A. Inducible Escherichia coli fermentation for increased plasmid DNA production. Biotechnol. Appl. Biochem. 2006, 45, 155–166. [Google Scholar] [PubMed]
  47. Tubulekas, I.; Berglund, P.; Fleeton, M.; Liljeström, P. Alphavirus expression vectors and their use as recombinant vaccines: A minireview. Gene 1997, 190, 191–195. [Google Scholar] [CrossRef] [PubMed]
  48. Leitner, W.W.; Ying, H.; Restifo, N.P. DNA and RNA-based vaccines: Principles, progress and prospects. Vaccine 1999, 18, 765–777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Franck, C.O.; Fanslau, L.; Bistrovic Popov, A.; Tyagi, P.; Fruk, L. Biopolymer-based carriers for DNA vaccine design. Angew. Chem. Int. Ed. 2021, 60, 13225–13243. [Google Scholar] [CrossRef]
  50. Sasaki, S.; Takeshita, F.; Xin, K.-Q.; Ishii, N.; Okuda, K. Adjuvant formulations and delivery systems for DNA vaccines. Methods 2003, 31, 243–254. [Google Scholar] [CrossRef]
  51. Almeida, A.M.; Queiroz, J.A.; Sousa, F.; Sousa, Â. Minicircle DNA: The future for DNA-based vectors? Trends Biotechnol. 2020, 38, 1047–1051. [Google Scholar] [CrossRef]
  52. Eusébio, D.; Almeida, A.M.; Alves, J.M.; Maia, C.J.; Queiroz, J.A.; Sousa, F.; Sousa, Â. The performance of minicircle DNA versus parental plasmid in p53 gene delivery into HPV-18-Infected cervical cancer cells. Nucleic Acid Ther. 2021, 31, 82–91. [Google Scholar] [CrossRef]
  53. Arévalo-Soliz, L.M.; Hardee, C.L.; Fogg, J.M.; Corman, N.R.; Noorbakhsh, C.; Zechiedrich, L. Improving therapeutic potential of non-viral minimized DNA vectors. Cell Gene Ther. Insights 2020, 6, 1489. [Google Scholar] [CrossRef] [PubMed]
  54. Hardee, C.L.; Arévalo-Soliz, L.M.; Hornstein, B.D.; Zechiedrich, L. Advances in non-viral DNA vectors for gene therapy. Genes 2017, 8, 65. [Google Scholar] [CrossRef] [Green Version]
  55. Schakowski, F.; Gorschlüter, M.; Buttgereit, P.; Märten, A.; Lilienfeld-Toal, M.V.; Junghans, C.; Schroff, M.; König-Merediz, S.A.; Ziske, C.; Strehl, J.; et al. Minimal Size MIDGE Vectors Improve Transgene Expression In Vivo. In Vivo 2007, 21, 17. [Google Scholar]
  56. Wang, H.-S.; Chen, Z.-J.; Zhang, G.; Ou, X.-L.; Yang, X.-L.; Wong, C.K.C.; Giesy, J.P.; Du, J.; Chen, S.-Y. A Novel Micro-Linear Vector for In Vitro and In Vivo Gene Delivery and Its Application for EBV Positive Tumors. PLoS ONE 2012, 7, e47159. [Google Scholar] [CrossRef] [Green Version]
  57. Karda, R.; Counsell, J.R.; Karbowniczek, K.; Caproni, L.J.; Tite, J.P.; Waddington, S.N. Production of lentiviral vectors using novel, enzymatically produced, linear DNA. Gene Ther. 2019, 26, 86–92. [Google Scholar] [CrossRef] [Green Version]
  58. Scott, V.L.; Patel, A.; Villarreal, D.O.; Hensley, S.E.; Ragwan, E.; Yan, J.; Sardesai, N.Y.; Rothwell, P.J.; Extance, J.P.; Caproni, L.J. Novel synthetic plasmid and Doggybone™ DNA vaccines induce neutralizing antibodies and provide protection from lethal influenza challenge in mice. Hum. Vaccines Immunother. 2015, 11, 1972–1982. [Google Scholar] [CrossRef]
  59. Grunwald, T.; Ulbert, S. Improvement of DNA vaccination by adjuvants and sophisticated delivery devices: Vaccine-platforms for the battle against infectious diseases. Clin. Exp. Vaccine Res. 2015, 4, 1–10. [Google Scholar] [CrossRef] [Green Version]
  60. Chellasamy, G.; Kiriyanthan, R.M.; Govindaraju, S.; Radha, A.; Yun, K. Recent trends in the development of vaccine technologies to combat pandemic outbreaks and challenges. In Pandemic Outbreaks in the 21st Century; Elsevier: Amsterdam, The Netherlands, 2021; pp. 235–243. [Google Scholar]
  61. Tews, B.A.; Meyers, G. Self-Replicating RNA vaccines. Methods Mol. Biol. 2017, 1499, 15–35. [Google Scholar] [PubMed]
  62. Rosa, S.S.; Prazeres, D.M.F.; Azevedo, A.M.; Marques, M.P.C. mRNA vaccines manufacturing: Challenges and bottlenecks. Vaccine 2021, 39, 2190–2200. [Google Scholar] [CrossRef] [PubMed]
  63. Chaudhary, N.; Weissman, D.; Whitehead, K.A. mRNA vaccines for infectious diseases: Principles, delivery and clinical translation. Nat. Rev. Drug Discov. 2021, 20, 817–838. [Google Scholar] [CrossRef]
  64. Kim, S.C.; Sekhon, S.S.; Shin, W.-R.; Ahn, G.; Cho, B.-K.; Ahn, J.-Y.; Kim, Y.-H. Modifications of mRNA vaccine structural elements for improving mRNA stability and translation efficiency. Mol. Cell. Toxicol. 2022, 18, 1–8. [Google Scholar] [CrossRef] [PubMed]
  65. Pardi, N.; Hogan, M.J.; Porter, F.W.; Weissman, D. mRNA vaccines—A new era in vaccinology. Nat. Rev. Drug Discov. 2018, 17, 261–279. [Google Scholar] [CrossRef] [Green Version]
  66. Bloom, K.; van den Berg, F.; Arbuthnot, P. Self-amplifying RNA vaccines for infectious diseases. Gene Ther. 2021, 28, 117–129. [Google Scholar] [CrossRef] [PubMed]
  67. Perri, S.; Greer, C.E.; Thudium, K.; Doe, B.; Legg, H.; Liu, H.; Romero, R.E.; Tang, Z.; Bin, Q.; Dubensky, T.W., Jr.; et al. An alphavirus replicon particle chimera derived from venezuelan equine encephalitis and sindbis viruses is a potent gene-based vaccine delivery vector. J. Virol. 2003, 77, 10394–10403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Hajj, K.A.; Whitehead, K.A. Tools for translation: Non-viral materials for therapeutic mRNA delivery. Nat. Rev. Mater. 2017, 2, 17056. [Google Scholar] [CrossRef]
  69. Caillaud, M.-C. Anionic lipids: A pipeline connecting key players of plant cell division. Front. Plant Sci. 2019, 10, 419. [Google Scholar] [CrossRef] [PubMed]
  70. Kowalski, P.S.; Rudra, A.; Miao, L.; Anderson, D.G. Delivering the messenger: Advances in technologies for therapeutic mRNA delivery. Mol. Ther. 2019, 27, 710–728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Han, X.; Zhang, H.; Butowska, K.; Swingle, K.L.; Alameh, M.-G.; Weissman, D.; Mitchell, M.J. An ionizable lipid toolbox for RNA delivery. Nat. Commun. 2021, 12, 7233. [Google Scholar] [CrossRef]
  72. Hou, X.; Zaks, T.; Langer, R.; Dong, Y. Lipid nanoparticles for mRNA delivery. Nat. Rev. Mater. 2021, 6, 1078–1094. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, C.; Ma, Y.; Zhang, J.; Kuo, J.C.-T.; Zhang, Z.; Xie, H.; Zhu, J.; Liu, T. Modification of lipid-based nanoparticles: An efficient delivery system for nucleic acid-based immunotherapy. Molecules 2022, 27, 1943. [Google Scholar] [CrossRef]
  74. Lynn, D.M.; Langer, R. Degradable poly (β-amino esters): Synthesis, characterization, and self-assembly with plasmid DNA. J. Am. Chem. Soc. 2000, 122, 10761–10768. [Google Scholar] [CrossRef]
  75. Ko, Y.T.; Kale, A.; Hartner, W.C.; Papahadjopoulos-Sternberg, B.; Torchilin, V.P. Self-assembling micelle-like nanoparticles based on phospholipid–polyethyleneimine conjugates for systemic gene delivery. J. Control. Release 2009, 133, 132–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Mitra, A.; Dey, B. Chitosan microspheres in novel drug delivery systems. Indian J. Pharm. Sci. 2011, 73, 355. [Google Scholar]
  77. Ramirez, J.E.V.; Sharpe, L.A.; Peppas, N.A. Current state and challenges in developing oral vaccines. Adv. Drug Deliv. Rev. 2017, 114, 116–131. [Google Scholar] [CrossRef] [PubMed]
  78. Cecil, R.L.F.; Goldman, L.; Schafer, A.I. Goldman’s Cecil Medicine, Expert Consult Premium Edition—Enhanced Online Features and Print, Single Volume, 24: Goldman’s Cecil Medicine; Elsevier Health Sciences: Philadelphia, PA, USA, 2012; Volume 1. [Google Scholar]
  79. Rich, R.R.; Fleisher, T.A.; Shearer, W.T.; Schroeder, H.W., Jr.; Frew, A.J.; Weyand, C.M. Clinical Immunology E-Book: Principles and Practice; Elsevier Health Sciences: Philadelphia, PA, USA, 2012. [Google Scholar]
  80. Golos, M.; Eliakim-Raz, N.; Stern, A.; Leibovici, L.; Paul, M. Conjugated pneumococcal vaccine versus polysaccharide pneumococcal vaccine for prevention of pneumonia and invasive pneumococcal disease in immunocompetent and immunocompromised adults and children. Cochrane Database Syst. Rev. 2019, 2019, CD012306. [Google Scholar] [CrossRef]
  81. Pichichero, M.E. Protein carriers of conjugate vaccines: Characteristics, development, and clinical trials. Hum. Vaccines Immunother. 2013, 9, 2505–2523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Giannini, G.; Rappuoli, R.; Ratti, G. The amino-acid sequence of two non-toxic mutants of diphtheria toxin: CRM45 and CRM197. Nucleic Acids Res. 1984, 12, 4063–4069. [Google Scholar] [CrossRef] [PubMed]
  83. Sahu, U.; Khare, P. Chapter 2—Evolution and development of vaccines against major human infections. In System Vaccinology; Prajapati, V., Ed.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 17–30. [Google Scholar]
  84. Harrison, L.H.; Mohan, N.; Kirkpatrick, P. Meningococcal group A, C, Y and W-135 conjugate vaccine. Nat. Rev. Drug Discov. 2010, 9, 429. [Google Scholar] [CrossRef]
  85. Ruan, M.R.; Akkoyunlu, M.; Grubb, A.; Forsgren, A. Protein D of Haemophilus influenzae. A novel bacterial surface protein with affinity for human IgD. J. Immunol. 1990, 145, 3379–3384. [Google Scholar] [CrossRef] [PubMed]
  86. Brisse, M.; Vrba, S.M.; Kirk, N.; Liang, Y.; Ly, H. Emerging concepts and technologies in vaccine development. Front. Immunol. 2020, 11, 583077. [Google Scholar] [CrossRef]
  87. Donaldson, B.; Al-Barwani, F.; Young, V.; Scullion, S.; Ward, V.; Young, S. Virus-like particles, a versatile subunit vaccine platform. In Subunit Vaccine Delivery; Springer: New York, NY, USA, 2015; pp. 159–180. [Google Scholar]
  88. Yadav, D.; Yadav, N.; Khurana, S.M. Vaccines: Present Status and Application; Elsevier: Amsterdam, The Netherlands, 2013; pp. 491–508. [Google Scholar]
  89. Jones, R.G.A.; Liu, Y.; Rigsby, P.; Sesardic, D. An improved method for development of toxoid vaccines and antitoxins. J. Immunol. Methods 2008, 337, 42–48. [Google Scholar] [CrossRef]
  90. Huisman, W.; Martina, B.E.; Rimmelzwaan, G.F.; Gruters, R.A.; Osterhaus, A.D. Vaccine-induced enhancement of viral infections. Vaccine 2009, 27, 505–512. [Google Scholar] [CrossRef]
  91. Yang, S.; Li, Y.; Dai, L.; Wang, J.; He, P.; Li, C.; Fang, X.; Wang, C.; Zhao, X.; Huang, E.; et al. Safety and immunogenicity of a recombinant tandem-repeat dimeric RBD-based protein subunit vaccine (ZF2001) against COVID-19 in adults: Two randomised, double-blind, placebo-controlled, phase 1 and 2 trials. Lancet Infect. Dis. 2021, 21, 1107–1119. [Google Scholar] [CrossRef]
  92. He, Y.; Li, J.; Heck, S.; Lustigman, S.; Jiang, S. Antigenic and immunogenic characterization of recombinant baculovirus-expressed severe acute respiratory syndrome coronavirus spike protein: Implication for vaccine design. J. Virol. 2006, 80, 5757–5767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Li, J.; Ulitzky, L.; Silberstein, E.; Taylor, D.R.; Viscidi, R. Immunogenicity and protection efficacy of monomeric and trimeric recombinant SARS coronavirus spike protein subunit vaccine candidates. Viral Immunol. 2013, 26, 126–132. [Google Scholar] [CrossRef] [Green Version]
  94. Wong, S.-S.; Webby Richard, J. Traditional and new influenza vaccines. Clin. Microbiol. Rev. 2013, 26, 476–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zhang, N.; Zheng, B.-J.; Lu, L.; Zhou, Y.; Jiang, S.; Du, L. Advancements in the development of subunit influenza vaccines. Microbes Infect. 2015, 17, 123–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Nobusawa, E. Structure and function of the hemagglutinin of influenza viruses. Nihon Rinsho Jpn. J. Clin. Med. 1997, 55, 2562–2569. [Google Scholar]
  97. Lee, Y.; Lee, Y.S.; Cho, S.Y.; Kwon, H.-J. Chapter four—perspective of peptide vaccine composed of epitope peptide, CpG-DNA, and liposome complex without carriers. In Advances in Protein Chemistry and Structural Biology; Donev, R., Ed.; Elsevier: Amsterdam, The Netherlands, 2015; Volume 99, pp. 75–97. [Google Scholar]
  98. Van der Weken, H.; Cox, E.; Devriendt, B. Advances in oral subunit vaccine design. Vaccines 2021, 9, 1. [Google Scholar] [CrossRef] [PubMed]
  99. Deng, J.; Cai, W.; Jin, F. A novel oil-in-water emulsion as a potential adjuvant for influenza vaccine: Development, characterization, stability and in vivo evaluation. Int. J. Pharm. 2014, 468, 187–195. [Google Scholar] [CrossRef] [PubMed]
  100. Sharma, T.; Kumar, G.S.; Chon, B.H.; Sangwai, J.S. Thermal stability of oil-in-water Pickering emulsion in the presence of nanoparticle, surfactant, and polymer. J. Ind. Eng. Chem. 2015, 22, 324–334. [Google Scholar] [CrossRef]
  101. Siva, S.P.; Kow, K.-W.; Chan, C.-H.; Tang, S.Y.; Ho, Y.K. Prediction of droplet sizes for oil-in-water emulsion systems assisted by ultrasound cavitation: Transient scaling law based on dynamic breakup potential. Ultrason. Sonochem. 2019, 55, 348–358. [Google Scholar] [CrossRef] [PubMed]
  102. Sun, Z.; Li, W.; Lenzo, J.; Holden, J.; McCullough, M.; O’Connor, A.; O’Brien-Simpson, N. The potential of calcium phosphate nanoparticles as adjuvants and vaccine delivery vehicles. Front. Mater. 2021, 8, 788373. [Google Scholar] [CrossRef]
  103. Ferrando, R.M.; Lay, L.; Polito, L. Gold nanoparticle-based platforms for vaccine development. Drug Discov. Today Technol. 2020, 38, 57–67. [Google Scholar] [CrossRef] [PubMed]
  104. Zhou, H.; Ge, J.; Miao, Q.; Zhu, R.; Wen, L.; Zeng, J.; Gao, M. Biodegradable inorganic nanoparticles for cancer theranostics: Insights into the degradation behavior. Bioconjug Chem. 2020, 31, 315–331. [Google Scholar] [CrossRef]
  105. Duewell, P.; Kisser, U.; Heckelsmiller, K.; Hoves, S.; Stoitzner, P.; Koernig, S.; Morelli, A.B.; Clausen, B.E.; Dauer, M.; Eigler, A. ISCOMATRIX adjuvant combines immune activation with antigen delivery to dendritic cells in vivo leading to effective cross-priming of CD8+ T cells. J. Immunol. 2011, 187, 55–63. [Google Scholar] [CrossRef] [Green Version]
  106. Lövgren Bengtsson, K.; Morein, B.; Osterhaus, A.D.M.E. ISCOM technology-based Matrix M™ adjuvant: Success in future vaccines relies on formulation. Expert Rev. Vaccines 2011, 10, 401–403. [Google Scholar] [CrossRef] [PubMed]
  107. Levast, B.; Awate, S.; Babiuk, L.; Mutwiri, G.; Gerdts, V.; van Drunen Littel-van den Hurk, S. Vaccine potentiation by combination adjuvants. Vaccines 2014, 2, 297–322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Brewer, J.M.; Tetley, L.; Richmond, J.; Liew, F.Y.; Alexander, J. Lipid vesicle size determines the Th1 or Th2 response to entrapped antigen. J. Immunol. 1998, 161, 4000–4007. [Google Scholar] [CrossRef] [PubMed]
  109. Attia, M.A.; Essa, E.A.; Elebyary, T.T.; Faheem, A.M.; Elkordy, A.A. Brief on recent application of liposomal vaccines for lower respiratory tract viral infections: From influenza to COVID-19 vaccines. Pharmaceuticals 2021, 14, 1173. [Google Scholar] [CrossRef]
  110. Gregoriadis, G. Liposomes and mRNA: Two technologies together create a COVID-19 vaccine. Med. Drug Discov. 2021, 12, 100104. [Google Scholar] [CrossRef]
  111. Mann, J.F.S.; Shakir, E.; Carter, K.C.; Mullen, A.B.; Alexander, J.; Ferro, V.A. Lipid vesicle size of an oral influenza vaccine delivery vehicle influences the Th1/Th2 bias in the immune response and protection against infection. Vaccine 2009, 27, 3643–3649. [Google Scholar] [CrossRef]
  112. Corthésy, B.; Bioley, G. Lipid-based particles: Versatile delivery systems for mucosal vaccination against infection. Front. Immunol. 2018, 9, 431. [Google Scholar] [CrossRef] [Green Version]
  113. Zafar, A.; Alruwaili, N.K.; Imam, S.S.; Alotaibi, N.H.; Alharbi, K.S.; Afzal, M.; Ali, R.; Alshehri, S.; Alzarea, S.I.; Elmowafy, M. Bioactive Apigenin loaded oral nano bilosomes: Formulation optimization to preclinical assessment. Saudi Pharm. J. 2021, 29, 269–279. [Google Scholar] [CrossRef]
  114. Adamiak, N.; Krawczyk, K.T.; Locht, C.; Kowalewicz-Kulbat, M. Archaeosomes and gas vesicles as tools for vaccine development. Front. Immunol. 2021, 12, 746235. [Google Scholar] [CrossRef]
  115. Krishnan, L.; Dicaire, C.J.; Patel, G.B.; Sprott, G.D. Archaeosome vaccine adjuvants induce strong humoral, cell-mediated, and memory responses: Comparison to conventional liposomes and alum. Infect. Immun. 2000, 68, 54–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. McCluskie, M.J.; Deschatelets, L.; Krishnan, L. Sulfated archaeal glycolipid archaeosomes as a safe and effective vaccine adjuvant for induction of cell-mediated immunity. Hum. Vaccines Immunother. 2017, 13, 2772–2779. [Google Scholar] [CrossRef] [PubMed]
  117. Wilschut, J. Influenza vaccines: The virosome concept. Immunol. Lett. 2009, 122, 118–121. [Google Scholar] [CrossRef] [PubMed]
  118. Liu, H.; Tu, Z.; Feng, F.; Shi, H.; Chen, K.; Xu, X. Virosome, a hybrid vehicle for efficient and safe drug delivery and its emerging application in cancer treatment. Acta Pharm. 2015, 65, 105–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Patravale, V.; Dandekar, P.; Jain, R. Nanoparticulate Drug Delivery: Perspectives on the Transition from Laboratory to Market; Elsevier: Amsterdam, The Netherlands, 2012. [Google Scholar]
  120. Asadi, K.; Gholami, A. Virosome-based nanovaccines; a promising bioinspiration and biomimetic approach for preventing viral diseases: A review. Int. J. Biol. Macromol. 2021, 182, 648–658. [Google Scholar] [CrossRef] [PubMed]
  121. Bourquin, C.; Anz, D.; Zwiorek, K.; Lanz, A.-L.; Fuchs, S.; Weigel, S.; Wurzenberger, C.; von der Borch, P.; Golic, M.; Moder, S. Targeting CpG oligonucleotides to the lymph node by nanoparticles elicits efficient antitumoral immunity. J. Immunol. 2008, 181, 2990–2998. [Google Scholar] [CrossRef] [Green Version]
  122. Sudheesh, M.S.; Vyas, S.P.; Kohli, D.V. Nanoparticle-based immunopotentiation via tetanus toxoid-loaded gelatin and aminated gelatin nanoparticles. Drug Deliv. 2011, 18, 320–330. [Google Scholar] [CrossRef]
  123. Hong, S.; Choi, D.W.; Kim, H.N.; Park, C.G.; Lee, W.; Park, H.H. Protein-based nanoparticles as drug delivery systems. Pharmaceutics 2020, 12, 604. [Google Scholar] [CrossRef]
  124. Sahoo, N.; Sahoo, R.K.; Biswas, N.; Guha, A.; Kuotsu, K. Recent advancement of gelatin nanoparticles in drug and vaccine delivery. Int. J. Biol. Macromol. 2015, 81, 317–331. [Google Scholar] [CrossRef]
  125. Zhu, G.; Lynn, G.M.; Jacobson, O.; Chen, K.; Liu, Y.; Zhang, H.; Ma, Y.; Zhang, F.; Tian, R.; Ni, Q. Albumin/vaccine nanocomplexes that assemble in vivo for combination cancer immunotherapy. Nat. Commun. 2017, 8, 1954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Zhu, Y.; Xue, J.; Chen, W.; Bai, S.; Zheng, T.; He, C.; Guo, Z.; Jiang, M.; Du, G.; Sun, X. Albumin-biomineralized nanoparticles to synergize phototherapy and immunotherapy against melanoma. J. Control. Release 2020, 322, 300–311. [Google Scholar] [CrossRef]
  127. Hassanin, I.; Elzoghby, A. Albumin-based nanoparticles: A promising strategy to overcome cancer drug resistance. Cancer Drug Resist. 2020, 3, 930. [Google Scholar] [CrossRef] [PubMed]
  128. Li, F.; Chen, Y.; Liu, S.; Pan, X.; Liu, Y.; Zhao, H.; Yin, X.; Yu, C.; Kong, W.; Zhang, Y. The effect of size, dose, and administration route on zein nanoparticle immunogenicity in BALB/c mice. Int. J. Nanomed. 2019, 14, 9917–9928. [Google Scholar] [CrossRef] [Green Version]
  129. Li, M.; Yu, M. Development of a nanoparticle delivery system based on zein/polysaccharide complexes. J. Food Sci. 2020, 85, 4108–4117. [Google Scholar] [CrossRef] [PubMed]
  130. Ko, E.-J.; Kang, S.-M. Immunology and efficacy of MF59-adjuvanted vaccines. Hum. Vaccines Immunother. 2018, 14, 3041–3045. [Google Scholar] [CrossRef] [Green Version]
  131. Grigoryan, L.; Lee, A.; Walls, A.C.; Lai, L.; Franco, B.; Arunachalam, P.S.; Feng, Y.; Luo, W.; Vanderheiden, A.; Floyd, K. Adjuvanting a subunit SARS-CoV-2 vaccine with clinically relevant adjuvants induces durable protection in mice. NPJ Vaccines 2022, 7, 55. [Google Scholar] [CrossRef] [PubMed]
  132. Lodaya, R.; Gregory, S.; Amiji, M.; O’Hagan, D. Overview of Vaccine Adjuvants. In Practical Aspects of Vaccine Development; Elsevier: Amsterdam, The Netherlands, 2021; pp. 9–25. [Google Scholar]
  133. De Rosa, S.C.; Cohen, K.W.; Bonaparte, M.; Fu, B.; Garg, S.; Gerard, C.; Goepfert, P.A.; Huang, Y.; Larocque, D.; McElrath, M.J. Whole-blood cytokine secretion assay as a high-throughput alternative for assessing the cell-mediated immunity profile after two doses of an adjuvanted SARS-CoV-2 recombinant protein vaccine candidate. Clin. Transl. Immunol. 2022, 11, e1360. [Google Scholar] [CrossRef]
  134. Klucker, M.F.; Dalençon, F.; Probeck, P.; Haensler, J. AF03, an alternative squalene emulsion-based vaccine adjuvant prepared by a phase inversion temperature method. J. Pharm. Sci. 2012, 101, 4490–4500. [Google Scholar] [CrossRef]
  135. Gao, L.; Li, J.; Song, T. Poly lactic-co-glycolic acid-based nanoparticles as delivery systems for enhanced cancer immunotherapy. Front. Chem. 2022, 10, 973666. [Google Scholar] [CrossRef]
  136. Bennewitz, N.L.; Babensee, J.E. The effect of the physical form of poly (lactic-co-glycolic acid) carriers on the humoral immune response to co-delivered antigen. Biomaterials 2005, 26, 2991–2999. [Google Scholar] [CrossRef] [PubMed]
  137. Broaders, K.E.; Cohen, J.A.; Beaudette, T.T.; Bachelder, E.M.; Fréchet, J.M.J. Acetalated dextran is a chemically and biologically tunable material for particulate immunotherapy. Proc. Natl. Acad. Sci. USA 2009, 106, 5497–5502. [Google Scholar] [CrossRef]
  138. Chen, N.; Collier, M.A.; Gallovic, M.D.; Collins, G.C.; Sanchez, C.C.; Fernandes, E.Q.; Bachelder, E.M.; Ainslie, K.M. Degradation of acetalated dextran can be broadly tuned based on cyclic acetal coverage and molecular weight. Int. J. Pharm. 2016, 512, 147–157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Peres, C.; Matos, A.I.; Conniot, J.; Sainz, V.; Zupančič, E.; Silva, J.M.; Graca, L.; Gaspar, R.S.; Preat, V.; Florindo, H.F. Poly (lactic acid)-based particulate systems are promising tools for immune modulation. Acta Biomater. 2017, 48, 41–57. [Google Scholar] [CrossRef]
  140. Baek, S.-W.; Kim, J.H.; Song, D.H.; Kim, D.-S.; Park, C.G.; Han, D.K. Enhanced Mechanical Properties and Anti–Inflammation of Poly (L–Lactic Acid) by Stereocomplexes of PLLA/PDLA and Surface–Modified Magnesium Hydroxide Nanoparticles. Polymers 2022, 14, 3790. [Google Scholar] [CrossRef] [PubMed]
  141. Wen, Z.-S.; Xu, Y.-L.; Zou, X.-T.; Xu, Z.-R. Chitosan nanoparticles act as an adjuvant to promote both Th1 and Th2 immune responses induced by ovalbumin in mice. Mar. Drugs 2011, 9, 1038–1055. [Google Scholar] [CrossRef] [Green Version]
  142. Lemke, C.D.; Graham, J.B.; Geary, S.M.; Zamba, G.; Lubaroff, D.M.; Salem, A.K. Chitosan is a surprising negative modulator of cytotoxic CD8+ T cell responses elicited by adenovirus cancer vaccines. Mol. Pharm. 2011, 8, 1652–1661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Lin, Y.; Wang, X.; Huang, X.; Zhang, J.; Xia, N.; Zhao, Q. Calcium phosphate nanoparticles as a new generation vaccine adjuvant. Expert Rev. Vaccines 2017, 16, 895–906. [Google Scholar] [CrossRef]
  144. Wang, X.; Li, X.; Yoshiyuki, K.; Watanabe, Y.; Sogo, Y.; Ohno, T.; Tsuji, N.M.; Ito, A. Comprehensive mechanism analysis of mesoporous-silica-nanoparticle-induced cancer immunotherapy. Adv. Healthc. Mater. 2016, 5, 1169–1176. [Google Scholar] [CrossRef] [PubMed]
  145. Bharti, C.; Nagaich, U.; Pal, A.K.; Gulati, N. Mesoporous silica nanoparticles in target drug delivery system: A review. Int. J. Pharm. Investig. 2015, 5, 124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Anselmo, A.C.; Mitragotri, S. Nanoparticles in the clinic. Bioeng. Transl. Med. 2016, 1, 10–29. [Google Scholar] [CrossRef]
  147. Zhou, Q.; Zhang, Y.; Du, J.; Li, Y.; Zhou, Y.; Fu, Q.; Zhang, J.; Wang, X.; Zhan, L. Different-sized gold nanoparticle activator/antigen increases dendritic cells accumulation in liver-draining lymph nodes and CD8+ T cell responses. ACS Nano 2016, 10, 2678–2692. [Google Scholar] [CrossRef] [PubMed]
  148. Dey, A.K.; Gonon, A.; Pécheur, E.-I.; Pezet, M.; Villiers, C.; Marche, P.N. Impact of gold nanoparticles on the functions of macrophages and dendritic cells. Cells 2021, 10, 96. [Google Scholar] [CrossRef]
  149. Balfourier, A.; Luciani, N.; Wang, G.; Lelong, G.; Ersen, O.; Khelfa, A.; Alloyeau, D.; Gazeau, F.; Carn, F. Unexpected intracellular biodegradation and recrystallization of gold nanoparticles. Proc. Natl. Acad. Sci. USA 2020, 117, 103–113. [Google Scholar] [CrossRef]
  150. Nguyen, B.; Tolia, N.H. Protein-based antigen presentation platforms for nanoparticle vaccines. NPJ Vaccines 2021, 6, 70. [Google Scholar] [CrossRef] [PubMed]
  151. Kianfar, E. Protein nanoparticles in drug delivery: Animal protein, plant proteins and protein cages, albumin nanoparticles. J. Nanobiotechnol. 2021, 19, 159. [Google Scholar] [CrossRef] [PubMed]
  152. Xu, H.; Shen, L.; Xu, L.; Yang, Y. Controlled delivery of hollow corn protein nanoparticles via non-toxic crosslinking: In vivo and drug loading study. Biomed. Microdevices 2015, 17, 8. [Google Scholar] [CrossRef] [PubMed]
  153. Verma, D.; Gulati, N.; Kaul, S.; Mukherjee, S.; Nagaich, U. Protein based nanostructures for drug delivery. J. Pharm. 2018, 2018, 9285854. [Google Scholar] [CrossRef] [Green Version]
  154. Marasini, N.; Ghaffar, K.A.; Skwarczynski, M.; Toth, I. Chapter twelve—Liposomes as a vaccine delivery system. In Micro and Nanotechnology in Vaccine Development; Skwarczynski, M., Toth, I., Eds.; William Andrew Elsevier: Amsterdam, The Netherlands, 2017; pp. 221–239. [Google Scholar]
  155. Kisakova, L.A.; Apartsin, E.K.; Nizolenko, L.F.; Karpenko, L.I. Dendrimer-mediated delivery of DNA and RNA vaccines. Pharmaceutics 2023, 15, 1106. [Google Scholar] [CrossRef] [PubMed]
  156. Gutjahr, A.; Phelip, C.; Coolen, A.L.; Monge, C.; Boisgard, A.S.; Paul, S.; Verrier, B. Biodegradable polymeric nanoparticles-based vaccine adjuvants for lymph nodes targeting. Vaccines 2016, 4, 34. [Google Scholar] [CrossRef] [Green Version]
  157. Vaid, R.; Yildirim, E.; Pasquinelli, M.A.; King, M.W. Hydrolytic degradation of polylactic acid fibers as a function of pH and exposure time. Molecules 2021, 26, 7554. [Google Scholar] [CrossRef] [PubMed]
  158. Poon, C.; Patel, A.A. Organic and inorganic nanoparticle vaccines for prevention of infectious diseases. Nano Express 2020, 1, 012001. [Google Scholar] [CrossRef]
  159. Lang, S.; Huang, X. Carbohydrate conjugates in vaccine developments. Front. Chem. 2020, 8, 284. [Google Scholar] [CrossRef]
  160. de Carvalho Lima, E.N.; Diaz, R.S.; Justo, J.F.; Castilho Piqueira, J.R. Advances and perspectives in the use of carbon nanotubes in vaccine development. Int. J. Nanomed. 2021, 16, 5411–5435. [Google Scholar] [CrossRef]
  161. Bamrungsap, S.; Zhao, Z.; Chen, T.; Wang, L.; Li, C.; Fu, T.; Tan, W. Nanotechnology in therapeutics: A focus on nanoparticles as a drug delivery system. Nanomedicine 2012, 7, 1253–1271. [Google Scholar] [CrossRef] [PubMed]
  162. Liu, C.; Liu, X.; Xiang, X.; Pang, X.; Chen, S.; Zhang, Y.; Ren, E.; Zhang, L.; Liu, X.; Lv, P.; et al. A nanovaccine for antigen self-presentation and immunosuppression reversal as a personalized cancer immunotherapy strategy. Nat. Nanotechnol. 2022, 17, 531–540. [Google Scholar] [CrossRef] [PubMed]
  163. Das, A.; Ali, N. Nanovaccine: An emerging strategy. Expert Rev. Vaccines 2021, 20, 1273–1290. [Google Scholar] [CrossRef]
  164. Maina, T.W.; Grego, E.A.; Boggiatto, P.M.; Sacco, R.E.; Narasimhan, B.; McGill, J.L. Applications of Nanovaccines for Disease Prevention in Cattle. Front. Bioeng. Biotechnol. 2020, 8, 608050. [Google Scholar] [CrossRef]
  165. Azharuddin, M.; Zhu, G.H.; Sengupta, A.; Hinkula, J.; Slater, N.K.H.; Patra, H.K. Nano toolbox in immune modulation and nanovaccines. Trends Biotechnol. 2022, 40, 1195–1212. [Google Scholar] [CrossRef]
  166. Kim, L.; Martinez, C.J.; Hodgson, K.A.; Trager, G.R.; Brandl, J.R.; Sandefer, E.P.; Doll, W.J.; Liebowitz, D.; Tucker, S.N. Systemic and mucosal immune responses following oral adenoviral delivery of influenza vaccine to the human intestine by radio controlled capsule. Sci. Rep. 2016, 6, 37295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Mudie, D.M.; Amidon, G.L.; Amidon, G.E. Physiological parameters for oral delivery and in vitro testing. Mol. Pharm. 2010, 7, 1388–1405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Zhu, Q.; Talton, J.; Zhang, G.; Cunningham, T.; Wang, Z.; Waters, R.C.; Kirk, J.; Eppler, B.; Klinman, D.M.; Sui, Y.; et al. Large intestine–targeted, nanoparticle-releasing oral vaccine to control genitorectal viral infection. Nat. Med. 2012, 18, 1291–1296. [Google Scholar] [CrossRef] [Green Version]
  169. Le, T.; Aguilar, B.; Mangal, J.L.; Acharya, A.P. Oral drug delivery for immunoengineering. Bioeng. Transl. Med. 2022, 7, e10243. [Google Scholar] [CrossRef] [PubMed]
  170. Ahadian, S.; Finbloom, J.A.; Mofidfar, M.; Diltemiz, S.E.; Nasrollahi, F.; Davoodi, E.; Hosseini, V.; Mylonaki, I.; Sangabathuni, S.; Montazerian, H.; et al. Micro and nanoscale technologies in oral drug delivery. Adv. Drug Deliv. Rev. 2020, 157, 37–62. [Google Scholar] [CrossRef]
  171. Coffey, J.W.; Gaiha, G.D.; Traverso, G. Oral biologic delivery: Advances toward oral subunit, DNA, and mRNA vaccines and the potential for mass vaccination during pandemics. Annu. Rev. Pharmacol. Toxicol. 2021, 61, 517–540. [Google Scholar] [CrossRef]
  172. de Santa Barbara, P.; van den Brink, G.R.; Roberts, D.J. Development and differentiation of the intestinal epithelium. Cell. Mol. Life Sci. CMLS 2003, 60, 1322–1332. [Google Scholar] [CrossRef] [PubMed]
  173. Roopenian, D.C.; Akilesh, S. FcRn: The neonatal Fc receptor comes of age. Nat. Rev. Immunol. 2007, 7, 715–725. [Google Scholar] [CrossRef]
  174. Yang, S.; Yu, M. Role of goblet cells in intestinal barrier and mucosal imunity. J. Inflamm. Res. 2021, 14, 3171–3183. [Google Scholar] [CrossRef]
  175. Jung, C.; Hugot, J.-P.; Barreau, F. Peyer’s patches: The immune sensors of the intestine. Int. J. Inflamm. 2010, 2010, 823710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Homayun, B.; Lin, X.; Choi, H.-J. Challenges and Recent Progress in Oral Drug Delivery Systems for Biopharmaceuticals. Pharmaceutics 2019, 11, 129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. De Smet, R.; Allais, L.; Cuvelier, C.A. Recent advances in oral vaccine development: Yeast-derived β-glucan particles. Hum. Vaccines Immunother. 2014, 10, 1309–1318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. David, J.; Mastropietro, K.P.; Omidian, H. Polymers in Oral Drug Delivery; Elsevier: Amsterdam, The Netherlands, 2017; pp. 430–444. [Google Scholar]
  179. dos Santos, J.; da Silva, G.S.; Velho, M.C.; Beck, R.C. Eudragit®: A versatile family of polymers for hot melt extrusion and 3D printing processes in pharmaceutics. Pharmaceutics 2021, 13, 1424. [Google Scholar] [CrossRef] [PubMed]
  180. Franco, P.; De Marco, I. Eudragit: A Novel Carrier for Controlled Drug Delivery in Supercritical Antisolvent Coprecipitation. Polymers 2020, 12, 234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Snejdrova Eva, D.M. Pharmaceutical applications of plasticized polymers. Recent Adv. Plast. 2012, 21, 23–34. [Google Scholar]
  182. Vidal-Romero, G.; Rocha-Pérez, V.; Zambrano-Zaragoza, M.L.; Del Real, A.; Martínez-Acevedo, L.; Galindo-Pérez, M.J.; Quintanar-Guerrero, D. Development and characterization of pH-dependent cellulose acetate phthalate nanofibers by electrospinning technique. Nanomaterials 2021, 11, 3202. [Google Scholar] [CrossRef]
  183. Obara, S.; Tanno, F.K.; Sarode, A. Properties and applications of hypromellose acetate succinate (HPMCAS) for solubility enhancement using melt extrusion. In Melt Extrusion: Materials, Technology and Drug Product Design; Repka, M.A., Langley, N., DiNunzio, J., Eds.; Springer: New York, NY, USA, 2013; pp. 107–121. [Google Scholar]
  184. Bagre, A.P.; Jain, K.; Jain, N.K. Alginate coated chitosan core shell nanoparticles for oral delivery of enoxaparin: In vitro and in vivo assessment. Int. J. Pharm. 2013, 456, 31–40. [Google Scholar] [CrossRef] [PubMed]
  185. Lee, K.Y.; Mooney, D.J. Alginate: Properties and biomedical applications. Prog. Polym. Sci. 2012, 37, 106–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Subramanian, D.A.; Langer, R.; Traverso, G. Mucus interaction to improve gastrointestinal retention and pharmacokinetics of orally administered nano-drug delivery systems. J. Nanobiotechnol. 2022, 20, 362. [Google Scholar] [CrossRef] [PubMed]
  187. Bernkop-Schnürch, A. Mucoadhesive systems in oral drug delivery. Drug Discov. Today Technol. 2005, 2, 83–87. [Google Scholar] [CrossRef]
  188. Boddupalli, B.M.; Mohammed, Z.N.; Nath, R.A.; Banji, D. Mucoadhesive drug delivery system: An overview. J. Adv. Pharm. Technol. Res. 2010, 1, 381–387. [Google Scholar] [CrossRef] [Green Version]
  189. Cho, C.-S.; Hwang, S.-K.; Gu, M.-J.; Kim, C.-G.; Kim, S.-K.; Ju, D.-B.; Yun, C.-H.; Kim, H.-J. Mucosal vaccine delivery using mucoadhesive polymer particulate systems. Tissue Eng. Regen. Med. 2021, 18, 693–712. [Google Scholar] [CrossRef] [PubMed]
  190. Shaikh, R.; Singh, T.R.R.; Garland, M.J.; Woolfson, A.D.; Donnelly, R.F. Mucoadhesive drug delivery systems. J. Pharm. Bioallied Sci. 2011, 3, 89. [Google Scholar]
  191. Pérez-González, G.L.; Villarreal-Gómez, L.J.; Serrano-Medina, A.; Torres-Martínez, E.J.; Cornejo-Bravo, J.M. Mucoadhesive electrospun nanofibers for drug delivery systems: Applications of polymers and the parameters’ roles. Int. J. Nanomed. 2019, 14, 5271–5285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Li, S.; Zhang, H.; Chen, K.; Jin, M.; Vu, S.H.; Jung, S.; He, N.; Zheng, Z.; Lee, M.-S. Application of chitosan/alginate nanoparticle in oral drug delivery systems: Prospects and challenges. Drug Deliv. 2022, 29, 1142–1149. [Google Scholar] [CrossRef]
  193. Li, X.; Kong, X.; Shi, S.; Zheng, X.; Guo, G.; Wei, Y.; Qian, Z. Preparation of alginate coated chitosan microparticles for vaccine delivery. BMC Biotechnol. 2008, 8, 89. [Google Scholar] [CrossRef] [Green Version]
  194. Yang, X.; Forier, K.; Steukers, L.; Van Vlierberghe, S.; Dubruel, P.; Braeckmans, K.; Glorieux, S.; Nauwynck, H.J. Immobilization of pseudorabies virus in porcine tracheal respiratory mucus revealed by single particle tacking. PLoS ONE 2012, 7, e51054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Kaur, G.; Grewal, J.; Jyoti, K.; Jain, U.K.; Chandra, R.; Madan, J. Oral controlled and sustained drug delivery systems: Concepts, advances, preclinical, and clinical status. In Drug Targeting and Stimuli Sensitive Drug Delivery Systems; Elsevier: Amsterdam, The Netherlands, 2018; pp. 567–626. [Google Scholar]
  196. Cevher, E.; Taha, M.A.M.; Orlu, M.; Araman, A. Evaluation of mechanical and mucoadhesive properties of clomiphene citrate gel formulations containing carbomers and their thiolated derivatives. Drug Deliv. 2008, 15, 57–67. [Google Scholar] [CrossRef] [Green Version]
  197. Lai, S.K.; Wang, Y.-Y.; Hanes, J. Mucus-penetrating nanoparticles for drug and gene delivery to mucosal tissues. Adv. Drug Deliv. Rev. 2009, 61, 158–171. [Google Scholar] [CrossRef] [Green Version]
  198. Snook, J.D.; Chesson, C.B.; Peniche, A.G.; Dann, S.M.; Paulucci, A.; Pinchuk, I.V.; Rudra, J.S. Peptide nanofiber–CaCO3 composite microparticles as adjuvant-free oral vaccine delivery vehicles. J. Mater. Chem. B 2016, 4, 1640–1649. [Google Scholar] [CrossRef]
  199. Dong, W.; Ye, J.; Zhou, J.; Wang, W.; Wang, H.; Zheng, X.; Yang, Y.; Xia, X.; Liu, Y. Comparative study of mucoadhesive and mucus-penetrative nanoparticles based on phospholipid complex to overcome the mucus barrier for inhaled delivery of baicalein. Acta Pharm. Sin. B 2020, 10, 1576–1585. [Google Scholar] [CrossRef]
  200. Lamson, N.G.; Berger, A.; Fein, K.C.; Whitehead, K.A. Anionic nanoparticles enable the oral delivery of proteins by enhancing intestinal permeability. Nat. Biomed. Eng. 2020, 4, 84–96. [Google Scholar] [CrossRef]
  201. Lai, S.K.; Wang, Y.-Y.; Cone, R.; Wirtz, D.; Hanes, J. Altering Mucus Rheology to “Solidify” Human Mucus at the Nanoscale. PLoS ONE 2009, 4, e4294. [Google Scholar] [CrossRef] [PubMed]
  202. Xu, Q.; Boylan, N.J.; Suk, J.S.; Wang, Y.-Y.; Nance, E.A.; Yang, J.-C.; McDonnell, P.J.; Cone, R.A.; Duh, E.J.; Hanes, J. Nanoparticle diffusion in, and microrheology of, the bovine vitreous ex vivo. J. Control. Release 2013, 167, 76–84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Wang, Y.Y.; Lai, S.K.; Suk, J.S.; Pace, A.; Cone, R.; Hanes, J. Addressing the PEG mucoadhesivity paradox to engineer nanoparticles that "slip" through the human mucus barrier. Angew. Chem. Int. Ed. Engl. 2008, 47, 9726–9729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Liu, M.; Zhang, J.; Shan, W.; Huang, Y. Developments of mucus penetrating nanoparticles. Asian J. Pharm. Sci. 2015, 10, 275–282. [Google Scholar] [CrossRef] [Green Version]
  205. Ensign, L.M.; Lai, S.K.; Wang, Y.-Y.; Yang, M.; Mert, O.; Hanes, J.; Cone, R. Pretreatment of human cervicovaginal mucus with pluronic F127 enhances nanoparticle penetration without compromising mucus barrier properties to herpes simplex virus. Biomacromolecules 2014, 15, 4403–4409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Dünnhaupt, S.; Kammona, O.; Waldner, C.; Kiparissides, C.; Bernkop-Schnürch, A. Nano-carrier systems: Strategies to overcome the mucus gel barrier. Eur. J. Pharm. Biopharm. 2015, 96, 447–453. [Google Scholar] [CrossRef] [PubMed]
  207. Samuni, Y.; Goldstein, S.; Dean, O.M.; Berk, M. The chemistry and biological activities of N-acetylcysteine. Biochim. Biophys. Acta Gen. Subj. 2013, 1830, 4117–4129. [Google Scholar] [CrossRef]
  208. Partidos, C.D. Intranasal vaccines: Forthcoming challenges. Pharm. Sci. Technol. Today 2000, 3, 273–281. [Google Scholar] [CrossRef] [PubMed]
  209. Xu, H.; Cai, L.; Hufnagel, S.; Cui, Z. Intranasal vaccine: Factors to consider in research and development. Int. J. Pharm. 2021, 609, 121180. [Google Scholar] [CrossRef]
  210. Takaki, H.; Ichimiya, S.; Matsumoto, M.; Seya, T. Mucosal immune response in nasal-associated lymphoid tissue upon intranasal administration by adjuvants. J. Innate Immun. 2018, 10, 515–521. [Google Scholar] [CrossRef] [PubMed]
  211. Scherließ, R. Nasal formulations for drug administration and characterization of nasal preparations in drug delivery. Ther. Deliv. 2020, 11, 183–191. [Google Scholar] [CrossRef]
  212. Ramvikas, M.; Arumugam, M.; Chakrabarti, S.R.; Jaganathan, K.S. Nasal vaccine delivery. In Micro and Nanotechnology in Vaccine Development; Elsevier: Amsterdam, The Netherlands, 2017; pp. 279–301. [Google Scholar]
  213. Jadhav, K.R.; Gambhire, M.N.; Shaikh, I.M.; Kadam, V.J.; Pisal, S.S. Nasal drug delivery system-factors affecting and applications. Curr. Drug Ther. 2007, 2, 27–38. [Google Scholar] [CrossRef]
  214. Tirucherai, G.S.; Pezron, I.; Mitra, A.K. Novel approaches to nasal delivery of peptides and proteins. STP Pharma Sci. 2002, 12, 3–12. [Google Scholar]
  215. Marx, D.; Williams, G.; Birkhoff, M. Intranasal drug administration—An attractive delivery route for some drugs. In Drug Discovery and Development; IntechOpen: Rijeka, Croatia, 2015; pp. 299–320. [Google Scholar]
  216. Upadhyay, S.; Parikh, A.; Joshi, P.; Upadhyay, U.M.; Chotai, N.P. Intranasal drug delivery system-A glimpse to become maestro. J. Appl. Pharm. Sci. 2011, 1, 34–44. [Google Scholar]
  217. Han, K.; Woghiren, O.E.; Priefer, R. Surface tension examination of various liquid oral, nasal, and ophthalmic dosage forms. Chem. Cent. J. 2016, 10, 31. [Google Scholar] [CrossRef] [Green Version]
  218. Gao, M.; Shen, X.; Mao, S. Factors influencing drug deposition in the nasal cavity upon delivery via nasal sprays. J. Pharm. Investig. 2020, 50, 251–259. [Google Scholar] [CrossRef]
  219. Djupesland, P.G. Nasal drug delivery devices: Characteristics and performance in a clinical perspective—A review. Drug Deliv. Transl. Res. 2013, 3, 42–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Food and Drug Administration. FluMist Quadrivalent. Available online: https://www.fda.gov/vaccines-blood-biologics/vaccines/flumist-quadrivalent (accessed on 24 April 2023).
  221. Cardoso, F.M.C.; Petrovajová, D.; Horňáková, T. Viral vaccine stabilizers: Status and trends. Acta Virol. 2017, 61, 231–239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Chavda, V.P.; Vora, L.K.; Pandya, A.K.; Patravale, V.B. Intranasal vaccines for SARS-CoV-2: From challenges to potential in COVID-19 management. Drug Discov. Today 2021, 26, 2619–2636. [Google Scholar] [CrossRef] [PubMed]
  223. Kraan, H.; van Herpen, P.; Kersten, G.; Amorij, J.-P. Development of thermostable lyophilized inactivated polio vaccine. Pharm. Res. 2014, 31, 2618–2629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Herdis, H.; Surachman, M.; Darmawan, I.W.A.; Afifah, A. The role of sucrose as extracellular cryoprotectant in maintaining the Garut rams’ frozen semen quality. In AIP Conference Proceedings; AIP Publishing: New York, NY, USA, 2019. [Google Scholar]
  225. Henriques, P.; Fortuna, A.; Doktorovová, S. Spray dried powders for nasal delivery: Process and formulation considerations. Eur. J. Pharm. Biopharm. 2022, 176, 1–20. [Google Scholar] [CrossRef]
  226. Kanojia, G.; Have, R.t.; Soema, P.C.; Frijlink, H.; Amorij, J.-P.; Kersten, G. Developments in the formulation and delivery of spray dried vaccines. Hum. Vaccines Immunother. 2017, 13, 2364–2378. [Google Scholar] [CrossRef]
  227. Bahamondez-Canas, T.F.; Cui, Z. Intranasal immunization with dry powder vaccines. Eur. J. Pharm. Biopharm. 2018, 122, 167–175. [Google Scholar] [CrossRef]
  228. Nižić Nodilo, L.; Ugrina, I.; Špoljarić, D.; Amidžić Klarić, D.; Jakobušić Brala, C.; Perkušić, M.; Pepić, I.; Lovrić, J.; Saršon, V.; Safundžić Kučuk, M. A dry powder platform for nose-to-brain delivery of dexamethasone: Formulation development and nasal deposition studies. Pharmaceutics 2021, 13, 795. [Google Scholar] [CrossRef] [PubMed]
  229. Zhang, W.; Zhang, L.-J.; Zhan, L.-T.; Zhao, M.; Wu, G.-H.; Si, J.-Y.; Chen, L.; Lin, X.; Sun, Y.-P.; Lin, M. The Optimal Concentration of Formaldehyde is Key to Stabilizing the Pre-Fusion Conformation of Respiratory Syncytial Virus Fusion Protein. Viruses 2019, 11, 628. [Google Scholar] [CrossRef] [Green Version]
  230. Gockel, C.M.; Bao, S.; Beagley, K.W. Transcutaneous immunization induces mucosal and systemic immunity: A potent method for targeting immunity to the female reproductive tract. Mol. Immunol. 2000, 37, 537–544. [Google Scholar] [CrossRef]
  231. Karande, P.; Mitragotri, S. Transcutaneous immunization: An overview of advantages, disease targets, vaccines, and delivery technologies. Annu. Rev. Chem. Biomol. Eng. 2010, 1, 175–201. [Google Scholar] [CrossRef] [PubMed]
  232. Hettinga, J.; Carlisle, R. Vaccination into the dermal compartment: Techniques, challenges, and prospects. Vaccines 2020, 8, 534. [Google Scholar] [CrossRef]
  233. Hirobe, S.; Okada, N.; Nakagawa, S. Transcutaneous vaccines–current and emerging strategies. Expert Opin. Drug Deliv. 2013, 10, 485–498. [Google Scholar] [CrossRef] [PubMed]
  234. Manikkath, J.; Hegde, A.R.; Parekh, H.S.; Mutalik, S. Peptide dendrimers in delivery of bioactive molecules to skin. In Nanoscience in Dermatology; Elsevier: Amsterdam, The Netherlands, 2016; pp. 89–97. [Google Scholar]
  235. Tezel, A.; Paliwal, S.; Shen, Z.; Mitragotri, S. Low-frequency ultrasound as a transcutaneous immunization adjuvant. Vaccine 2005, 23, 3800–3807. [Google Scholar] [CrossRef]
  236. Gehl, J. Electroporation: Theory and methods, perspectives for drug delivery, gene therapy and research. Acta Physiol. Scand. 2003, 177, 437–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Schoppink, J.; Rivas, D.F. Jet injectors: Perspectives for small volume delivery with lasers. Adv. Drug Deliv. Rev. 2022, 182, 114109. [Google Scholar] [CrossRef]
  238. Trimzi, M.A.; Ham, Y.-B. A needle-free jet injection system for controlled release and repeated biopharmaceutical delivery. Pharmaceutics 2021, 13, 1770. [Google Scholar] [CrossRef]
  239. Lin, C.-H.; Aljuffali, I.A.; Fang, J.-Y. Lasers as an approach for promoting drug delivery via skin. Expert Opin. Drug Deliv. 2014, 11, 599–614. [Google Scholar] [CrossRef]
  240. Kakar, P.; Li, Z.; Li, Y.; Cao, Y.; Chen, X. Laser facilitates week-long sustained transdermal drug delivery at high doses. J. Control. Release 2020, 319, 428–437. [Google Scholar] [CrossRef] [PubMed]
  241. Menon, I.; Bagwe, P.; Gomes, K.B.; Bajaj, L.; Gala, R.; Uddin, M.N.; D’souza, M.J.; Zughaier, S.M. Microneedles: A new generation vaccine delivery system. Micromachines 2021, 12, 435. [Google Scholar] [CrossRef] [PubMed]
  242. Alimardani, V.; Abolmaali, S.S.; Yousefi, G.; Rahiminezhad, Z.; Abedi, M.; Tamaddon, A.; Ahadian, S. Microneedle arrays combined with nanomedicine approaches for transdermal delivery of therapeutics. J. Clin. Med. 2021, 10, 181. [Google Scholar] [CrossRef]
  243. Wang, F.-Y.; Chen, Y.; Huang, Y.-Y.; Cheng, C.-M. Transdermal drug delivery systems for fighting common viral infectious diseases. Drug Deliv. Transl. Res. 2021, 11, 1498–1508. [Google Scholar] [CrossRef] [PubMed]
  244. Cárcamo-Martínez, Á.; Mallon, B.; Domínguez-Robles, J.; Vora, L.K.; Anjani, Q.K.; Donnelly, R.F. Hollow microneedles: A perspective in biomedical applications. Int. J. Pharm. 2021, 599, 120455. [Google Scholar] [CrossRef] [PubMed]
  245. Prausnitz, M.R. Engineering microneedle patches for vaccination and drug delivery to skin. Annu. Rev. Chem. Biomol. Eng. 2017, 8, 177–200. [Google Scholar] [CrossRef]
  246. Tariq, N.; Ashraf, M.W.; Tayyaba, S. A review on solid microneedles for biomedical applications. J. Pharm. Innov. 2022, 17, 1464–1483. [Google Scholar] [CrossRef]
  247. Jung, J.H.; Jin, S.G. Microneedle for transdermal drug delivery: Current trends and fabrication. J. Pharm. Investig. 2021, 51, 503–517. [Google Scholar] [CrossRef]
  248. Aldawood, F.K.; Andar, A.; Desai, S. A comprehensive review of microneedles: Types, materials, processes, characterizations and applications. Polymers 2021, 13, 2815. [Google Scholar] [CrossRef] [PubMed]
  249. He, X.; Sun, J.; Zhuang, J.; Xu, H.; Liu, Y.; Wu, D. Microneedle system for transdermal drug and vaccine delivery: Devices, safety, and prospects. Dose-Response 2019, 17, 1559325819878585. [Google Scholar] [CrossRef] [Green Version]
  250. Leone, M.; Mönkäre, J.; Bouwstra, J.A.; Kersten, G. Dissolving microneedle patches for dermal vaccination. Pharm. Res. 2017, 34, 2223–2240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Sartawi, Z.; Blackshields, C.; Faisal, W. Dissolving microneedles: Applications and growing therapeutic potential. J. Control. Release 2022, 348, 186–205. [Google Scholar] [CrossRef]
  252. Turner, J.G.; White, L.R.; Estrela, P.; Leese, H.S. Hydrogel-forming microneedles: Current advancements and future trends. Macromol. Biosci. 2021, 21, 2000307. [Google Scholar] [CrossRef]
  253. Dardano, P.; Caliò, A.; Di Palma, V.; Bevilacqua, M.F.; Di Matteo, A.; De Stefano, L. A photolithographic approach to polymeric microneedles array fabrication. Materials 2015, 8, 8661–8673. [Google Scholar] [CrossRef] [Green Version]
  254. Jin, Z.; Zhao, B.; Wan, W.; Li, Y.; Bai, X.; Hu, C.; Wang, J.; Li, Y.; Xin, W.; Kang, L. Application of microneedle-based vaccines in biosecurity. J. Biosaf. Biosecur. 2022, 4, 75–83. [Google Scholar] [CrossRef]
  255. Donnelly, R.F.; McCrudden, M.T.C.; Zaid Alkilani, A.; Larrañeta, E.; McAlister, E.; Courtenay, A.J.; Kearney, M.-C.; Singh, T.R.R.; McCarthy, H.O.; Kett, V.L. Hydrogel-forming microneedles prepared from “super swelling” polymers combined with lyophilised wafers for transdermal drug delivery. PLoS ONE 2014, 9, e111547. [Google Scholar] [CrossRef]
  256. Jiang, X.; Zhao, H.; Li, W. Microneedle-mediated transdermal delivery of drug-carrying nanoparticles. Front. Bioeng. Biotechnol. 2022, 10, 840395. [Google Scholar] [CrossRef] [PubMed]
  257. Habibi, N.; Mauser, A.; Ko, Y.; Lahann, J. Protein nanoparticles: Uniting the power of proteins with engineering design approaches. Adv. Sci. 2022, 9, 2104012. [Google Scholar] [CrossRef]
  258. Karande, P.; Mitragotri, S. Enhancement of transdermal drug delivery via synergistic action of chemicals. Biochim. Biophys. Acta-Biomembr. 2009, 1788, 2362–2373. [Google Scholar] [CrossRef] [Green Version]
  259. Barry, B.W. Lipid-protein-partitioning theory of skin penetration enhancement. J. Control. Release 1991, 15, 237–248. [Google Scholar] [CrossRef]
  260. Tan, G.; Xu, P.; Lawson, L.B.; He, J.; Freytag, L.C.; Clements, J.D.; John, V.T. Hydration effects on skin microstructure as probed by high-resolution cryo-scanning electron microscopy and mechanistic implications to enhanced transcutaneous delivery of biomacromolecules. J. Pharm. Sci. 2010, 99, 730–740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Alonso, A.; Meirelles, N.C.; Yushmanov, V.E.; Tabak, M. Water increases the fluidity of intercellular membranes of stratum corneum: Correlation with water permeability, elastic, and electrical resistance properties. J. Investig. Dermatol. 1996, 106, 1058–1063. [Google Scholar] [CrossRef] [Green Version]
  262. Notman, R.; den Otter, W.K.; Noro, M.G.; Briels, W.J.; Anwar, J. The permeability enhancing mechanism of DMSO in ceramide bilayers simulated by molecular dynamics. Biophys. J. 2007, 93, 2056–2068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Williams, A.C.; Barry, B.W. Terpenes and the lipid–protein–partitioning theory of skin penetration enhancement. Pharm. Res. 1991, 8, 17–24. [Google Scholar] [CrossRef]
  264. Osborne, D.W.; Musakhanian, J. Skin penetration and permeation properties of Transcutol®—Neat or diluted mixtures. AAPS PharmSciTech 2018, 19, 3512–3533. [Google Scholar] [CrossRef]
  265. Food and Drug Administration. VAXCHORA. Available online: https://www.fda.gov/vaccines-blood-biologics/vaccines/vaxchora (accessed on 24 April 2023).
  266. Food and Drug Administration. ROTARIX. Available online: https://www.fda.gov/vaccines-blood-biologics/vaccines/rotarix (accessed on 24 April 2023).
  267. Food and Drug Administration. RotaTeq. Available online: https://www.fda.gov/vaccines-blood-biologics/vaccines/rotateq (accessed on 23 April 2023).
  268. Food and Drug Administration. VIVOTIF. Available online: https://www.fda.gov/vaccines-blood-biologics/vaccines/vivotif (accessed on 23 April 2023).
  269. Food and Drug Administration. Adenovirus Type 4 and Type 7 Vaccine, Live, Oral. Available online: https://www.fda.gov/vaccines-blood-biologics/vaccines/adenovirus-type-4-and-type-7-vaccine-live-oral (accessed on 23 April 2023).
  270. Kool, M.; Fierens, K.; Lambrecht, B.N. Alum adjuvant: Some of the tricks of the oldest adjuvant. J. Med. Microbiol. 2012, 61, 927–934. [Google Scholar] [CrossRef]
  271. HogenEsch, H.; O’Hagan, D.T.; Fox, C.B. Optimizing the utilization of aluminum adjuvants in vaccines: You might just get what you want. NPJ Vaccines 2018, 3, 51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Ding, Z.; Van Riet, E.; Romeijn, S.; Kersten, G.F.A.; Jiskoot, W.; Bouwstra, J.A. Immune modulation by adjuvants combined with diphtheria toxoid administered topically in BALB/c mice after microneedle array pretreatment. Pharm. Res. 2009, 26, 1635–1643. [Google Scholar] [CrossRef] [Green Version]
  273. Stark, F.C.; Akache, B.; Deschatelets, L.; Tran, A.; Stuible, M.; Durocher, Y.; McCluskie, M.J.; Agbayani, G.; Dudani, R.; Harrison, B.A. Intranasal immunization with a proteosome-adjuvanted SARS-CoV-2 spike protein-based vaccine is immunogenic and efficacious in mice and hamsters. Sci. Rep. 2022, 12, 9772. [Google Scholar] [CrossRef]
  274. He, P.; Zou, Y.; Hu, Z. Advances in aluminum hydroxide-based adjuvant research and its mechanism. Hum. Vaccines Immunother. 2015, 11, 477–488. [Google Scholar] [CrossRef]
  275. Kotloff, K.L.; Sztein, M.B.; Wasserman, S.S.; Losonsky, G.A.; DiLorenzo, S.C.; Walker, R.I. Safety and immunogenicity of oral inactivated whole-cell Helicobacter pylori vaccine with adjuvant among volunteers with or without subclinical infection. Infect. Immun. 2001, 69, 3581–3590. [Google Scholar] [CrossRef] [Green Version]
  276. Harro, C.; Bourgeois, A.L.; Sack, D.; Walker, R.; DeNearing, B.; Brubaker, J.; Maier, N.; Fix, A.; Dally, L.; Chakraborty, S. Live attenuated enterotoxigenic Escherichia coli (ETEC) vaccine with dmLT adjuvant protects human volunteers against virulent experimental ETEC challenge. Vaccine 2019, 37, 1978–1986. [Google Scholar] [CrossRef] [PubMed]
  277. Qadri, F.; Akhtar, M.; Bhuiyan, T.R.; Chowdhury, M.I.; Ahmed, T.; Rafique, T.A.; Khan, A.; Rahman, S.I.A.; Khanam, F.; Lundgren, A. Safety and immunogenicity of the oral, inactivated, enterotoxigenic Escherichia coli vaccine ETVAX in Bangladeshi children and infants: A double-blind, randomised, placebo-controlled phase 1/2 trial. Lancet Infect. Dis. 2020, 20, 208–219. [Google Scholar] [CrossRef] [Green Version]
  278. Norton, E.B.; Lawson, L.B.; Freytag, L.C.; Clements, J.D. Characterization of a mutant Escherichia coli heat-labile toxin, LT (R192G/L211A), as a safe and effective oral adjuvant. Clin. Vaccine Immunol. 2011, 18, 546–551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Sjökvist Ottsjö, L.; Flach, C.-F.; Clements, J.; Holmgren, J.; Raghavan, S. A double mutant heat-labile toxin from Escherichia coli, LT (R192G/L211A), is an effective mucosal adjuvant for vaccination against Helicobacter pylori infection. Infect. Immun. 2013, 81, 1532–1540. [Google Scholar] [CrossRef] [Green Version]
  280. Sjökvist Ottsjö, L.; Jeverstam, F.; Yrlid, L.; Wenzel, A.U.; Walduck, A.K.; Raghavan, S. Induction of mucosal immune responses against Helicobacter pylori infection after sublingual and intragastric route of immunization. Immunology 2017, 150, 172–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  281. Maldarelli, G.A.; Matz, H.; Gao, S.; Chen, K.; Hamza, T.; Yfantis, H.G.; Feng, H.; Donnenberg, M.S. Pilin vaccination stimulates weak antibody responses and provides no protection in a C57Bl/6 murine model of acute Clostridium difficile infection. J. Vaccines Vaccin. 2016, 7, 321. [Google Scholar] [CrossRef] [Green Version]
  282. Hayden, C.A.; Streatfield, S.J.; Lamphear, B.J.; Fake, G.M.; Keener, T.K.; Walker, J.H.; Clements, J.D.; Turner, D.D.; Tizard, I.R.; Howard, J.A. Bioencapsulation of the hepatitis B surface antigen and its use as an effective oral immunogen. Vaccine 2012, 30, 2937–2942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Lalsiamthara, J.; Lee, J.H. Immunization with Salmonella Enteritidis secreting mucosal adjuvant labile toxin confers protection against wild type challenge via augmentation of CD3+ CD4+ T-cell proliferation and enhancement of IFN-γ, IL-6 and IL-10 expressions in chicken. Vaccine 2017, 35, 767–773. [Google Scholar] [CrossRef] [PubMed]
  284. Holmgren, J.; Nordqvist, S.; Blomquist, M.; Jeverstam, F.; Lebens, M.; Raghavan, S. Preclinical immunogenicity and protective efficacy of an oral Helicobacter pylori inactivated whole cell vaccine and multiple mutant cholera toxin: A novel and non-toxic mucosal adjuvant. Vaccine 2018, 36, 6223–6230. [Google Scholar] [CrossRef]
  285. Lebens, M.; Terrinoni, M.; Karlsson, S.L.; Larena, M.; Gustafsson-Hedberg, T.; Källgård, S.; Nygren, E.; Holmgren, J. Construction and preclinical evaluation of mmCT, a novel mutant cholera toxin adjuvant that can be efficiently produced in genetically manipulated Vibrio cholerae. Vaccine 2016, 34, 2121–2128. [Google Scholar] [CrossRef] [PubMed]
  286. Stratmann, T. Cholera toxin subunit B as adjuvant––An accelerator in protective immunity and a break in autoimmunity. Vaccines 2015, 3, 579–596. [Google Scholar] [CrossRef] [PubMed]
  287. Zhao, W.; Wu, W.; Xu, X. Oral vaccination with liposome-encapsulated recombinant fusion peptide of urease B epitope and cholera toxin B subunit affords prophylactic and therapeutic effects against H. pylori infection in BALB/c mice. Vaccine 2007, 25, 7664–7673. [Google Scholar] [CrossRef]
  288. Schussek, S.; Bernasconi, V.; Mattsson, J.; Wenzel, U.A.; Strömberg, A.; Gribonika, I.; Schön, K.; Lycke, N.Y. The CTA1-DD adjuvant strongly potentiates follicular dendritic cell function and germinal center formation, which results in improved neonatal immunization. Mucosal Immunol. 2020, 13, 545–557. [Google Scholar] [CrossRef]
  289. Guo, T.; Li, X.; Lin, M.; Zhang, L. Mucosal adjuvant activity of chitosan encapsulated nanoparticles as helicobacter pylori epitope vaccine carrier. Nanosci. Nanotechnol. Lett. 2016, 8, 1106–1111. [Google Scholar] [CrossRef]
  290. Laue, C.; Stevens, Y.; van Erp, M.; Papazova, E.; Soeth, E.; Pannenbeckers, A.; Stolte, E.; Böhm, R.; Gall, S.L.; Falourd, X. Adjuvant effect of orally applied preparations containing non-digestible polysaccharides on influenza vaccination in healthy seniors: A double-blind, randomised, controlled pilot trial. Nutrients 2021, 13, 2683. [Google Scholar] [CrossRef] [PubMed]
  291. Rachmawati, H.; Winarsih, S.; Prawiro, S.R.; Barlianto, W.; Sanarto, S.; Djunaedi, D.; Endharti, A.T.; Sardjono, T.W.; Khotimah, H.; Nugraheni, R.W. AdhO36 liposomes from Salmonella Typhi in combination with β-Glucan Immuno-adjuvant From Candida albicans cell wall as oral vaccine against typhoid fever in mice model. Open Access Maced. J. Med. Sci. 2020, 8, 441–448. [Google Scholar] [CrossRef]
  292. Kournikakis, B.; Mandeville, R.; Brousseau, P.; Ostroff, G. Anthrax-protective effects of yeast beta 1, 3 glucans. MedGenMed 2003, 5, 1. [Google Scholar] [PubMed]
  293. Courtney, A.N.; Nehete, P.N.; Nehete, B.P.; Thapa, P.; Zhou, D.; Sastry, K.J. Alpha-galactosylceramide is an effective mucosal adjuvant for repeated intranasal or oral delivery of HIV peptide antigens. Vaccine 2009, 27, 3335–3341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Davitt, C.J.H.; Longet, S.; Albutti, A.; Aversa, V.; Nordqvist, S.; Hackett, B.; McEntee, C.P.; Rosa, M.; Coulter, I.S.; Lebens, M. Alpha-galactosylceramide enhances mucosal immunity to oral whole-cell cholera vaccines. Mucosal Immunol. 2019, 12, 1055–1064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Longet, S.; Abautret-Daly, A.; Davitt, C.J.H.; McEntee, C.P.; Aversa, V.; Rosa, M.; Coulter, I.S.; Holmgren, J.; Raghavan, S.; Lavelle, E.C. An oral alpha-galactosylceramide adjuvanted Helicobacter pylori vaccine induces protective IL-1R-and IL-17R-dependent Th1 responses. NPJ Vaccines 2019, 4, 45. [Google Scholar] [CrossRef] [Green Version]
  296. Clements, J.D.; Norton, E.B. The mucosal vaccine adjuvant LT (R192G/L211A) or dmLT. mSphere 2018, 3, e00215–e00218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. Elieh Ali Komi, D.; Sharma, L.; Dela Cruz, C.S. Chitin and its effects on inflammatory and immune responses. Clin. Rev. Allergy Immunol. 2018, 54, 213–223. [Google Scholar] [CrossRef] [Green Version]
  298. Honda-Okubo, Y.; Saade, F.; Petrovsky, N. Advax™, a polysaccharide adjuvant derived from delta inulin, provides improved influenza vaccine protection through broad-based enhancement of adaptive immune responses. Vaccine 2012, 30, 5373–5381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  299. Wang, W.; Li, Y.; Wang, Y.; Ren, S.; Li, Y.; Wang, B. Polyactin A is a novel and potent immunological adjuvant for peptide-based cancer vaccine. Int. Immunopharmacol. 2018, 54, 95–102. [Google Scholar] [CrossRef] [PubMed]
  300. Zhang, S.; Li, W.; Smith, C.J.; Musa, H. Cereal-derived arabinoxylans as biological response modifiers: Extraction, molecular features, and immune-stimulating properties. Crit. Rev. Food Sci. Nutr. 2015, 55, 1035–1052. [Google Scholar] [CrossRef]
  301. Ichinohe, T.; Watanabe, I.; Ito, S.; Fujii, H.; Moriyama, M.; Tamura, S.-i.; Takahashi, H.; Sawa, H.; Chiba, J.; Kurata, T. Synthetic double-stranded RNA poly (I: C) combined with mucosal vaccine protects against influenza virus infection. J. Virol. 2005, 79, 2910–2919. [Google Scholar] [CrossRef] [Green Version]
  302. Vanpouille-Box, C.; Galluzzi, L. Nucleic Acid Sensing and Immunity-PART B; Academic Press: Cambridge, MA, USA, 2019. [Google Scholar]
  303. Ohto, U.; Shimizu, T. Structural aspects of nucleic acid-sensing Toll-like receptors. Biophys. Rev. 2016, 8, 33–43. [Google Scholar] [CrossRef] [Green Version]
  304. Medzhitov, R. Toll-like receptors and innate immunity. Nat. Rev. Immunol. 2001, 1, 135–145. [Google Scholar] [CrossRef] [PubMed]
  305. Morere, J.; Hognon, C.; Miclot, T.; Jiang, T.; Dumont, E.; Barone, G.; Monari, A.; Bignon, E. How Fragile We Are: Influence of Stimulator of Interferon Genes (STING) Variants on Pathogen Recognition and Immune Response Efficiency. J. Chem. Inf. Model. 2022, 62, 3096–3106. [Google Scholar] [CrossRef] [PubMed]
  306. McNab, F.; Mayer-Barber, K.; Sher, A.; Wack, A.; O’Garra, A. Type I interferons in infectious disease. Nat. Rev. 2015, 15, 87–103. [Google Scholar] [CrossRef] [PubMed]
  307. Tandrup Schmidt, S.; Foged, C.; Smith Korsholm, K.; Rades, T.; Christensen, D. Liposome-based adjuvants for subunit vaccines: Formulation strategies for subunit antigens and immunostimulators. Pharmaceutics 2016, 8, 7. [Google Scholar] [CrossRef]
  308. Hayashi, F.; Smith, K.D.; Ozinsky, A.; Hawn, T.R.; Yi, E.C.; Goodlett, D.R.; Eng, J.K.; Akira, S.; Underhill, D.M.; Aderem, A. The innate immune response to bacterial flagellin is mediated by Toll-like receptor 5. Nature 2001, 410, 1099–1103. [Google Scholar] [CrossRef] [PubMed]
  309. Lee, W.; Lee, S.-H.; Ahn, D.-G.; Cho, H.; Sung, M.-H.; Han, S.H.; Oh, J.-W. The antiviral activity of poly-γ-glutamic acid, a polypeptide secreted by Bacillus sp., through induction of CD14-dependent type I interferon responses. Biomaterials 2013, 34, 9700–9708. [Google Scholar] [CrossRef]
  310. Jones, T.; Cyr, S.; Allard, F.; Bellerose, N.; Lowell, G.H.; Burt, D.S. Protollin™: A novel adjuvant for intranasal vaccines. Vaccine 2004, 22, 3691–3697. [Google Scholar] [CrossRef]
  311. Díaz-Dinamarca, D.A.; Salazar, M.L.; Castillo, B.N.; Manubens, A.; Vasquez, A.E.; Salazar, F.; Becker, M.I. Protein-based adjuvants for vaccines as immunomodulators of the innate and adaptive immune response: Current knowledge, challenges, and future opportunities. Pharmaceutics 2022, 14, 1671. [Google Scholar] [CrossRef]
  312. Food and Drug Administration. Common Ingredients in U.S. Licensed Vaccines. Available online: https://www.fda.gov/vaccines-blood-biologics/safety-availability-biologics/common-ingredients-us-licensed-vaccines (accessed on 23 April 2023).
  313. Petrovsky, N. Comparative safety of vaccine adjuvants: A summary of current evidence and future needs. Drug Saf. 2015, 38, 1059–1074. [Google Scholar] [CrossRef] [PubMed]
  314. McLenon, J.; Rogers, M.A.M. The fear of needles: A systematic review and meta-analysis. J. Adv. Nurs. 2019, 75, 30–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Zheng, Z.; Diaz-Arévalo, D.; Guan, H.; Zeng, M. Noninvasive vaccination against infectious diseases. Hum. Vaccines Immunother. 2018, 14, 1717–1733. [Google Scholar] [CrossRef]
  316. Mutsch, M.; Zhou, W.; Rhodes, P.; Bopp, M.; Chen, R.T.; Linder, T.; Spyr, C.; Steffen, R. Use of the inactivated intranasal influenza vaccine and the risk of Bell’s palsy in Switzerland. N. Engl. J. Med. 2004, 350, 896–903. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  317. Stephenson, I.; Zambon, M.C.; Rudin, A.; Colegate, A.; Podda, A.; Bugarini, R.; Del Giudice, G.; Minutello, A.; Bonnington, S.; Holmgren, J. Phase I evaluation of intranasal trivalent inactivated influenza vaccine with nontoxigenic Escherichia coli enterotoxin and novel biovector as mucosal adjuvants, using adult volunteers. J. Virol. 2006, 80, 4962–4970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Valli, E.; Harriett, A.J.; Nowakowska, M.K.; Baudier, R.L.; Provosty, W.B.; McSween, Z.; Lawson, L.B.; Nakanishi, Y.; Norton, E.B. LTA1 is a safe, intranasal enterotoxin-based adjuvant that improves vaccine protection against influenza in young, old and B-cell-depleted (μMT) mice. Sci. Rep. 2019, 9, 15128. [Google Scholar] [CrossRef] [Green Version]
  319. Fan, X.T.; Wang, Y.L.; Su, Q.D.; Feng, Q.I.U.; Yao, Y.I.; Jia, Z.Y.; Da Yan, W.; Kun, Q.I.N.; Zou, Y.N.; Bi, S.L. Intranasal immunization using CTA1-DD as a mucosal adjuvant for an inactivated influenza vaccine. Biomed. Environ. Sci. 2019, 32, 531–540. [Google Scholar]
  320. Li, H.; Ren, H.; Zhang, Y.; Cao, L.; Xu, W. Intranasal vaccination with a recombinant protein CTA1-DD-RBF protects mice against hRSV infection. Sci. Rep. 2021, 11, 18641. [Google Scholar] [CrossRef]
  321. Andersen, C.S.; Dietrich, J.; Agger, E.M.; Lycke, N.Y.; Lövgren, K.; Andersen, P. The combined CTA1-DD/ISCOMs vector is an effective intranasal adjuvant for boosting prior Mycobacterium bovis BCG immunity to Mycobacterium tuberculosis. Infect. Immun. 2007, 75, 408–416. [Google Scholar] [CrossRef] [Green Version]
  322. De Filette, M.; Ramne, A.; Birkett, A.; Lycke, N.; Löwenadler, B.; Jou, W.M.; Saelens, X.; Fiers, W. The universal influenza vaccine M2e-HBc administered intranasally in combination with the adjuvant CTA1-DD provides complete protection. Vaccine 2006, 24, 544–551. [Google Scholar] [CrossRef] [PubMed]
  323. Asahi-Ozaki, Y.; Itamura, S.; Ichinohe, T.; Strong, P.; Tamura, S.-i.; Takahashi, H.; Sawa, H.; Moriyama, M.; Tashiro, M.; Sata, T. Intranasal administration of adjuvant-combined recombinant influenza virus HA vaccine protects mice from the lethal H5N1 virus infection. Microbes Infect. 2006, 8, 2706–2714. [Google Scholar] [CrossRef] [PubMed]
  324. Li, J.; Arévalo, M.T.; Chen, Y.; Posadas, O.; Smith, J.A.; Zeng, M. Intranasal immunization with influenza antigens conjugated with cholera toxin subunit B stimulates broad spectrum immunity against influenza viruses. Hum. Vaccines Immunother. 2014, 10, 1211–1220. [Google Scholar] [CrossRef]
  325. Li, M.; Jiang, Y.; Gong, T.; Zhang, Z.; Sun, X. Intranasal vaccination against HIV-1 with adenoviral vector-based nanocomplex using synthetic TLR-4 agonist peptide as adjuvant. Mol. Pharm. 2016, 13, 885–894. [Google Scholar] [CrossRef] [PubMed]
  326. Ren, S.-T.; Zhang, X.-M.; Sun, P.-F.; Sun, L.-J.; Guo, X.; Tian, T.; Zhang, J.; Guo, Q.-Y.; Li, X.; Guo, L.-J. Intranasal immunization using mannatide as a novel adjuvant for an inactivated influenza vaccine and its adjuvant effect compared with MF59. PLoS ONE 2017, 12, e0169501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  327. Luczo, J.M.; Bousse, T.; Johnson, S.K.; Jones, C.A.; Pearce, N.; Neiswanger, C.A.; Wang, M.-X.; Miller, E.A.; Petrovsky, N.; Wentworth, D.E. Intranasal powder live attenuated influenza vaccine is thermostable, immunogenic, and protective against homologous challenge in ferrets. NPJ Vaccines 2021, 6, 59. [Google Scholar] [CrossRef]
  328. Doavi, T.; Mousavi, S.L.; Kamali, M.; Amani, J.; Ramandi, M.F. Chitosan-based intranasal vaccine against Escherichia coli O157: H7. Iran. Biomed. J. 2016, 20, 97. [Google Scholar]
  329. Mann, A.J.; Noulin, N.; Catchpole, A.; Stittelaar, K.J.; De Waal, L.; Veldhuis Kroeze, E.J.B.; Hinchcliffe, M.; Smith, A.; Montomoli, E.; Piccirella, S. Intranasal H5N1 vaccines, adjuvanted with chitosan derivatives, protect ferrets against highly pathogenic influenza intranasal and intratracheal challenge. PLoS ONE 2014, 9, e93761. [Google Scholar] [CrossRef] [Green Version]
  330. Wu, Y.; Wei, W.; Zhou, M.; Wang, Y.; Wu, J.; Ma, G.; Su, Z. Thermal-sensitive hydrogel as adjuvant-free vaccine delivery system for H5N1 intranasal immunization. Biomaterials 2012, 33, 2351–2360. [Google Scholar] [CrossRef] [PubMed]
  331. Svindland, S.C.; Jul-Larsen, Å.; Pathirana, R.; Andersen, S.; Madhun, A.; Montomoli, E.; Jabbal-Gill, I.; Cox, R.J. The mucosal and systemic immune responses elicited by a chitosan-adjuvanted intranasal influenza H5N1 vaccine. Influenza Other Respir. Viruses 2012, 6, 90–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  332. Wang, Z.-B.; Shan, P.; Li, S.-Z.; Zhou, Y.; Deng, X.; Li, J.-L.; Zhang, Y.; Gao, J.-S.; Xu, J. The mechanism of action of acid-soluble chitosan as an adjuvant in the formulation of nasally administered vaccine against HBV. RSC Adv. 2016, 6, 96785–96797. [Google Scholar] [CrossRef]
  333. Hasegawa, H.; Ichinohe, T.; Strong, P.; Watanabe, I.; Ito, S.; Tamura, S.i.; Takahashi, H.; Sawa, H.; Chiba, J.; Kurata, T. Protection against influenza virus infection by intranasal administration of hemagglutinin vaccine with chitin microparticles as an adjuvant. J. Med. Virol. 2005, 75, 130–136. [Google Scholar] [CrossRef]
  334. Okamoto, S.; Matsuoka, S.; Takenaka, N.; Haredy, A.M.; Tanimoto, T.; Gomi, Y.; Ishikawa, T.; Akagi, T.; Akashi, M.; Okuno, Y. Intranasal immunization with a formalin-inactivated human influenza A virus whole-virion vaccine alone and intranasal immunization with a split-virion vaccine with mucosal adjuvants show similar levels of cross-protection. Clin. Vaccine Immunol. 2012, 19, 979–990. [Google Scholar] [CrossRef] [Green Version]
  335. Sasaki, E.; Kuramitsu, M.; Momose, H.; Kobiyama, K.; Aoshi, T.; Yamada, H.; Ishii, K.J.; Mizukami, T.; Hamaguchi, I. A novel vaccinological evaluation of intranasal vaccine and adjuvant safety for preclinical tests. Vaccine 2017, 35, 821–830. [Google Scholar] [CrossRef]
  336. Lin, Y.-L.; Chow, Y.-H.; Huang, L.-M.; Hsieh, S.-M.; Cheng, P.-Y.; Hu, K.-C.; Chiang, B.-L. A CpG-adjuvanted intranasal enterovirus 71 vaccine elicits mucosal and systemic immune responses and protects human SCARB2-transgenic mice against lethal challenge. Sci. Rep. 2018, 8, 10713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  337. Ho, H.-M.; Huang, C.-Y.; Yang, C.-H.; Liu, S.-J.; Chen, H.-W.; Yu, G.-Y.; Chen, J.-K.; Chuang, T.-H.; Huang, M.-H. Formulation of SARS-CoV-2 spike protein with CpG oligodeoxynucleotides and squalene nanoparticles modulates immunological aspects following intranasal delivery. Pharmaceutics 2022, 14, 2539. [Google Scholar] [CrossRef] [PubMed]
  338. Naito, Y.; Hamaoka, S.; Kinoshita, M.; Kainuma, A.; Shimizu, M.; Katoh, H.; Moriyama, K.; Ishii, K.J.; Sawa, T. The protective effects of nasal PcrV-CpG oligonucleotide vaccination against Pseudomonas aeruginosa pneumonia. Microbiol. Immunol. 2018, 62, 774–785. [Google Scholar] [CrossRef] [PubMed]
  339. Poggio, T.V.; La Torre, J.L.; Scodeller, E.A. Intranasal immunization with a recombinant truncated FimH adhesin adjuvanted with CpG oligodeoxynucleotides protects mice against uropathogenic Escherichia coli challenge. Can. J. Microbiol. 2006, 52, 1093–1102. [Google Scholar] [CrossRef]
  340. Jearanaiwitayakul, T.; Limthongkul, J.; Kaofai, C.; Apichirapokey, S.; Chawengkirttikul, R.; Sapsutthipas, S.; Sunintaboon, P.; Ubol, S. The STING ligand and delivery system synergistically enhance the immunogenicity of an intranasal spike SARS-CoV-2 vaccine candidate. Biomedicines 2022, 10, 1142. [Google Scholar] [CrossRef]
  341. Du, Y.; Xu, Y.; Feng, J.; Hu, L.; Zhang, Y.; Zhang, B.; Guo, W.; Mai, R.; Chen, L.; Fang, J. Intranasal administration of a recombinant RBD vaccine induced protective immunity against SARS-CoV-2 in mouse. Vaccine 2021, 39, 2280–2287. [Google Scholar] [CrossRef]
  342. Falkeborn, T.; Asahara, N.; Hayashi, M.; Arai, M.; Hinkula, J.; Maltais, A.-K. Comparison of the mucosal adjuvant Endocine™ with two well-known adjuvants: Cholera toxin and alum. Jacobs J. Vaccine Vaccin. 2015, 1, 6. [Google Scholar]
  343. Yang, J.; Liu, M.-Q.; Liu, L.; Li, X.; Xu, M.; Lin, H.; Liu, S.; Hu, Y.; Li, B.; Liu, B. A triple-RBD-based mucosal vaccine provides broad protection against SARS-CoV-2 variants of concern. Cell. Mol. Immunol. 2022, 19, 1279–1289. [Google Scholar] [CrossRef]
  344. Hong, S.H.; Byun, Y.-H.; Nguyen, C.T.; Kim, S.Y.; Seong, B.L.; Park, S.; Woo, G.-J.; Yoon, Y.; Koh, J.T.; Fujihashi, K. Intranasal administration of a flagellin-adjuvanted inactivated influenza vaccine enhances mucosal immune responses to protect mice against lethal infection. Vaccine 2012, 30, 466–474. [Google Scholar] [CrossRef]
  345. Treanor, J.; Nolan, C.; O’Brien, D.; Burt, D.; Lowell, G.; Linden, J.; Fries, L. Intranasal administration of a proteosome-influenza vaccine is well-tolerated and induces serum and nasal secretion influenza antibodies in healthy human subjects. Vaccine 2006, 24, 254–262. [Google Scholar] [CrossRef]
  346. Novotny, L.A.; Clements, J.D.; Bakaletz, L.O. Transcutaneous immunization as preventative and therapeutic regimens to protect against experimental otitis media due to nontypeable Haemophilus influenzae. Mucosal Immunol. 2011, 4, 456–467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  347. Riddle, M.S.; Maciel, M., Jr.; Porter, C.K.; Poole, S.T.; Gutierrez, R.L.; Gormley, R.; Laird, R.M.; Sebeny, P.J.; Dori, K.E.; Greenleaf, M.E. A first in human clinical trial assessing the safety and immunogenicity of transcutaneously delivered enterotoxigenic Escherichia coli fimbrial tip adhesin with heat-labile enterotoxin with mutation R192G. Vaccine 2020, 38, 7040–7048. [Google Scholar] [CrossRef]
  348. Stickings, P.; Peyre, M.; Coombes, L.; Muller, S.; Rappuoli, R.; Del Giudice, G.; Partidos, C.D.; Sesardic, D. Transcutaneous immunization with cross-reacting material CRM197 of diphtheria toxin boosts functional antibody levels in mice primed parenterally with adsorbed diphtheria toxoid vaccine. Infect. Immun. 2008, 76, 1766–1773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  349. Tierney, R.; Beignon, A.-S.; Rappuoli, R.; Muller, S.; Sesardic, D.; Partidos, C.D. Transcutaneous immunization with tetanus toxoid and mutants of Escherichia coli heat-labile enterotoxin as adjuvants elicits strong protective antibody responses. J. Infect. Dis. 2003, 188, 753–758. [Google Scholar] [CrossRef] [Green Version]
  350. Skountzou, I.; Quan, F.-S.; Jacob, J.; Compans, R.W.; Kang, S.-M. Transcutaneous immunization with inactivated influenza virus induces protective immune responses. Vaccine 2006, 24, 6110–6119. [Google Scholar] [CrossRef]
  351. Guo, L.; Qiu, Y.; Chen, J.; Zhang, S.; Xu, B.; Gao, Y. Effective transcutaneous immunization against hepatitis B virus by a combined approach of hydrogel patch formulation and microneedle arrays. Biomed. Microdevices 2013, 15, 1077–1085. [Google Scholar] [CrossRef] [PubMed]
  352. Bal, S.M.; Ding, Z.; Kersten, G.F.A.; Jiskoot, W.; Bouwstra, J.A. Microneedle-based transcutaneous immunisation in mice with N-trimethyl chitosan adjuvanted diphtheria toxoid formulations. Pharm. Res. 2010, 27, 1837–1847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  353. Weldon, W.C.; Zarnitsyn, V.G.; Esser, E.S.; Taherbhai, M.T.; Koutsonanos, D.G.; Vassilieva, E.V.; Skountzou, I.; Prausnitz, M.R.; Compans, R.W. Effect of adjuvants on responses to skin immunization by microneedles coated with influenza subunit vaccine. PLoS ONE 2012, 7, e41501. [Google Scholar] [CrossRef]
  354. Liu, C.; Hashizume, T.; Kurita-Ochiai, T.; Yamamoto, M. CpG oligodeoxynucleotide is an effective adjuvant for transcutaneous immunization. Int. J. Oral-Med. Sci. 2007, 6, 91–96. [Google Scholar] [CrossRef]
  355. Berry, L.J.; Hickey, D.K.; Skelding, K.A.; Bao, S.; Rendina, A.M.; Hansbro, P.M.; Gockel, C.M.; Beagley, K.W. Transcutaneous immunization with combined cholera toxin and CpG adjuvant protects against Chlamydia muridarum genital tract infection. Infect. Immun. 2004, 72, 1019–1028. [Google Scholar] [CrossRef] [Green Version]
  356. Belyakov, I.M.; Hammond, S.A.; Ahlers, J.D.; Glenn, G.M.; Berzofsky, J.A. Transcutaneous immunization induces mucosal CTLs and protective immunity by migration of primed skin dendritic cells. J. Clin. Investig. 2004, 113, 998–1007. [Google Scholar] [CrossRef] [PubMed]
  357. Beignon, A.-S.; Brown, F.; Eftekhari, P.; Kramer, E.; Briand, J.-P.; Muller, S.; Partidos, C.D. A peptide vaccine administered transcutaneously together with cholera toxin elicits potent neutralising anti-FMDV antibody responses. Vet. Immunol. Immunopathol. 2005, 104, 273–280. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of limitations of invasive vaccines.
Figure 1. Schematic illustration of limitations of invasive vaccines.
Pharmaceutics 15 02114 g001
Figure 2. Schematic illustration of immune responses upon vaccine administration.
Figure 2. Schematic illustration of immune responses upon vaccine administration.
Pharmaceutics 15 02114 g002
Figure 3. Schematic illustration of DNA vaccine production methods. Production of circular DNA (A) and linear DNA (B).
Figure 3. Schematic illustration of DNA vaccine production methods. Production of circular DNA (A) and linear DNA (B).
Pharmaceutics 15 02114 g003
Figure 4. Representation of immune responses of an intestinal M cell.
Figure 4. Representation of immune responses of an intestinal M cell.
Pharmaceutics 15 02114 g004
Figure 5. Schematic illustration of mucus-penetrating mechanisms. (A) Size-dependent penetration, (B) non-reactive coating treatment, and (C) permeation enhancer treatment.
Figure 5. Schematic illustration of mucus-penetrating mechanisms. (A) Size-dependent penetration, (B) non-reactive coating treatment, and (C) permeation enhancer treatment.
Pharmaceutics 15 02114 g005
Table 1. Nano- and micro-sized carriers with adjuvanticity.
Table 1. Nano- and micro-sized carriers with adjuvanticity.
Carriers/AdjuvantsTypeImmune ResponsesLimitationsStatus
Lipid-based carriersImmune-stimulating complexes (ISCOM)
  • Activation of CD8 T cells [105]
  • Th1 and Th2 activation [106]
  • High dose of saponin is required to manufacture (immunoreactive) [106]
  • Limitation in encapsulation of non-membrane-derived antigens [106]
In clinical stage [107]
Liposome
  • Th1 and Th2 activation [108]
  • Limitations in the selection of lipid due to rigidity [109]
In use in the vaccine industry; Pfizer/BioNTech and Moderna (SARS-CoV-2) [110]
Bilosome
  • Th1 activation [111]
  • High stability but low immunogenicity [112]
In preclinical stage [113]
Archaeosomes
  • Activation of CD8 T cells [114]
  • Th1 and Th2 activation [115]
  • Mechanism-of-action- and safety-related research should be performed [115]
In preclinical stage [116]
Virosomes
  • Activation of CD8 T cells [114,117]
  • Th1 and Th2 activation [115,117]
  • Rapid disintegration in the blood [118]
In use in the vaccine industry; Epaxal® and Inflexal®V (influenza vaccine) [119,120]
Protein-based carriersGelatin
  • Activation of CD8 T cells [121]
  • Th1 and Th2 activation [122]
  • Large particle size
  • Rapid degradation [123]
In clinical stage [124]
Albumin
  • Activation of CD8 T cells [125]
  • Th1 activation [126]
  • Potential immunogenic reaction [127]
FDA-approved [123]
Zein
  • Th1 and Th2 activation [128]
  • Low stability [129]
In preclinical stage [128]
EmulsionMF59
(oil-in-water emulsion)
  • Activation of CD8 T cells [130]
  • Th2 activation [130]
  • Limitation in drug release control
In use in the vaccine industry; Fluad®, Focetria®, and Celtura® (influenza vaccine) [130]
AS03
(oil-in-water emulsion)
  • Th1 and Th2 activation [131]
  • Limitation in drug release control
In use in the vaccine industry; Pandemrix and Arepanrix (influenza vaccin) [132]
AF03
(oil-in-water emulsion)
  • Th1 and Th2 activation [133]
  • Lower immune response compared to AS03 [133]
In use in the vaccine industry; Humenza™ (influenza vaccine) [134]
PolymerPoly(lactic-co-glycolic acid)
  • Activation of CD8 T cells [135]
  • Th1 and Th2 activation (Th2 dominant) [136]
  • Slow and non-tunable degradation rate [137]
  • Degradation into acidic byproduct (inflammation) [138]
FDA-approved
Poly(lactic acid)
  • Activation of CD8 T cells
  • Th1 and Th2 activation [139]
  • Degradation into acidic byproduct (inflammation) [140]
FDA-approved
Chitosan
  • Th1 and Th2 activation [141]
  • Reduction in CD8 T cell response [142]
FDA-approved
Dextran (acetalated)
  • Activation of CD8 T cells [137]
  • Th1 and Th2 activation [137]
  • Fast and tunable degradation rate [137]
  • Degradation into pH-neutral byproduct [137,138]
FDA-approved
InorganicCalcium phosphate nanoparticles
  • Activation of CD8 T cells [102]
  • Th1 and Th2 activation [143]
  • Low antigen loading capacity and rapid aggregation [102]
FDA-approved
Silica nanoparticles
  • Activation of CD8 T cells [144]
  • Th1 and Th2 activation [144]
  • Hemolysis through interaction of silanol groups and the phospholipids of the red blood cells [145]
In clinical stage [146]
Gold nanoparticles
  • Activation of CD8 T cells [147]
  • Th1 and Th2 activation [148]
  • Poor biodegradability [149]
FDA-approved
Table 2. Adjuvants for oral vaccines.
Table 2. Adjuvants for oral vaccines.
AdjuvantTypeVaccine NameVaccine TypeTarget DiseaseStatus
Toxin-based adjuvantHeat-labile enterotoxin (LT) -
  • Inactivated vaccine [275]
Helicobacter pyloriIn clinical
Double-mutant heat-labile toxin (dmLT)ETEC vaccine (ACE527) [276]
  • Live attenuated vaccine
Enterotoxigenic Escherichia coli (ETEC)In clinical
ETVAX [277]
  • Inactivated vaccine
Enterotoxigenic Escherichia coli (ETEC)In clinical
-
  • Inactivated vaccine [278]
  • Subunit vaccine [279,280]
Helicobacter pyloriIn preclinical
-
  • Subunit vaccine [281]
Clostridium difficileIn preclinical
-
  • Subunit vaccine [278]
Clostridium tetaniIn preclinical
-
  • Subunit vaccine [282]
Hepatitis B virusIn preclinical
-
  • Live attenuated vaccine [283]
Salmonella enteritidisIn preclinical
Multiple-mutated cholera toxin (mmCT)-
  • Inactivated vaccine [284]
Helicobacter pyloriIn preclinical
-
  • Inactivated bacteria (Vibrio cholerae) [285]
  • Subunit vaccine (influenza virus) [285]
Vibrio cholerae and influenza virusIn preclinical
Recombinant cholera toxin B subunit (rCTB)Dukoral [286]
  • Inactivated vaccine
Vibrio choleraePrequalified by WHO
-
  • Live bacteria [287]
Helicobacter pyloriIn preclinical
Cholera-toxin-derived adjuvant (CTA1DD) CTA1-3M2e-DD [288]
  • Subunit vaccine
Influenza virusIn preclinical
Polysaccharide-based adjuvantChitosan-
  • Inactivated vaccine [289]
Helicobacter pyloriIn preclinical
β-glucans-
  • Inactivated vaccine [290]
Influenza virusIn clinical
-
  • Subunit vaccine [291]
Salmonella TyphiIn preclinical
-
  • Inactivated vaccine [292]
Bacillus anthracisIn preclinical
Arabinoxylan (AX) -
  • Inactivated vaccine [290]
Influenza virusIn clinical
Bacterial exopolysaccharide (EPS) -
  • Inactivated vaccine [290]
Influenza virusIn clinical
Lipid-based adjuvantα-galactosyl ceramide-
  • Subunit vaccine [293]
Human immunodeficiency virus In preclinical
-
  • Inactivated vaccine [294]
Vibrio choleraeIn preclinical
-
  • Inactivated vaccine [295]
Helicobacter pyloriIn preclinical
Table 3. Adjuvants for intranasal vaccines.
Table 3. Adjuvants for intranasal vaccines.
AdjuvantTypeVaccine NameVaccine TypeTarget DiseaseStatus
Toxin-based adjuvantHeat-labile enterotoxin (LT)Nasalflu (Berna Biotech) [316]Inactivated vaccineInfluenza virusIn clinical
-Inactivated vaccine [317]Influenza virusIn clinical
Enzymatic A1 domain of LT (LTA1) -Subunit vaccine [318]Influenza virusIn preclinical
Cholera-toxin-derived adjuvant (CTA1DD) -Inactivated vaccine [319]Influenza virusIn clinical
-Inactivated vaccine [320]Human respiratory syncytial virus (hRSV)In preclinical
-Subunit vaccine [321]Mycobacterium tuberculosisIn preclinical
-Subunit vaccine [322,323]Influenza virusIn preclinical
Cholera toxin B subunit (CTB)-Live attenuated vaccine [323]Influenza virusIn preclinical
-Subunit vaccine [324]Influenza virusIn preclinical
Polysaccharide-based adjuvantRS09 (Ala-Pro-Pro-His-Ala-Leu-Ser) -Viral vector vaccine [325]Human immunodeficiency virusIn preclinical
Mannatide (polyactin A)-Inactivated vaccine [326]Influenza virusIn preclinical
Advax™-Live attenuated [327]Influenza virusIn preclinical
Chitosan-Non-viral vector vaccine [328]Enterohemorrhagic Escherichia coli (EHEC) O157:H7In preclinical
-Inactivated vaccine [329]Influenza virusIn preclinical
-Subunit vaccine [330,331]Influenza virusIn preclinical
-Subunit vaccine [332]Hepatitis B virusIn preclinical
Chitin-Live attenuated [323]Influenza virusIn preclinical
-Subunit vaccine [333]Influenza virusIn preclinical
Nucleotide-based adjuvant, synthetic adjuvantSynthetic double-stranded RNA polyriboinosinic acid-polyribocytidylic acid [poly (I:C)]-Live attenuated vaccine [323]Influenza virusIn preclinical
-Inactivated vaccine [334]Influenza virusIn preclinical
Cytosine-phosphate-guanine oligodeoxynucleotide (CpG ODN)-Subunit vaccine [335]Influenza virusIn preclinical
-Inactivated vaccine [336]EnterovirusIn preclinical
-Subunit vaccine [337]SARS-CoV-2In preclinical
-Non-viral vector vaccine [338]Pseudomonas aeruginosaIn preclinical
-Non-viral vector vaccine [339]Escherichia coliIn preclinical
Cyclic guanosine monophosphate-adenosine monophosphate (2′3′-cGAMP or cGAMP)-Subunit vaccine [340]SARS-CoV-2In preclinical
Aluminum-based adjuvantAlhydrogel® (aluminum oxyhydroxide gel)-Subunit vaccine [341]SARS-CoV-2In preclinical
Lipid-based adjuvantEndocine™ (lipids monoolein and oleic acid)-Subunit vaccine [342]Influenza virusIn preclinical
Protein-based adjuvantFlagellin-Inactivated vaccine [343]SARS-CoV-2In preclinical
-Inactivated vaccine [344]Influenza virusIn preclinical
Poly(γ-glutamic acid)-Inactivated vaccine [334]Influenza virusIn preclinical
Proteosome-Subunit vaccine [273]SARS-CoV-2In preclinical
-Inactivated vaccine [345]Influenza virusIn preclinical
ProtollinTM-Subunit vaccine [310]Influenza virusIn preclinical
Table 4. Adjuvants for transcutaneous vaccines.
Table 4. Adjuvants for transcutaneous vaccines.
AdjuvantTypeVaccine NameVaccine TypeTarget DiseaseAdministration MethodStatus
Toxin derivatesDouble-mutant heat-labile toxin (dmLT)-Subunit vaccine [346]Nontypeable Haemophilus influenzae (NTHI)Hydration (sterile, pyrogen-free 0.9% sodium chloride)In preclinical
Single-mutant heat-labile enterotoxin-Subunit vaccine [347]Enterotoxigenic Escherichia coli (ETEC)Hydration and abrasion (70% isopropyl alcohol, glycerol 10%, and 10 swipes of medical grade sandpaper)In clinical
-Subunit vaccine [348]Corynebacterium diphtheriaeHydration (sterile PBS)In preclinical
-Subunit vaccine [349]Clostridium tetaniHydrationIn preclinical
Cholera toxin (CT)-Subunit vaccine [272]Corynebacterium diphtheriaeHollow microneedle (pre-treatment purpose)In preclinical
-Inactivated vaccine [350]Influenza virusHydration (saline-soaked gauze)In preclinical
Cholera toxin B subunit (CTB)-Subunit vaccine [351]hepatitis B
virus
Solid microneedle and hydrogel patchIn preclinical
Polysaccharide-based adjuvantChitosan-Subunit vaccine [352]Corynebacterium diphtheriaeHollow microneedle (pre- or post-treatment purpose)In preclinical
Nucleotide-based adjuvant, synthetic adjuvantSynthetic double-stranded RNA polyriboinosinic acid-polyribocytidylic acid [poly (I:C)]-Subunit vaccine [353]Influenza virusCoated microneedleIn preclinical
Cytosine-phosphate-guanine oligodeoxynucleotide (CpG ODN)-Subunit vaccine [354]Clostridium tetaniHydration (PBS-drenched
wound plaster)
In preclinical
-Subunit vaccine [272]Corynebacterium diphtheriaeHollow microneedle (pre-treatment purpose)In preclinical
-Non-viral vector vaccine [355]Chlamydia muridarumHydration (sterile PBS)In preclinical
-Subunit vaccine [356]Human immunodeficiency virusHydration (saline-drenched gauze)In preclinical
-Subunit vaccine [357]Foot-and-mouth disease
virus
HydrationIn preclinical
Immune-stimulating complexes (ISCOMs)Quil A-Subunit vaccine [272]Corynebacterium diphtheriaeHollow microneedle (pre-treatment purpose)In preclinical
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

Han, S.; Lee, P.; Choi, H.-J. Non-Invasive Vaccines: Challenges in Formulation and Vaccine Adjuvants. Pharmaceutics 2023, 15, 2114. https://doi.org/10.3390/pharmaceutics15082114

AMA Style

Han S, Lee P, Choi H-J. Non-Invasive Vaccines: Challenges in Formulation and Vaccine Adjuvants. Pharmaceutics. 2023; 15(8):2114. https://doi.org/10.3390/pharmaceutics15082114

Chicago/Turabian Style

Han, Sumin, Panjae Lee, and Hyo-Jick Choi. 2023. "Non-Invasive Vaccines: Challenges in Formulation and Vaccine Adjuvants" Pharmaceutics 15, no. 8: 2114. https://doi.org/10.3390/pharmaceutics15082114

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