Next Article in Journal
Influenza Vaccination in Children Younger than 5 Years in the Region of Murcia (Spain), a Comparative Analysis among Vaccinating and Non-Vaccinating Parents: Data from the FLUTETRA Study
Next Article in Special Issue
Trends in COVID-19 Vaccine Development: Vaccine Platform, Developer, and Nationality
Previous Article in Journal
Intradermal Inoculation of Inactivated Foot-and-Mouth Disease Vaccine Induced Effective Immune Responses Comparable to Conventional Intramuscular Injection in Pigs
Previous Article in Special Issue
A Perspective on Oral Immunotherapeutic Tools and Strategies for Autoimmune Disorders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Vaccine Strategies to Elicit Mucosal Immunity

by
Yufeng Song
1,
Frances Mehl
1 and
Steven L. Zeichner
1,2,*
1
Department of Pediatrics, University of Virginia, Charlottesville, VA 22908, USA
2
Department of Microbiology, Immunology, and Cancer Biology, University of Virginia, Charlottesville, VA 22908, USA
*
Author to whom correspondence should be addressed.
Vaccines 2024, 12(2), 191; https://doi.org/10.3390/vaccines12020191
Submission received: 1 December 2023 / Revised: 29 January 2024 / Accepted: 8 February 2024 / Published: 13 February 2024
(This article belongs to the Special Issue New Trends in Vaccine Characterization, Formulations, and Development)

Abstract

:
Vaccines are essential tools to prevent infection and control transmission of infectious diseases that threaten public health. Most infectious agents enter their hosts across mucosal surfaces, which make up key first lines of host defense against pathogens. Mucosal immune responses play critical roles in host immune defense to provide durable and better recall responses. Substantial attention has been focused on developing effective mucosal vaccines to elicit robust localized and systemic immune responses by administration via mucosal routes. Mucosal vaccines that elicit effective immune responses yield protection superior to parenterally delivered vaccines. Beyond their valuable immunogenicity, mucosal vaccines can be less expensive and easier to administer without a need for injection materials and more highly trained personnel. However, developing effective mucosal vaccines faces many challenges, and much effort has been directed at their development. In this article, we review the history of mucosal vaccine development and present an overview of mucosal compartment biology and the roles that mucosal immunity plays in defending against infection, knowledge that has helped inform mucosal vaccine development. We explore new progress in mucosal vaccine design and optimization and novel approaches created to improve the efficacy and safety of mucosal vaccines.

1. History of Vaccine Development

Vaccines, one of the most important medical inventions in human history, are essential tools for preventing disease. They have saved countless lives and continue to be a key component of public health efforts, especially in lower-income countries [1].
Historical practices, like variolation, which had their origins in ancient history in the civilizations of East and South Asia, constitute the earliest examples of efforts by humans to prevent infectious disease by prior exposure to infectious agents [2,3,4]. Interestingly, some of these early variolation techniques involved exposing individuals to infectious agents via a mucosal route, such as nasal insufflation of infectious material, pulverized smallpox lesion scabs [3,5]. The public health and military importance of immunization in historical settings is illustrated by the well-known history of General George Washington mandating variolation for troops under his command during the American Revolutionary War [6].
Vaccination by a modified or homolog pathogen began in the late 18th century with Edward Jenner, who found that deliberate infection with cowpox could protect against the more fatal smallpox [7,8,9]. In the 1880s, Louis Pasteur expanded the target of vaccines from smallpox to other pathogens when he discovered that pathogens could be attenuated through repeated passage in environments in which the full complement of pathogenic features was not required for pathogen survival or inactivated by chemical exposure. In 1886, the first inactivated whole-organism vaccines were created against typhoid and cholera [10,11]. The 20th century saw the invention of vaccines that used only parts of the pathogen, including toxoids (1923), pathogen subunits (1944), capsular polysaccharides (1977), protein-conjugated capsular polysaccharides (1987), and recombinant proteins (1986). The use of cell culture to attenuate viruses also became a much-used tool in the 1950s and 1960s [11]. Viral vector-based and mRNA vaccines are some of the more recent vaccine developments of the 21st century [12].
The first vaccines were delivered parenterally using a lancet [9]. Even now, the most common mechanisms of delivery remain the parenteral routes: intramuscular, subcutaneous, or intradermal [13]. Vaccines are evaluated for efficacy through clinical studies, with an important corollary being their ability to provoke a systemic immune response, measured by assessing levels of antibodies against the pathogen [11,14]. While parenterally administered vaccines have made great contributions to human health, and some mucosally-administered vaccines have proved to be safe and effective, such as the live attenuated polio [15,16], rotavirus [17], and influenza [18,19] vaccines and the killed whole-cell cholera vaccine [20,21], most vaccines are not mucosally administered. The advantages of vaccines administered via a mucosal route have led to increasing interest in the development of new mucosal vaccines [13,14,22,23,24]. Routes that have been used for mucosal vaccination include mainly intranasal administration and oral administration, which can deliver vaccine antigens to the GI tract beyond the stomach. Intraoral administration of vaccines, such as sublingual or buccal exposure, has also been used [25,26,27], although these routes have not been as well developed. Rectal vaccine administration has also been studied, and vaccination across urogenital mucosa is also possible, but such routes are generally less appealing due to practical administration and patient acceptability concerns [28].
Parenteral vaccines made of virus-like particles have induced the production of IgA in addition to IgG [29,30,31]. DNA parenteral vaccines have also induced mucosal immunity in animal models [32,33,34]. Some, but not all, adjuvants in parenteral vaccines have promoted mucosal immunity; the mechanism behind this is not yet understood [35]. Nevertheless, most workers in the field consider that an optimal mucosal immune response will result following mucosal exposure to the immunogen [36,37,38,39,40].

2. Importance of Mucosal Vaccines and Their Development

Mucosal immunity is of particular importance because mucosal surfaces are the point of entry for more than 90% of human pathogens [41]. The mucosal surfaces represent the first line of the host against infection. Strong mucosal immunity can prevent infection by pathogens across mucosal surfaces and reduce secondary transmission of pathogens shed from a mucosal surface [41,42,43,44]. Even in the event of infection, mucosal immunity at sites of pathogen shedding could prevent forward transmission, lowering infection rates, thereby slowing the emergence of new mutants, reducing morbidity, and blunting potential future pandemics [43,45]. Most vaccines produce a systemic immune response, and while some animal studies have suggested that injected vaccines produce good mucosal immune responses, many parenterally administered vaccines that produce excellent humoral immune responses yield substantially worse mucosal immune responses [35,42]. Several studies have shown administering vaccines at a mucosal surface elicits the most effective mucosal immune responses [36,37,38,39,40].
A mucosal vaccine is a type of vaccine designed to stimulate effective immune responses at the sites of infection, which can help prevent the spread of diseases and is administered through the mucosal surfaces of the host body. Mucosal vaccines have several advantages over traditional vaccines. They can provide more effective protection against diseases by blocking transmission through mucosal surfaces at the site of infection [43,46]. Additionally, most mucosal vaccines can be easier to administer than traditional vaccines, as they do not require needles, syringes, and material to disinfect the injection site [47], which decreases costs and logistical problems. Non-parenteral vaccines can be administered by personnel with less training than is required for parenteral administration [48]. Non-parenteral vaccines may also be more readily accepted by vaccinees who are averse to receiving injections [49,50].
Where mucosal immunity is induced depends on the route of administration. Mucosal vaccines administered across one mucosal surface elicit mucosal immunity where they are administered and in certain other mucosal compartments depending on the administration route. Both nasal and sublingual vaccines can elicit mucosal immunity in the upper and lower respiratory tract and the reproductive tract. Sublingual vaccines also elicit mucosal immunity in the stomach and small intestine. Oral, rectal, and vaginal vaccines induce mucosal immunity in the GI tract, colon and rectum, and reproductive tract [51,52].
The first modern, licensed mucosal vaccine was produced by Albert Sabin, who created an oral polio vaccine with three attenuated strains [10,53]. This oral vaccine induced systemic immunity but, importantly, also local immunity in the GI tract. Replication of the vaccine poliovirus strains also led to secondary exposure to the vaccine strains in non-vaccinated persons, strengthening herd immunity. Mucosal immunity elicited by oral polio vaccines addressed the problem of polio being able to replicate in the gut of people vaccinated with the injectable inactivated vaccine and then infect non-vaccinated individuals [53].
Most of the approved mucosal vaccines employ attenuated pathogens administered orally (Table 1). There have been approximately 30 clinical trials for nasal vaccines, but FluMist® (a live, attenuated influenza vaccine) remains the only one approved in the United States [54]. FluMist® is composed of influenza strains that are temperature-sensitive and cold-adapted, becoming attenuated at body temperature [55,56]. Vaccination with intranasal live, attenuated influenza vaccines yields higher serum IgA and T-cell immunity and decreased viral replication compared to intramuscular inactivated vaccines [57,58]. However, they also have decreased hemagglutination inhibition, microneutralization, and IgGs [57,59].
Both parenteral and mucosal vaccines may promote cellular immunity; however, only mucosal vaccines promote mucosal tissue-resident memory T-cells and B-cells. Therefore, mucosal vaccines enable a more direct and immediate response to an invading pathogen [59,60,61]. This effect for viral vaccines depends on the pathogen being live or in another viral vector; an inactivated virus administered intranasally does not induce cellular immunity [59,60], although it may be possible to develop new strategies to enhance the production of cell-mediated immunity by non-living vaccines with adjuvants, vehicles, or novel dosing schedules. One study has shown that inactivated influenza virus administered intranasally induced a Th1, Th2, and Th17 response, whereas the same vaccine administered intramuscularly only induced Th1 and Th2 [62].
Killed whole-cell cholera vaccines (Dukoral®, ShanChol™, and Euvichol-Plus®, available outside the USA) are mucosal vaccines approved by the WHO, using an inactivated whole organism [63,64,65]. VAXCHORA®, a live attenuated bacterial vaccine, is another FDA-approved single-dose oral vaccine for people between the ages of 2 and 64 traveling to areas of active cholera transmission [66,67,68]. There is research into other live attenuated oral bacterial vaccines for tuberculosis and enterotoxigenic E. coli and a live attenuated intranasal vaccine for pertussis [69,70,71,72,73,74].
Although much effort has been directed at the development of mucosal vaccines, significant challenges remain. The number of approved mucosal vaccines for humans is low compared to the number of pathogens that enter a mucosal surface. Only 8 out of 96 licensed vaccines in the United States are mucosal (https://www.fda.gov/vaccines-blood-biologics/vaccines/vaccines-licensed-use-united-states (accessed on 21 January 2024)).
Table 1. List of FDA and WHO-approved mucosal vaccines for human use [75,76,77].
Table 1. List of FDA and WHO-approved mucosal vaccines for human use [75,76,77].
FDA-Approved Vaccines
PathogenProduct NameTrade NameManufacturerDelivery Route (Form)FormulationInitial Approval
Adenovirus Types 4 and 7Adenovirus Type 4 and Type 7 Vaccine, Live, OralAdenovirus Type 4 and Type 7 Vaccine, Live, OralBarr Labs, Inc. North Wales, PA, USAOral (tablets)Live attenuated2011 (military use only)
Influenza Type A (H1N1)Influenza Vaccine, Live, Intranasal (Monovalent, Type A H1N1)Influenza A (H1N1) 2009 Monovalent Vaccine Live, IntranasalMedImmune, LLC Gaithersburg, MD, USAIntranasal (spray)Live attenuated2003
Influenza Types A and BInfluenza Vaccine, Live, Intranasal (Trivalent, Types A and B)FluMistMedImmune, LLC Gaithersburg, MD, USAIntranasal (spray)Live attenuated2003
Influenza Types A and BInfluenza Vaccine, Live, Intranasal (Quadrivalent, Types A and B)FluMist QuadrivalentMedImmune, LLC Gaithersburg, MD, USAIntranasal (spray)Live attenuated2003
RotavirusRotavirus Vaccine, Live, OralRotarixGlaxoSmithKline Biologicals Rixensart, BelgiumOral (liquid)Live attenuated2008
RotavirusRotavirus Vaccine, Live, Oral, PentavalentRotaTeqMerck Sharp & Dohme Corp Whitehouse Station, NJ, USAOral (liquid)Live attenuated2006
Salmonella typhiTyphoid Vaccine, Live, Oral, Ty21aVivotifBerna Biotech, Ltd. Berne, SwitzerlandOral (capsule)Live attenuated1989
Vibrio choleraeCholera Vaccine, Live, OralVaxchoraEmergent Travel Health, Redwood City, CA, USAOral (liquid)Live attenuated2016
WHO (but not FDA)-Approved Vaccines
PathogenProduct NameTrade NameManufacturerDelivery Route (Form)FormulationInitial Approval
Influenza Type A (H1N1)Influenza Vaccine, Live, Oral, PandemicNasovacSerum Institute of India Pvt. Ltd., Hadapsar, IndiaIntranasal (spray)Live attenuated2012
Influenza Type A (H5N1)Influenza Vaccine, Live, PandemicPandemic Live Attenuated VaccineAstraZeneca Pharmaceuticals LP., Nijmegen, NetherlandsIntranasal (spray)Live attenuated2020
Influenza Types A and BInfluenza Vaccine, Live, Oral, SeasonalNasovac-SSerum Institute of India Pvt. Ltd., Hadapsar, IndiaIntranasal (spray)Live attenuated2015
PoliovirusPolio Vaccine—Oral (OPV) Bivalent Types 1 and 3Biopolio B1/3Bharat Biotech International Limited, Hyderabad, IndiaOral (liquid)Live attenuated2017
Bivalent OPV Type 1 and 3 Poliomyelitis Vaccine, Live (Oral)Panacea Biotec Ltd., Malpur, IndiaOral (liquid)Live attenuated2018
Bivalent Oral Poliomyelitis Vaccine Type 1 & 3 (bOPV 1 & 3)PT Bio Farma (Persero), Bandung, IndonesiaOral (liquid)Live attenuated2010
Bivalent type 1 & 3 Oral Poliomyelitis vaccine, IP (bOPV1 & 3)Haffkine Bio Pharmaceutical Corporation Ltd., Mumbai, IndiaOral (liquid)Live attenuated2010
Oral Bivalent Types 1 and 3 Poliomyelitis VaccineSanofi Pasteur, Lyon, FranceOral (liquid)Live attenuated2011
Polio Sabin One and ThreeGlaxoSmithKline Biologicals SA, Rixensart, BelgiumOral (liquid)Live attenuated2009
Poliomyelitis Vaccine (live, oral attenuated, human Diploid Cell), type 1 and 3Beijing Institute of Biological Products Co., Ltd., Beijing, ChinaOral (liquid)Live attenuated2017
Poliomyelitis Vaccine (Oral), Bivalent types 1 and 3Serum Institute of India Pvt. Ltd., Hadapsar, IndiaOral (liquid)Live attenuated2013
PoliovirusPolio Vaccine—Oral (OPV) Monovalent Type 1Monovalent Oral Poliomyelitis Vaccine Type 1 (mOPV1)PT Bio Farma (Persero), Bandung, IndonesiaOral (liquid)Live attenuated2009
Monovalent type 1 Oral Poliomyelitis vaccine, IP (mOPV1)Haffkine Bio Pharmaceutical Corporation Ltd., Mumbai, IndiaOral (liquid)Live attenuated2009
Polio Sabin Mono T1GlaxoSmithKline Biologicals SA, Rixensart, BelgiumOral (liquid)Live attenuated2009
PoliovirusPolio Vaccine—Oral (OPV) Monovalent Type 2Monovalent Oral Poliomyelitis Vaccine Type 2PT Bio Farma (Persero), Bandung, IndonesiaOral (liquid)Live attenuated2019
ORAL MONOVALENT TYPE 2 POLIOMYELITIS VACCINE (mOPV2)Sanofi Pasteur, Lyon, FranceOral (liquid)Live attenuated2016
Polio Sabin Mono Two (oral)GlaxoSmithKline Biologicals SA, Rixensart, BelgiumOral (liquid)Live attenuated2011
PoliovirusPolio Vaccine—Oral (OPV) Monovalent Type 3Polio Sabin Mono Three (oral)GlaxoSmithKline Biologicals SA, Rixensart, BelgiumOral (liquid)Live attenuated2010
PoliovirusPolio Vaccine—Oral (OPV) TrivalentBIOPOLIOBharat Biotech International Limited, Hadapsar, IndiaOral (liquid)Live attenuated2015
Oral Poliomyelitis Vaccines (Oral Drops)PT Bio Farma (Persero), Bandung, IndonesiaOral (liquid)Live attenuated2020
Poliomyelitis vaccine (Oral) IP Haffkine Bio Pharmaceutical Corporation Ltd., Mumbai, IndiaOral (liquid)Live attenuated2006
RotavirusRotavirus Vaccine, Live, OralRotasilSerum Institute of India Pvt. Ltd., Hadapsar, IndiaOral (liquid)Live attenuated2018
Rotasil ThermoSerum Institute of India Pvt. Ltd., Hadapsar, IndiaOral (liquid)Live attenuated2020
RotavacBharat Biotech International Limited, Hyderabad, IndiaOral (liquid)Live attenuated2018
Rotovac 5DBharat Biotech International Limited, Hyderabad, IndiaOral (liquid)Live attenuated2021
Vibrio choleraeCholera Vaccine, Inactivated, OralDukoralValneva Sweden AB, Stockholm, SwedenOral (liquid)Inactivated (with recombinant cholera toxin subunit B)2001
EuvicholEuBiologics Co., Ltd., Chuncheon-si, South KoreaOral (liquid)Inactivated2015
ShanCholSanofi Healthcare India Private Limited, Medchal, IndiaOral (liquid)Inactivated2011
The veterinary vaccine industry has many more licensed mucosal vaccines, targeting a variety of pathogens and given to many different types of animals, including cattle, poultry, and fish. In addition to the immunological benefits, these vaccines are cheaper and easier to administer to large groups than parenteral vaccines [78]. Many swine and some poultry vaccines can be administered in drinking water, including Coliprotec F4, which consists of avirulent live E. coli and protects against post-weaning diarrhea in swine [79]. Live attenuated Salmonella poultry vaccines can be administered to flocks using a large sprayer [80,81,82]. While most mucosal vaccines are live, the rabies vaccine for control of rabies in wildlife uses a viral vector to express the rabies glycoprotein and is administered orally in bait [83,84].
In addition to a need for mucosal vaccines against more pathogens, there is also a need to develop new, more effective, less expensive mucosal vaccine platforms. Many of the vaccine platforms invented in the last hundred years and used for parenteral vaccines have not yet been successfully applied to mucosal vaccines or only applied in limited ways with minimal successful dissemination worldwide. For example, the intranasal hepatitis B vaccine, HeberNasvac®, which uses recombinant virus-like particles, represents one successful application, but it is only approved in Cuba [54].
There are many new platforms in the research phase for mucosal vaccines. Some include viral vectors, which can mimic natural infections and elicit both an innate and adaptive immune response. At least one viral-vector vaccine for SARS-CoV-2 is in the human clinical trial stage, and a viral-vector vaccine for tuberculosis has been tested in mice [85,86,87]. A non-viral vector (N1,N3-dicarbamimidoyl-5-methylisophthalamide (BGG)) carrying antigen-expressing plasmid DNA has also been developed and used in an intranasal SARS-CoV-2 vaccine [88,89]. Oral mRNA vaccines are also being developed [90]. Several intranasal vaccines that combine protein subunits and adjuvants are being designed and tested against pertussis, tuberculosis, and SARS-CoV-2 [91,92,93,94,95,96]. Outer membrane vesicles, which have previously been used in parenteral meningitis B vaccines, are now being used for intranasal pertussis and SARS-CoV-2 vaccines [97,98,99]. Recent progress on improved vaccines against pertussis and tuberculosis via nasal delivery demonstrates that they elicit more effective immunity compared with traditional vaccine formulations that have been used for decades [100,101]. Broad spectrum antigen-specific antibody responses and cellular-mediated immunity are achieved after those mucosal vaccinations in clinical trials, even with a single dose [101]. Sustained protective immunity up to 12 months after vaccination [101] and sterilizing immunity was observed in animal models [100].
Additionally, our lab is developing a vaccine platform that uses killed whole-cell genome-reduced bacteria, that uses an autotransporter to express the antigen of interest and immune modulators that can be used non-parenterally [102].

3. The Mucosal Barrier

The body’s mucosal surfaces have an enormous area, about 200 times greater than the skin. Mucosae line many important organs and have been classified into two different types (Figure 1). Type I mucosa are found in the respiratory tract, most of the gastrointestinal tract, tonsils, and adenoids, whereas type II mucosa includes the oral cavity and urogenital tract [42,43,103]. The mucosal surface is where a pathogen present in the environment first encounters and then enters the host [104]. The mucosal barrier, in most organs, is just a single layer of epithelial cells separating the environment from the interior, serves as both a barrier to pathogens and other potentially harmful factors in the environment and the interior of the organism, and in some organs, a surface that exchanges essential nutrients and gases. The mucosal barrier also, in some organs, like the respiratory and GI tracts, enables the immune system to sense and sample the environment for antigens and transmit those antigens on to the immune system to initiate a response.
Most mucosal tissue consists of three layers: epithelium, lamina propria, and muscularis mucosae. The mucosal epithelium lines the inner surfaces of organs and body cavities that are exposed to the external environment. The epithelium is composed of cells joined by tight junctions and faces a complex environment rich in microorganisms [42]. Epithelial cells protect the host from irritants and toxins by secreting a thick, gel-like mucus and arrange themselves in different ways to accomplish different functions, such as epithelia having one or several layers of cells [105]. Epithelial cells are one of the main types of cells in the human body. They are found in the lining of organs and body cavities, and they also form glands. Epithelial cells have many functions, including secretion, absorption, sensation, protection, and transport [106,107]. Mucosal epithelium can be defined as the following different types [108] (Figure 2): (1) Nonkeratinized epithelium, which is found in the oral cavity, esophagus, vagina, and anus is composed of several layers of cells, with the outermost layer being moist and alive. This type of epithelium mostly provides protection against mechanical stress and abrasion [107,108,109]; (2) columnar epithelium is a type of epithelium found in the stomach, intestines, respiratory tract except for alveoli, and gallbladder. It is composed of tall, narrow cells that are tightly packed together and are mainly specialized for absorption and secretion [107,108,109]; (3) squamous epithelium is a type of epithelium found in the lungs, blood vessels, and body cavities. It is composed of flat, scale-like cells that are tightly packed together. This type of epithelium serves as a barrier against infection and injury [107,108,109]. In the lungs, in the alveoli, gas exchange occurs across thin, Type I squamous epithelial cells, while cuboidal Type II cells secrete surfactant. In other parts of the respiratory tract, the epithelia include ciliated cells, nonciliated columnar cells, mucous (goblet) cells, brush cells, basal cells, and associated submucosal cells. Ciliated cells play critical roles in mucociliary clearance to propel pathogens and inhaled particles trapped in the mucous layer out of airways by beating in metachronal waves of cilia [110]. The classification of epithelial cells into different types according to their shapes and cell layer numbers reflects their physiological and host defense functions. Simple squamous cells are thin, flat cells that line the blood vessels, lungs, and body cavities. They are involved in the exchange of gases and nutrients [107,108]. Simple cuboidal cells are cube-shaped cells that line the kidney tubules and the ducts of glands. They are involved in secretion and absorption [106,108]. Simple columnar cells are elongated cells that line the digestive tract and the uterus. Their major functions are secretion, absorption, and protection [111]. Stratified squamous cells are flat cells that form the outer layer of the skin and the lining of the mouth, throat, and vagina. They are mainly involved in physical protection [112]. Stratified cuboidal cells are cube-shaped cells that form the lining of the sweat glands and the salivary glands. Their functions are mostly in secretion [108]. Stratified columnar cells are elongated cells that form the lining of the male urethra and the ducts of some glands [107]. They are involved in secretion and protection. Pseudostratified columnar cells are elongated cells that line the respiratory tract. They are involved in secretion and protection.
The epithelial layer is the site of immunological activities and constitutes both a physical and biochemical barrier that enables the transfer of nutrients, antigens, and other immune factors while resisting the transmission of pathogens and toxins [42,43,105]. Epithelia, together with associated glands, provide innate non-specific defenses via the production and secretion of materials such as mucins, antimicrobial proteins, enzymes, secretory antibodies, buffers, prebiotics for microbiome modulation, salts, water, hormones, and other bioactive molecules, and other components of the intraluminal mucosal environment [42,113,114]. The epithelium of many organ systems also has important neural sensing, including temperature and molecular sensors for taste and smell and neuroendocrine functions [115,116,117,118].
Epithelial cells replace themselves frequently; they have a high cell turnover rate, which helps to protect against pathogen invasion. Some epithelia include cells with cilia, which move mucus, and small particles trapped in mucus, including pathogens. Besides their physical barrier function, epithelial cells play active and important roles in mucosal immune defense. They can function in immune surveillance by sensing dangerous pathogen components through pattern-recognition receptors (PRRs), such as Toll-like receptors (TLRs), to stimulate innate and adaptive immune responses by secreting cytokine and chemokine signals as a result of cell activation following receptor binding, to signal underlying mucosal immune cells, like dendritic cells (DCs) and macrophages [42,43,109,119]. Some epithelial cells have specialized functions in immune sensing. For example, M cells are found in the epithelium of the gut-associated lymphoid tissues (GALT) and are involved in the uptake and transport of antigens from the lumen of the gut to the underlying immune cells [120]. Similarly, analogous cells in the respiratory epithelium are involved in the uptake and transport of antigens from the airway lumen to the underlying immune cells [121]. In addition, other components of the epithelium, like tuft cells, are found in the respiratory and gastrointestinal epithelium and are involved in the detection of parasitic infections [122,123] as well as intraepithelial lymphocytes (IELs) that function in immune surveillance [124,125].
The middle layer of the mucosa is called lamina propria, which is a loose connective tissue attached to the epithelium and composed of connective tissue, nerves, blood and lymphatic vessels, and associated lymphoid tissue. The major functions of the lamina propria include a structural role, supplying blood to the epithelium, hosting lymphoid tissues, and binding epithelial structures to smooth muscle [113,126]. Lamina propria neural components work together with smooth and striated muscle to help modulate the conformation of the epithelium [127].
The muscularis mucosae, a layer of smooth muscle, is the deepest layer of the mucosa. It varies in thickness throughout the digestive, respiratory, and genitourinary tracts and, in humans, is most active in the stomach and GI tract. The muscularis mucosae provides a motor function that keeps the mucosa in motion, helping the mucosa to stretch and contract along with the organs [113,126]. Impairment of the mucosal barrier can lead to impaired immune functions and responses as well as inflammation [128,129].

4. Mucosal Immunity

Mucosal immunity is the immune response that occurs at the mucosal surfaces of the body. The mucosal immune system constitutes a complex network of innate immune components, immune cells, and antibodies working together to maintain the integrity of the mucosal barrier and mediate immune responses to pathogens (Figure 1) [43,45,103]. An ideal mucosal vaccine should elicit durable, effective local immunity. Detailed knowledge of the mucosal immune system can help inform the development of new mucosal vaccines.
Innate immune responses activate the adaptive immune system, which is composed of T and B lymphocytes that can recognize specific antigens from different pathogen-derived molecules. The mucosal cellular immune response begins with antigen-presenting cells (APC). DC cells are classical APCs. They process captured antigens and, after activation and maturation, present antigens to naïve T-cells, yielding clonally expanded and differentiated Th cell subsets including Th1, Th2, Th17, and Treg cells. Each cell type subset has cell type-specific cytokine secretion profiles [104,130]. Once activated, the lymphocytes proliferate and differentiate into effector cells that can eliminate the pathogen via direct cell killing by T-cells and/or immunoglobulins, aided by secretion of cytokines and chemokines that promote inflammation and serve as chemoattractants for additional lymphocytes, monocyte/macrophages, neutrophils, and in some circumstances, eosinophils and basophils [103,131].
The mucosal immune system is comprised of anatomic and physiologic components with distinct functions, which can vary in the mucosa of different tissues [132,133]. Based on functional and anatomical properties, mucosal tissues can be divided into inductive and effective sites [132]. The mucosal defense mechanisms can be defined as physical barriers (epithelial lining, mucus, cilia function) and biochemical factors (pH, antimicrobial peptides) [134]. Antimicrobial peptides (AMPs) are a class of chemical factors produced by epithelial cells and immune cells, often at increased levels in response to microbial invasion [135,136,137,138]. They are small peptides with broad-spectrum antimicrobial activity against infections. Mechanisms of action of AMPs include disrupting the microbial cell membrane, inhibition of protein synthesis, and inducing an oxidative stress response [135]. AMPs are an important component of the non-adaptive immune system because they provide immediate protection against pathogens before the adaptive immune system can mount a specific response [138]. In addition to their antimicrobial activity, AMPs also have immunomodulatory functions, such as promoting wound healing, regulating inflammation, and enhancing the recruitment of immune cells [136].
In addition to the AMPs, the microbiome of mucosal compartments is another component of the mucosal immune system, consisting of a complex ecosystem of microorganisms that inhabit the mucosal surfaces of the body and lumens of body cavities [139,140,141]. The microbiome plays an important role in maintaining the health of the host by regulating the immune system, protecting against pathogens, and aiding in digestion [139]. The composition of the microbiome varies depending on the location of the mucosal sites. The microbiome of the respiratory tract is dominated by bacteria such as Streptococcus, Haemophilus, and Moraxella [140], while the microbiome of the gastrointestinal tract is more diverse and includes bacteria such as Bacteroidaceae, Firmicutes, and Actinomycetota [140]. The microbiome of mucosal compartments can interact with the host immune system by stimulating the production of antimicrobial peptides to protect against pathogens and regulating the development and function of immune cells, T-cells, and B-cells [142]. The microbiome of some mucosal compartments can help defend against invading pathogens by altering the mucosal environment to make it more hostile to invading pathogens, occupying ecological niches to exclude invaders, and producing antimicrobial substances, and the host can work to modulate the microbiome to help accomplish this. An interesting example is the female genital tract. A healthy vagina has a low pH, promoted by the production of organic acids produced by Lactobacillae, the dominant component of a healthy vaginal microbiome. The vaginal epithelium secretes glycogen, which helps “feed” the lactobacilli, so the vaginal epithelium helps defend against potential pathogens by providing endogenous prebiotics that help modulate the microbiome to provide a more effective defense [143,144,145,146,147,148].
Cytokines secreted by mucosal cells play crucial roles in controlling the growth and activity of other immune cells by establishing cell–cell interactions and facilitating cellular signaling. The tightly regulated cytokine signaling network regulates both innate and adaptive immune responses across mucosal barriers [149,150].
The mucosal effector sites are comprised of the lamina propria regions in mucosal barriers of the upper respiratory, gastrointestinal, and reproductive tracts, along with secretory glandular tissues. There are several different types of antigen-specific mucosal effector cells in effector sites, including T helper cells (TH), regulatory T-cells (Tregs), and IgA-producing plasma cells, which, together, compose the humoral and cellular immune responses localized in mucosal sites [132,133,151]. Humoral immune responses are mediated by immunoglobulin molecules produced by plasma cells [152]. Although some IgG is excreted into mucosal cavities, IgA is the predominant immunoglobulin isotype in mucosal immunity and widely exists in mucosa in the respiratory, gastrointestinal, and urogenital tracts. It is also present in tears, saliva, and colostrum [153,154]. IgA in sera is found as monomers. In contrast, mucosal IgA is polymeric [153,155]. The induction of SIgA, secretory IgA, serves as the first line of defense to protect the mucosal surface from dangerous substances such as toxins and microorganisms by preventing their access and entry through epithelial receptors. Major IgA production sites in the mucosa are gut-associated lymphoid tissues (GALT), including Peyer’s patches (PPs), isolated lymphoid follicles, and mesenteric lymph nodes (MLNs), producing more than 90% of the SIgA made by the body [153,156]. PPs are the principal precursor source of IgA-producing plasma cells [157], secreting an average of 3 g of SIgAs into the gut lumen of an adult human each day [158]. Antigens are taken up by mucosal immune cells such as dendritic cells (DCs) [159], follicle-associated epithelium containing microfold (M) cells [160], follicle-associated epithelium (FAE) [161], and small intestine goblet cells (GCs) [162]. Upon antigen processing and presentation, APCs, T-cells, and B-cells are activated, and IgA class switch recombination and somatic hypermutation (SHM) are later initiated in the mucosal B-cells [163,164] in which activation of SMAD3/4 and Runx3 downstream of the TGF-β recognition signaling pathway also contributes to IgA production [154,164,165]. The mechanisms of IgA class switch recombination are divided into T-cell-dependent (TD), which requires the interaction of CD40 on B-cells with CD40L, derived from T-cells, resulting in high-affinity antigen-specific neutralizing IgA production [166], and T-cell-independent (TI) which produce commensal-reactive IgA through stimulation of B-cell activating factor of the TNF family (BAFF) and A proliferation-inducing ligand (APRIL) secreted from innate immune cells such as innate lymphoid cells (ILCs) and plasmacytoid dendritic cells (pDCs) [166,167,168,169]. Human B-cells produce two subclass of IgA molecules, IgA1 and IgA2, with similar receptor binding profiles but different topographies [164]. IgA1 exists in both systemic and mucosal districts, whereas IgA2 is mostly present in mucosal districts colonized by plentiful microbiota, including the mucosa of the distal intestinal tract and the urogenital tract [164].
B-cell activation occurs at multiple mucosal sites, promoting B-cell differentiation and maturation and production of antigen-specific IgG antibodies as well as secretory IgA (SIgA), which can neutralize the pathogens and prevent invasion through the mucosal surface by antibody neutralization activities or Fc-mediated functions including Antibody-Dependent Cellular Cytotoxicity (ADCC), Antibody-Dependent Cellular Phagocytosis (ADCP), and Antibody-Dependent Cell-mediated Viral Inhibition (ADCVI) [103,170,171]. IgG excretion is mediated by the neonatal Fc receptor (FcRn) [172], whereas IgA molecules are secreted via polymeric Ig receptor (pIgR), which is an epithelial membrane receptor that interacts with and transcytoses J chain-containing dimeric IgA (and pentameric IgM) produced by lamina propria plasma cells [173]. Most of the antibodies produced in the host take place locally in the mucosal surfaces where the predominant form is the dimeric IgA by pIgR-mediated export, which demonstrates the importance of IgA in mucosal immunity as most pathogens are encountered by the host mucosal surfaces [151,173]. Besides pathogen clearance via ‘immune exclusion’, SIgA also has other functions like reinforcement of the epithelial barrier, facilitation of APC to capture antigens by promoting downstream local immune responses, and excludes the pathogenic microorganisms from the epithelial surface by forming a biofilm of non-pathogenic microorganisms as a result of anchoring within the mucus [151,174].
IgG is the most abundant antibody in blood and is involved in the systemic immune response. IgG can also be found in type I mucosae, such as the respiratory and gastrointestinal tracts; specific IgG is increased in response to local infections or vaccination [154,156,175,176,177]. IgA is mostly found in type I mucosae like the upper respiratory tract, whereas the IgG is the predominant protective isotype in type II mucosae, such as corneal, oral, esophageal, lower respiratory tract, and lower female reproductive tracts, due to the expression of FcRn but not pIgR [175]. The production of IgG in mucosal immune responses is regulated by a complex interplay between local immune cells and the systemic immune system [176]. B-cells in mucosal tissues can be activated by APCs. Activated B-cells can then differentiate into plasma cells that produce IgG [177]. The production of IgG in mucosal tissues can also be influenced by cytokines and chemokines produced by immune cells in response to infection or inflammation [121,177,178]. In addition, some IgM could be found in respiratory and urogenital tract mucosa, whereas IgD constitutes a significant fraction of antibodies in the upper segments of human respiratory and digestive tracts [154].
The mucosal immune sensing sites are comprised of mucosa-associated lymphoid tissue (MALT), including the gut-associated lymphoid tissues (GALT), nasopharyngeal-associated lymphoid tissue (NALT), and lymphoid sites. MALT is composed mostly of dendritic cells, macrophages, innate lymphoid cells, mucosal-associated invariant T-cells, intraepithelial T-cells, regulatory T-cells (Treg), as well as IgA secreting plasma cells and a source of memory B- and T-cells that migrate to effector sites to elicit immune responses [103,132].
A typical pathogen invasion begins from exposure through routes such as inhalation, ingestion, or direct contact with a mucosal surface, such as sexual intercourse. Then, pathogens attach to the epithelium and penetrate through the mucosal layer to enter the host either by infecting epithelial cells or by gaining access to cells in the submucosa through a microabrasion in the mucosal barrier, or transport across a mucosal cell or through the actions of a mucosal immune cell, such as dendritic cells, followed by establishment of target cell infection in the lamina propria [175]. The immune system responds to pathogens in the mucosa by initiating a series of immune responses, which are composed of a variety of immune cells, including T-cells, B-cells, macrophages, dendritic cells (DCs), natural killer (NK) cells, innate lymphoid cells (ILCs), and plasma cells. When the mucosa encounters a pathogen, the pathogen is first recognized by the innate immune system, which is composed of antigen-presenting immune cells through PRRs. These innate immune cells can engulf and destroy pathogens or present them to other immune cells for further processing [179]. The follicle-associated epithelium containing microfold (M) cells, one type of lymphoid follicle cell found in MALT, are responsible for antigen capture and uptake in the local lumen sites, followed by antigen presentation to APCs like DCs and macrophages, where the antigens are further processed and presented to lymphocytes [179,180,181]. Th cells play critical roles in mucosal immune responses, such as the Th17 cell, a Th subset that mostly secretes interleukin-17 (IL-17), which has been reported to help promote protective mucosal immunity induced by vaccines with increased SIgA levels by upregulating the polymeric Ig receptor expression levels in mucosal epithelial cell surfaces and promote T-cell-independent B-cell maturation and differentiation [103,104].
After mucosal vaccination or pathogen infection, a subset of memory lymphocytes, known as tissue-resident memory cells, are preferentially elicited, while little to no numbers of those cells were observed following parenteral immunizations [182]. Those cells, maintained within non-lymphoid barrier tissues, act as sentinels, providing rapid local recall of immune responses to previously encountered antigens, driving accelerated pathogen clearance via mechanisms that involve enhanced cytokine expression and functional effector molecules secretion [61]. Reactivation of long-lived tissue-resident memory lymphocytes can trigger enhanced innate and adaptive immune responses in the host as a whole, beyond the mucosal compartment, mediated by cytokines [61].
When local immune responses are activated after mucosal vaccination, antigens are taken up and processed by APCs, like mucosal epithelial cells and DCs [47]. Then, immune cells are polarized by the antigen-laden APCs, and those activated immune cells migrate via lymphatic vessels to draining lymph nodes, which promotes a systemic immune response, leading to further T- and B-cell proliferation, differentiation, and maturation. Activated T- and B-cells then circulate through the bloodstream, establishing systemic immunity [42]. During the crosstalk between mucosal and systemic immunity, activated lymphocytes express specific homing receptors that guide them in their migrations between mucosal and systemic sites. This creates a feedback loop where systemic immune cells contribute to local mucosal immunity and vice versa [42,47]. Cytokines produced by both local and systemic immune cells also play a crucial role in coordinating the overall immune response, such as IL-17 produced by Th17 cells, promoting both local mucosal inflammation and systemic antibody production [42,47].
Immunological memory was long defined as a characteristic only belonging to adaptive immunity. This traditional thinking has been challenged recently as evidence has been developed showing that, after infection or vaccination, some prototypical innate immune cells, such as myeloid cells or natural killer cells, can also display long-term nonspecific heterologous protective immunity when encountering a second stimulation [101,183,184]. This previously unrecognized property of immune phenomena of hyper-reactive types of innate immune memory, which leads to an enhanced innate immune protective outcome, has been termed “trained innate immunity” (TII [183]). Mechanistic studies have demonstrated epigenetic, metabolic, and functional reprogramming of myeloid cells and myeloid progenitors in the bone marrow and peripheral tissues, such as gut and lung mucosa, that mediate a persistent change in the homeostatic state of the innate immune cells that endure long after the initial antigenic or microbial exposure/clearance, with increased responsiveness and production of inflammatory mediators upon secondary stimulation, as well as enhanced capacity to eliminate infection. This yields an enhanced innate host defense against the same or unrelated pathogens [183,185]. TII is now recognized as a potential correlation of protection against M. tuberculosis infection, and studies also reveal that the respiratory mucosal route of TB vaccination-induced robust memory airway macrophages in the lung provides a type of TII capable of potent protection against M. tuberculosis in the early stage of infection, independent of T-cell immunity [101,186]. Thus, the concept of TII has received renewed attention not only in light of new knowledge with regard to innate immune memory but also in aiding in designing new generations of vaccines that combine classical immunological memory and TII.

5. Mucosal Vaccine Formulation and Delivery Strategies

Knowledge of the mucosal immune system can inform mucosal vaccine design, enhancing effective humoral immune responses mediated by IgA and IgG through the promotion of optimal Th cell responses with maximal B-cell stimulation, more profound CTL responses by incorporation of well-defined protection-correlated antigen-specific T-cell epitopes in the context of Human leukocyte antigens (HLA) alleles, incorporation of adjuvants that can augment innate immune responses by APCs to elicit enhanced and long-lasting protective immunity, and technologies like nanoparticles, liposomes, or virus-like particles to improve the antigen delivery efficacy to mucosa [23,187].
Several currently approved mucosal vaccines are mostly based on traditional live attenuated or killed whole-cell (KWC) vaccines. Live attenuated vaccines have been extensively explored for mucosal immunization in humans and animals for decades. They are made from weakened versions of infectious pathogens, produced using physical, chemical, or biological methods, and contain antigens that closely mimic a true infection [188]. The most important advantages of live attenuated vaccines include long-lasting immunity, high efficacy, and non-parenteral administration. The manufacture and storage of those vaccines, however, can be more difficult than other types of vaccines. In addition, since the agent is living, there is a chance that the agent can revert to wild type or that secondary mutations may transform attenuated vaccine strains into more contagious and virulent forms [47,188,189].
Killed whole-cell (KWC) bacterial vaccines are produced from inactivated or killed microorganisms that are incapable of replicating within the hosts. KWC vaccines are typically more stable, requiring easier-to-maintain storage conditions [188]. They can also be lyophilized and stored at ambient temperature for long times without affecting stability and efficacy, like inactivated whole-cell-based cholera vaccines containing cholera toxin B subunit [190]. However, a replicating or quasi-replicating vaccine, like a viral-vectored vaccine, may elicit better immune responses, as initial increases in antigen exposure may be interpreted by the immune system as a signal of a threatening infection [191], although dosing schedules or dose delivery technologies to non-live or quasi-live vaccines may yield improve immune responses. Several studies have shown that limited antigen exposure can lead to poor immunogenicity, especially to highly conserved epitopes, which could potentially induce epitope-specific broadly immune responses but are mostly immune subdominant in nature, such as influenza hemagglutinin stalk domain or the fusion peptides from some envelope proteins of viruses with Type 1 viral fusion proteins [102,192,193,194,195]. To overcome this, an adjuvant can be included with a KWC vaccine to augment efficacy by stimulating immune responses or modulating the presentation of antigen, but with potential risks of increased reactogenicity and adverse reactions like inflammation [189]. Some non-replicating vaccines require booster doses and reimmunization [47,189]. However, the disadvantages seen with KWC vaccines exist for all non-replicating vaccines, including, for example, vaccines based on proteins and polysaccharides and mRNA vaccines. A significant advantage of KWC vaccine vector platforms and the other non-replicating vaccine platforms is that antivector immunity will not be a barrier to the use of the platform for multiple different vaccines. Antivector immunity can create significant problems for viral-vectored vaccines, for example, adenovirus or poxvirus-vectored vaccines [196,197,198], making it difficult to use the same virus vector for more than one disease antigen. For KWC vaccines, antivector immunity could even have a potential adjuvating effect.
A vaccine approach that can be particularly useful for mucosal vaccines is subunit vaccines, which are made from isolated components of microorganisms instead of whole organisms. Subunit vaccines are often easier to produce and store than traditional live attenuated vaccines. Subunit vaccines are the basis for many vaccine technologies, like isolated protein and polysaccharide and mRNA vaccines. Subunit vaccines have the advantage that they aim to elicit immune responses against the component of the pathogen that is known to be the specific target of immune responses that confers protective immunity against the pathogen, for example, a viral envelope protein involved in virion binding and entry into host cells, or microbial toxins that are responsible for the pathogenesis of disease. Another compelling advantage of subunit vaccines is that they can be designed to elicit immunity against a part of a pathogen antigen that is highly conserved across different pathogen variants, enabling a route to a variant-resistant vaccine or even a vaccine that provides protection against a large family of pathogens. A vaccine that can elicit protection against a highly conserved antigen may be particularly useful for pathogens that rapidly evolve variants within an infected individual, like HIV, or across a population, like coronaviruses. One licensed mucosal vaccine based on pentameric cholera toxin B subunit (CTB) effectively prevents disease. CTB has promoted the effective delivery of bound heterologous antigens to mucosal APCs, such as enterocytes, M cells, macrophages, and DCs [199,200,201,202]. Human clinical studies have demonstrated that CTB included in a vaccine preparation promotes the induction of broadly reactive antigen-specific local IgA and systemic IgG responses via rectal and intranasal administration [203,204,205]. Recent studies have also shown that CTB elicits mucosal T-cell immune responses [206,207], including local and intestinal protective tissue-resident memory T (TRM)-cells [47,208].
Consideration of the advantages and disadvantages of traditional vaccine formulations has led to the development of several next-generation mucosal vaccines. These include viral vector-based vaccines, particle-based vaccines, and vaccine platforms employing native or modified whole bacterial cells to deliver and present antigens. Viral vector-based vaccines use biomodified replicating or non-replicating viral vectors to deliver DNA or RNA instructions for antigen production to vaccine cells, enabling antigen production by the cells, followed by the induction of immune responses against the vaccine antigen expressed by the cells. Several viral vectors have been successfully developed as vaccine-delivery vectors. A widely used approach employs adenoviral vectors to deliver the genetic instructions. Adenoviral vector vaccines can elicit both strong humoral and cellular immune responses, even to the point of sterilizing immunity [209,210,211,212,213,214,215]. Viral vector-based vaccines have been shown to elicit good immunity against diseases that are transmitted across mucosal surfaces with ease of administration [216,217,218,219]. The formulation and delivery of viral vector-based vaccines can vary depending on the type of vaccine and the disease it is designed to prevent. One major hurdle of viral vector-based vaccine application is that previous exposure to the vector can reduce effectiveness, which prevents multiple uses of the same vector. Pre-existing immunity to microbial vectors can limit the vaccine’s effectiveness, even for priming with viral-vectored vaccines [47]. Although a recent human clinical study of a COVID-19 mucosal vaccine based on recombinant adenovirus serotype 5 (Ad5) vector has shown some efficacy, with safety and tolerability [220], there are still concerns about how vaccine efficacy may be affected by pre-existing antivector immunity, as observed in two HIV-1 vaccine human clinical studies, Step and HVTN 503/Phambili, more than a decade ago [221,222,223]. Both studies used recombinant adenovirus serotype 5-vectored vaccines and failed after phase 2b studies as it was found that populations who were Ad5 seropositive before vaccination had an increased risk of HIV-1 acquisition after vaccination [224,225]. Later, several follow-up studies explored the potential underlying mechanisms of how vector-specific pre-existing immunity correlates with risk of infection, including an association of high frequencies of preimmunization Ad5-specific T-cells with decreased magnitude of HIV-specific CD4 responses and decreased breadth of HIV-specific CD8 responses [226], enhancement of HIV-1 replication in CD4 T-cells by Ad5 immune complexes activation of the dendritic cell–T-cell axis [227], increased susceptibility to HIV infection by Ad5-specific CD4 T-cells [228], and studies on non-human primate challenge models further implicated increased the risk of SIV acquisition from low-dose SIV penile challenge associated with higher pre-immune responses to Ad5 vectors [229], suggesting that pre-existing Ad5 immunity might dampen desired vaccine-induced responses and increase the risks of viral acquisition of the HIV-1 infected population.
Particle-based vaccines are relatively new types of vaccines. These include nanoparticle, microparticle-based vaccines, and virus-like particles (VLPs). Particle vaccines serve not only to deliver the antigen cargo to the host but also to modulate antigen release rates since prolonged antigen release may enhance the level and quality of immune responses [230,231,232]. Some KWC microbial vectors that express vaccine antigens can be considered biologically produced microparticles. Other advantages of particle vaccines include (1) protection of the antigens; (2) due to the size of the particles, they can facilitate capture and uptake by APCs followed by maturation and stimulation of those APCs to further activate innate and adaptive immunities; (3) particle vaccines are versatile platforms to delivery several different antigens or combination of antigen with immune stimulatory molecules. Some particles can also act as adjuvants [232,233]. Indeed, some of the impressive effectiveness of the recently developed mRNA vaccines may be attributed to the adjuvating effects of the lipid nanoparticles used to encapsulate the mRNA component of the vaccines [234]. Several different chemical or biomaterials have been utilized to develop mucosal particle-based vaccines such as chitosan [235,236], poly lactic-co-glycolic acid (PLGA) [236,237], recombinant self-assemble protein subunits [238,239,240], lipid bilayers (FDA approved mRNA COVID-19 vaccines) [241,242], and VLPs that mimic viruses but without infectious components [243,244,245]. Those particle-based mucosal vaccines, either administrated alone or in combination with adjuvants, have been proven to be good at stimulating the immune system and eliciting effective immune protection against infections [230,232,244,246]. Particle-based vaccines beyond the licensed mRNA–lipid nanoparticle vaccines are still in the early stages of development, and more research is needed to determine their effectiveness and safety. However, they have the potential to be a promising new approach to vaccination.
Benefiting from the fast development of modern molecular biotechnologies, traditional whole-cell antigen platforms have the potential for combination with purified recombinant subunits but also for overexpression of desired antigens on surfaces of those whole-cell vectors [47]. Results from several groups attempting to use whole bacterial cells engineered to overexpress antigens on the cell surface suggest that expression of antigens on killed or inactivated whole bacteria cell surfaces offers a promising approach to elicit effective immune protection. The killed bacteria cells not only deliver the antigens but also act as adjuvants and help minimize the required vaccine doses [20,102,247,248]. Moreover, those killed whole-cell vaccines are easily produced by industry on a large scale at a low cost, are stable at refrigerator temperatures for prolonged storage, and are easy to administer through mucosal immunizations [247,248]. All these characteristics make the killed whole-cell vaccine a potentially affordable vaccine candidate, especially for lower-income countries, to help reduce the burden of infectious diseases, particularly in response to pandemics [63].
Mucosal vaccine delivery strategies include administration across various mucosal surfaces like the nose, mouth, and lungs. Compared with traditional injection-based vaccination, mucosal vaccine delivery methods can stimulate both mucosal and systemic immunity to provide not only quicker but also better protection against infections, given that infection most frequently happens across mucosal surfaces. Researchers have also found that some mucosal vaccines can induce strong resident memory T-cell responses in mucosal-associated lymphoid tissues to provide durable protective immunity [249]. The delivery of mucosal vaccines can also vary depending on the type of vaccine and the disease they are designed to prevent. Several strategies have been used to deliver mucosal vaccines, including oral delivery by pill, liquid, dissolvable substrate, or small patches. Intranasal and respiratory delivery mechanisms have included nebulizers, nasal sprays, and inhalers. There is an approved nasal vaccine to prevent influenza and an oral live-attenuated bacterial vaccine, TY21A (Vivotif®), against Salmonella to prevent typhoid fever [43]. The vaccine delivery method, along with factors such as vaccine formulation, target antigen/pathogen, and host immune status, will significantly affect the development of protective immunity.

6. Mucosal Adjuvants

Many challenges hamper the efficacy of mucosal vaccines. One approach to improving mucosal vaccine efficacy involves the use of adjuvants. Mucosal vaccine adjuvants are substances and vaccine components that enhance immune responses to antigens delivered through mucosal surfaces [103,104]. Adjuvants act by various mechanisms, such as increasing antigen stability, optimizing antigen delivery and exposure pharmacokinetics, activating innate immune receptors, modulating mucosal barriers and secretions, recruiting and activating immune cells, and stimulating mucosal cellular immune responses [250,251,252,253]. Mucosal vaccine adjuvants can also allow for a reduced antigen dose or enable good immune responses to vaccines with fewer doses. Mucosal vaccine adjuvants that have been studied include bacterial toxins and their derivatives, CpG-containing DNA, cytokines and chemokines, nanoparticles, and biopolymers (Table 2) [254].
Bacterial enterotoxins and their derivatives have been studied and used for mucosal vaccine adjuvants, including cholera toxin (CT) and heat-labile enterotoxin from E. coli (LT) [279]. These adjuvants induce high IgA antibody levels and long-lasting memory to target antigens when administrated mucosally [280]. The mechanisms underlining the enhanced immunogenicity have also been investigated extensively. Recent studies have demonstrated that bacterial enterotoxin adjuvants promote DC cell migration from the subepithelial dome to the follicular-associated epithelium after vaccine administration, followed by antigen capture and presentation in mesenteric lymph nodes [104,280,281,282]. These results showed the critical role of APC stimulation by adjuvants plays in inducing both local mucosal responses and mucosal surface crosstalk [282]. Enterotoxin adjuvants can also induce mixed Th1/Th2 immune responses [280], and the adjuvants can promote Th17 responses that are correlated with high levels of IgA antibody, indicating the importance of IL-17, expressed by Th17 cells, in promoting vaccine-induced protection [283,284]. Although bacterial enterotoxins and their derivatives serve as effective mucosal adjuvants, toxicity concerns have prevented widespread use in clinical applications [285]. However, some toxin subunits or otherwise modified or mutated toxins are used in approved vaccines, such as the recombinant cholera toxin subunit B used in Dukoral® [41].
Another class of adjuvants include pattern recognition receptor (PRR) ligands. These adjuvants yield enhanced immune responses by toll-like receptor (TLR)-mediated immune cell activation through pathogen-associated molecular pattern (PAMP) recognition. Synthetic oligodeoxynucleotides (ODNs) containing unmethylated CpG is a TLR9 agonist [286]. It activates the innate immune responses to yield strong Th1 responses and proinflammatory cytokines, inducing cytotoxic T lymphocyte (CTL) responses and interferon (INF)-γ secretion [287]. CpG also activates antigen-presenting cells (APCs) by upregulating the expression of MHC, CD40, and CD80 on DCs to enhance antigen processing and presentation [287]. CpG ODNs have been found to stimulate TLR9-expressing B lymphocytes, leading to increased IgA levels [288]. In 2017, the US FDA approved HEPLISAV-B® for hepatitis B, the first vaccine to employ a CpG ODN adjuvant [289]. Other nucleotide-based adjuvants include cyclic dinucleotides, such as c-di-GMP and c-di-AMP, which are ligands for the stimulator of interferon genes (STING) pathway that can induce type I interferon and proinflammatory cytokines to enhance antigen presentation as well as adaptive immune responses [290,291,292]. Cyclic dinucleotides improve mucosal immunity against influenza, tuberculosis, and anthrax [254].
Monophosphoryl lipid A (MPL), derived from lipopolysaccharide (LPS) of Gram-negative bacteria, is another promising adjuvant that stimulates immune responses by activating TLR4 [293]. When used in complex emulsion liposome formulations, plus the active fraction of the bark of Quillaja saponaria (QS-21) or alum, MPL induces strong, antigen-specific CD4+ T-cell responses [294,295,296]. Flagellin, a key component of bacterial flagella, is a natural agonist of TLR5 and another mucosal adjuvant. After binding to TLR5, flagellin activates the MyD88-dependent pathway, which mobilizes nuclear factor NF-κB and stimulates tumor necrosis factor (TNF)-α production [297,298]. Flagellin is also recognized by two NOD-like receptors (NLD), NLRC4 and NAIP5, to activate the inflammasome response [103,299]. Flagellin is a potent adjuvant that can stimulate both innate and adaptive immune responses as it activates proinflammatory cytokine and chemokine production in various immune cells in the mucosa, including DCs, natural killer cells (NKs), epithelial cells, and lymph node stromal cells [103,300].
Cytokines and chemokines, including IL-6, IL-12, and IL-15, help induce CTL responses and antigen-specific IgA antibodies at mucosal sites, while cytokines from the IL-1 family induce both IgA and IgG [256,301]. Other cytokines and chemokines, such as IFN-γ, IL-2, IL-18, IL-21, and granulocyte-macrophage colony-stimulating factor (GM-CSF), exhibit adjuvant activity at mucosal sites [103,302]. However, dose-related toxicities of those cytokines and chemokines, as well as concerns about the stabilities and efficient delivery methods, hinder their clinical application [302,303].
Several natural-product-based mucosal vaccine adjuvant candidates have been studied and may have advantages over conventional adjuvants [304] as they promote specific and non-specific immunity in the host with lower toxicity and side effects due to their biocompatibility and biodegradability [304,305,306]. These adjuvants include biopolymers, nanoparticles, virus-like particles, and extracellular vesicles [304]. N-dihydrogalactochitosan (GC), a novel biopolymer synthesized from galactose and chitosan, has mucoadhesive, pH-resistant, and biocompatible properties. GC can induce type I interferon production and enhance humoral and cellular immunity against SARS-CoV-2. Nanoparticles, such as ferritin and virus-like particles (VLPs), can encapsulate or display antigens in combination with stimulatory molecules on their surfaces with adjuvant activities. The nanoparticles can protect antigens and adjuvant compounds from degradation, increase antigen uptake by mucosal cells, and target specific immune cells and tissues to further enhance mucosal immunity [304].
Mucosal vaccine adjuvants are an active area of research, as mucosal vaccines made using the adjuvants offer many advantages over conventional parenteral vaccines. However, mucosal vaccine adjuvants also face challenges, such as safety, toxicity, stability, and regulatory approval [103,104]. The development of mucosal vaccine adjuvants is at an early stage, and more research is required before they become widely employed in licensed vaccines.

7. Challenges in Mucosal Vaccine Development

Mucosal vaccines promise to elicit protective immunity against infection with substantial advantages over traditional parenteral vaccines. While the development of new mucosal vaccines is highly desirable, more work is needed as there are still challenges in this field. Identification of the correlates of protection is one of the major concerns. This includes a better understanding of mucosal immunity and the key components that contribute to durable and broadly effective protection at mucosal sites after natural infections and vaccinations. Also needed is an improved understanding of the underlying molecular and cellular mechanisms responsible for optimal mucosal immunity, which will facilitate preclinical research, clinical development, and regulatory approval. Important efforts would include basic research aimed at revealing the specific immune cells critical for protective mucosal immunity and the underlying molecular mechanisms that lead to the induction of immunity at mucosal surfaces [47,307,308,309]. Improved understanding is also needed of the host mucosal microenvironment, including the mucosal microbiome, that can affect biophysiological features of the mucosal barrier to help regulate immune responses [310,311,312].
Improved animal models will also be needed to further the development of mucosal vaccines. Existing, widely used animal models, particularly small animal models, may not be adequate and may not correctly predict the activity of mucosal vaccines in humans or other larger animals, such as agriculturally important animals. Beyond simple differences in the sizes of the organs and systems with mucosal surfaces and the fact that several commonly used small animal models are not native hosts for diseases that would be the targets for mucosal vaccine development, the mucosa structures and those associated immune systems in animal gastrointestinal, respiratory, and urogenital tracts vary from those in humans and other larger animals [45]. For example, although mouse models have been widely used for influenza vaccine development, there are limitations since the species is not the natural host, and the sialic acid receptors that mediate viral entry and their distribution in the respiratory tract are different from those in humans. Ferrets are considered more suitable models since they are more biologically similar to human beings in being able to be infected with human influenza virus isolates, exhibit pathogenicity that is closer to that experienced by humans, have more homologous sialic acid receptor distribution, can be infected via airborne transmission, and will mount protective immune responses after vaccination [313,314]. Besides the natural features of the model species, the immune responses elicited by vaccination and by pathogen challenge also need to be validated in comparison to human natural infection or vaccination [313,314,315]. In general, more predictive animal models are needed to help develop new safe and effective mucosal vaccines for a large variety of pathogens.
Determining the ideal vaccine formulation and delivery system is another major hurdle for mucosal vaccine development. Efforts include defining vaccine formulations based on the type of pathogen and the desired immune responses, especially immune responses that correlate well with protection [42,45,47,104]. Determination of dosage and timing are also important [45,47]. Finding better, safe, and effective adjuvants is another critical aspect [103,104]. The methods and devices for vaccine administration are additional areas needing development. Factors such as manufacturing, supply chain, stability, acceptability, and cost are key practical considerations for mucosal vaccine development [45]. Other challenges for mucosal vaccine development include conducting clinical trials that are both acceptable to regulatory authorities internationally and appropriate for at-risk patient populations [45,47].
Mucosal vaccination, beyond the currently approved vaccines, holds great promise for the future. While substantial challenges still hamper the development of safe and effective mucosal vaccines, researchers are continuing to create exciting new technologies to enable the development of future mucosal vaccines.

8. Disadvantages and Side Effects of Mucosal Vaccines

Currently, licensed mucosal vaccines can produce adverse effects. The live-attenuated influenza intranasal vaccine has caused some rare incidences of asthma exacerbation [316]. This has been a concern for parenterally administered live attenuated influenza vaccines as well, but they are now generally considered safe for patients with asthma [317]. More common adverse effects for some of the oral vaccines include fatigue, headaches, and gastrointestinal symptoms [318,319], but such adverse effects are common for many vaccines or even many therapeutics. Some rare cases of anaphylaxis and Guillain–Barré syndrome have also been reported for adenovirus/adenovirus-vectored vaccine [320]. None of these side effects are limited to mucosal vaccines, and similar adverse effects have been caused by various parenteral vaccines [321,322].
Live attenuated vaccines, which constitute the most common mucosal vaccines, come with an inherent risk of reversion to the pathogenic wild type [16]. Vaccine-associated poliomyelitis is contracted at a rate of 4.7 cases per million births by both vaccinees and their contacts [323].
Intranasal vaccines, in particular, must be rigorously tested for safety since the proximity of the nasal passages to the olfactory bulbs may allow access to the central nervous system. This is more of a concern for living or quasi-living (viral-vectored) intranasal vaccines. One study showed that adenovirus vectors are able to infect the CNS when administered intranasally [324]. There are concerns that live attenuated intranasal vaccines, particularly influenza, could cause encephalitis [28]. However, this has not been proven to be a problem [316].
Inactivated or subunit mucosal vaccines may be safer than live vaccines, but they usually require the addition of an adjuvant [324]. These adjuvants, especially enterotoxins, have safety concerns [104]. For example, a cholera toxin adjuvant in an inactivated influenza vaccine increased the risk of Bell’s palsy by almost twenty times when the vaccine was administered intranasally [325,326]. For oral vaccines, it remains a challenge to find adjuvants that are potent enough to stimulate an immune response and survive passage through the upper gastrointestinal tract while minimizing gastrointestinal symptoms [327].
One of the biggest disadvantages of oral vaccine development is the potential for tolerance to the antigen. Chronic environmental enteropathy and deficiencies in vitamin A or zinc, which are more likely to occur in developing countries, significantly decrease the immune response to oral vaccines [52,328]. Additionally, repeated doses of an antigen given orally can result in immune tolerance [329,330,331]. This must be taken into account when considering the number of boosts an oral vaccine needs in order to avoid promoting tolerance to the pathogen.

9. Conclusions and Perspectives

Mucosal vaccines hold great promise to prevent infection and transmission of pathogens, most of which enter hosts across mucosal surfaces. Mucosal vaccines can elicit local immunity at mucosal sites while also inducing strong and long-lasting systemic immune responses. Many mucosal vaccines for humans and animals have been developed and proved effective in blocking infection and transmission of pathogens. Nevertheless, there are still drawbacks to current mucosal vaccines and challenges for new mucosal vaccine development. The importance of rapid production of effective, safe, cost-efficient, distribution-friendly, easily administered mucosal vaccines manufactured at large scale in locations critically affected by disease has become increasingly obvious, particularly given recent experience with COVID-19. Dedicated efforts to advance the mucosal vaccine field and lower the barrier of mucosal vaccine development are essential to address current societal needs and to prepare for the next potential pandemic.

Author Contributions

Conceptualization, Y.S., F.M. and S.L.Z.; writing—original draft preparation, Y.S. and F.M.; writing—review and editing, Y.S., F.M. and S.L.Z.; supervision, project administration, and funding acquisition, S.L.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The study was supported by internal funding from the University of Virginia, including the Manning Fund for COVID-19 Research at UVA, the Ivy Foundation, the Pendleton Laboratory Fund for Pediatric Infectious Disease Research, the Coulter Foundation, and NIAID, NIH (R01 AI176515), and the Commonwealth Commercialization Fund, Virginia Innovation Partnership Corporation.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Figure 1 and Figure 2 are created with BioRender.com.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Woodland, D.L. Vaccine Development. Viral Immunol. 2017, 30, 141. [Google Scholar] [CrossRef]
  2. Alivisatos, C.N. The First Immunologist, James Pylarino (1659–1718), and the Introduction of Variolation: (Section of the History of Medicine). Proc. R. Soc. Med. 1934, 27, 1099–1104. [Google Scholar]
  3. Boylston, A. The Origins of Inoculation. J. R. Soc. Med. 2012, 105, 309–313. [Google Scholar] [CrossRef]
  4. Dinc, G.; Ulman, Y.I. The Introduction of Variolation “A La Turca” to the West by Lady Mary Montagu and Turkey’s Contribution to This. Vaccine 2007, 25, 4261–4265. [Google Scholar] [CrossRef] [PubMed]
  5. Leung, A.K.C. “Variolation” and Vaccination in Late Imperial China, Ca 1570–1911. In History of Vaccine Development; Plotkin, S.A., Ed.; Springer: New York, NY, USA, 2011; pp. 5–12. ISBN 978-1-4419-1338-8. [Google Scholar]
  6. Cantey, J.B. Smallpox Variolation during the Revolutionary War. Pediatr. Infect. Dis. J. 2011, 30, 821. [Google Scholar] [CrossRef] [PubMed]
  7. Jenson, A.B.; Ghim, S.-J.; Sundberg, J.P. An Inquiry into the Causes and Effects of the Variolae (or Cow-Pox. 1798). Exp. Dermatol. 2016, 25, 178–180. [Google Scholar] [CrossRef] [PubMed]
  8. Jenner, E. An Inquiry into the Causes and Effects of the Variolae Vaccinae: A Disease Discovered in Some of the Western Counties of England, Particularly Gloucestershire, and Known by the Name of the Cow Pox; Ashley & Brewer: Springfield, MA, USA, 1798. [Google Scholar]
  9. Riedel, S. Edward Jenner and the History of Smallpox and Vaccination. Proc. Bayl. Univ. Med. Cent. 2005, 18, 21–25. [Google Scholar] [CrossRef] [PubMed]
  10. Plotkin, S.A.; Plotkin, S.L. The Development of Vaccines: How the Past Led to the Future. Nat. Rev. Microbiol. 2011, 9, 889–893. [Google Scholar] [CrossRef] [PubMed]
  11. Plotkin, S.A. Vaccines: Past, Present and Future. Nat. Med. 2005, 11, S5–S11. [Google Scholar] [CrossRef]
  12. Kayser, V.; Ramzan, I. Vaccines and Vaccination: History and Emerging Issues. Hum. Vaccin. Immunother. 2021, 17, 5255–5268. [Google Scholar] [CrossRef]
  13. Zhang, L.; Wang, W.; Wang, S. Effect of Vaccine Administration Modality on Immunogenicity and Efficacy. Expert Rev. Vaccines 2015, 14, 1509–1523. [Google Scholar] [CrossRef] [PubMed]
  14. Kiyono, H.; Yuki, Y.; Nakahashi-Ouchida, R.; Fujihashi, K. Mucosal Vaccines: Wisdom from Now and Then. Int. Immunol. 2021, 33, 767–774. [Google Scholar] [CrossRef] [PubMed]
  15. Yeh, M.T.; Bujaki, E.; Dolan, P.T.; Smith, M.; Wahid, R.; Konz, J.; Weiner, A.J.; Bandyopadhyay, A.S.; Van Damme, P.; De Coster, I.; et al. Engineering the Live-Attenuated Polio Vaccine to Prevent Reversion to Virulence. Cell Host Microbe 2020, 27, 736–751.e8. [Google Scholar] [CrossRef] [PubMed]
  16. Bandyopadhyay, A.S.; Garon, J.; Seib, K.; Orenstein, W.A. Polio Vaccination: Past, Present and Future. Future Microbiol. 2015, 10, 791–808. [Google Scholar] [CrossRef] [PubMed]
  17. Glass, R.I.; Tate, J.E.; Jiang, B.; Parashar, U. The Rotavirus Vaccine Story: From Discovery to the Eventual Control of Rotavirus Disease. J. Infect. Dis. 2021, 224, S331–S342. [Google Scholar] [CrossRef]
  18. Jhaveri, R. Live Attenuated Influenza Vaccine: Is Past Performance a Guarantee of Future Results? Clin. Ther. 2018, 40, 1246–1254. [Google Scholar] [CrossRef]
  19. Tanner, A.R.; Dorey, R.B.; Brendish, N.J.; Clark, T.W. Influenza Vaccination: Protecting the Most Vulnerable. Eur. Respir. Rev. 2021, 30, 200258. [Google Scholar] [CrossRef]
  20. Bi, Q.; Ferreras, E.; Pezzoli, L.; Legros, D.; Ivers, L.C.; Date, K.; Qadri, F.; Digilio, L.; Sack, D.A.; Ali, M.; et al. Protection against Cholera from Killed Whole-Cell Oral Cholera Vaccines: A Systematic Review and Meta-Analysis. Lancet Infect. Dis. 2017, 17, 1080–1088. [Google Scholar] [CrossRef]
  21. Ali, M.; Clemens, J. Assessing Vaccine Herd Protection by Killed Whole-Cell Oral Cholera Vaccines Using Different Study Designs. Front. Public. Health 2019, 7, 211. [Google Scholar] [CrossRef] [PubMed]
  22. Huang, M.; Zhang, M.; Zhu, H.; Du, X.; Wang, J. Mucosal Vaccine Delivery: A Focus on the Breakthrough of Specific Barriers. Acta Pharm. Sin. B 2022, 12, 3456–3474. [Google Scholar] [CrossRef] [PubMed]
  23. Anggraeni, R.; Ana, I.D.; Wihadmadyatami, H. Development of Mucosal Vaccine Delivery: An Overview on the Mucosal Vaccines and Their Adjuvants. Clin. Exp. Vaccine Res. 2022, 11, 235–248. [Google Scholar] [CrossRef]
  24. Hellfritzsch, M.; Scherließ, R. Mucosal Vaccination via the Respiratory Tract. Pharmaceutics 2019, 11, 375. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, T.; Hashizume, T.; Kurita-Ochiai, T.; Yamamoto, M. Sublingual Vaccination with Outer Membrane Protein of Porphyromonas Gingivalis and Flt3 Ligand Elicits Protective Immunity in the Oral Cavity. Biochem. Biophys. Res. Commun. 2009, 390, 937–941. [Google Scholar] [CrossRef] [PubMed]
  26. Freestone, D.S. Clinical Trials Carried out to Assess Non-Parenteral Routes for Administration of Wistar RA 27/3 Strain Live Attenuated Rubella Vaccine. Dev. Biol. Stand. 1976, 33, 237–240. [Google Scholar] [PubMed]
  27. Gala, R.P.; Popescu, C.; Knipp, G.T.; McCain, R.R.; Ubale, R.V.; Addo, R.; Bhowmik, T.; Kulczar, C.D.; D’Souza, M.J. Physicochemical and Preclinical Evaluation of a Novel Buccal Measles Vaccine. AAPS PharmSciTech 2017, 18, 283–292. [Google Scholar] [CrossRef]
  28. 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]
  29. Waerlop, G.; Janssens, Y.; Jacobs, B.; Jarczowski, F.; Diessner, A.; Leroux-Roels, G.; Klimyuk, V.; Leroux-Roels, I.; Thieme, F. Immune Responses in Healthy Adults Elicited by a Bivalent Norovirus Vaccine Candidate Composed of GI.4 and GII.4 VLPs without Adjuvant. Front. Immunol. 2023, 14, 1188431. [Google Scholar] [CrossRef]
  30. Lu, B.; Huang, Y.; Huang, L.; Li, B.; Zheng, Z.; Chen, Z.; Chen, J.; Hu, Q.; Wang, H. Effect of Mucosal and Systemic Immunization with Virus-like Particles of Severe Acute Respiratory Syndrome Coronavirus in Mice. Immunology 2010, 130, 254–261. [Google Scholar] [CrossRef] [PubMed]
  31. Shuttleworth, G.; Eckery, D.C.; Awram, P. Oral and Intraperitoneal Immunization with Rotavirus 2/6 Virus-like Particles Stimulates a Systemic and Mucosal Immune Response in Mice. Arch. Virol. 2005, 150, 341–349. [Google Scholar] [CrossRef]
  32. Yan, R.; Wang, J.; Xu, L.; Song, X.; Li, X. DNA Vaccine Encoding Haemonchus Contortus Actin Induces Partial Protection in Goats. Acta Parasitol. 2014, 59, 698–709. [Google Scholar] [CrossRef]
  33. Jia, R.; Guo, J.H.; Fan, M.W.; Bian, Z.; Chen, Z.; Fan, B.; Yu, F.; Xu, Q.A. Immunogenicity of CTLA4 Fusion Anti-Caries DNA Vaccine in Rabbits and Monkeys. Vaccine 2006, 24, 5192–5200. [Google Scholar] [CrossRef] [PubMed]
  34. Zhao, G.; Yan, R.; Muleke, C.I.; Sun, Y.; Xu, L.; Li, X. Vaccination of Goats with DNA Vaccines Encoding H11 and IL-2 Induces Partial Protection against Haemonchus contortus Infection. Vet. J. 2012, 191, 94–100. [Google Scholar] [CrossRef] [PubMed]
  35. Su, F.; Patel, G.B.; Hu, S.; Chen, W. Induction of Mucosal Immunity through Systemic Immunization: Phantom or Reality? Hum. Vaccin. Immunother. 2016, 12, 1070–1079. [Google Scholar] [CrossRef] [PubMed]
  36. Atmar, R.L.; Keitel, W.A.; Cate, T.R.; Munoz, F.M.; Ruben, F.; Couch, R.B. A Dose–Response Evaluation of Inactivated Influenza Vaccine given Intranasally and Intramuscularly to Healthy Young Adults. Vaccine 2007, 25, 5367–5373. [Google Scholar] [CrossRef] [PubMed]
  37. Ogra, P.L.; Faden, H.; Welliver, R.C. Vaccination Strategies for Mucosal Immune Responses. Clin. Microbiol. Rev. 2001, 14, 430–445. [Google Scholar] [CrossRef] [PubMed]
  38. Tsai, C.J.Y.; Loh, J.M.S.; Fujihashi, K.; Kiyono, H. Mucosal Vaccination: Onward and Upward. Expert. Rev. Vaccines 2023, 22, 885–899. [Google Scholar] [CrossRef]
  39. Van Der Ley, P.A.; Zariri, A.; Van Riet, E.; Oosterhoff, D.; Kruiswijk, C.P. An Intranasal OMV-Based Vaccine Induces High Mucosal and Systemic Protecting Immunity Against a SARS-CoV-2 Infection. Front. Immunol. 2021, 12, 781280. [Google Scholar] [CrossRef]
  40. Hoft, D.F.; Xia, M.; Zhang, G.L.; Blazevic, A.; Tennant, J.; Kaplan, C.; Matuschak, G.; Dube, T.J.; Hill, H.; Schlesinger, L.S.; et al. PO and ID BCG Vaccination in Humans Induce Distinct Mucosal and Systemic Immune Responses and CD4+ T Cell Transcriptomal Molecular Signatures. Mucosal Immunol. 2018, 11, 486–495. [Google Scholar] [CrossRef]
  41. Miquel-Clopés, A.; Bentley, E.G.; Stewart, J.P.; Carding, S.R. Mucosal Vaccines and Technology. Clin. Exp. Immunol. 2019, 196, 205–214. [Google Scholar] [CrossRef]
  42. Neutra, M.R.; Kozlowski, P.A. Mucosal Vaccines: The Promise and the Challenge. Nat. Rev. Immunol. 2006, 6, 148–158. [Google Scholar] [CrossRef]
  43. Woodrow, K.A.; Bennett, K.M.; Lo, D.D. Mucosal Vaccine Design and Delivery. Annu. Rev. Biomed. Eng. 2012, 14, 17–46. [Google Scholar] [CrossRef] [PubMed]
  44. Hoft, D.F.; Brusic, V.; Sakala, I.G. Optimizing Vaccine Development. Cell. Microbiol. 2011, 13, 934–942. [Google Scholar] [CrossRef]
  45. Knisely, J.M.; Buyon, L.E.; Mandt, R.; Farkas, R.; Balasingam, S.; Bok, K.; Buchholz, U.J.; D’Souza, M.P.; Gordon, J.L.; King, D.F.L.; et al. Mucosal Vaccines for SARS-CoV-2: Scientific Gaps and Opportunities-Workshop Report. NPJ Vaccines 2023, 8, 53. [Google Scholar] [CrossRef] [PubMed]
  46. Morens, D.M.; Taubenberger, J.K.; Fauci, A.S. Rethinking Next-Generation Vaccines for Coronaviruses, Influenzaviruses, and Other Respiratory Viruses. Cell Host Microbe 2023, 31, 146–157. [Google Scholar] [CrossRef] [PubMed]
  47. Lavelle, E.C.; Ward, R.W. Mucosal Vaccines—Fortifying the Frontiers. Nat. Rev. Immunol. 2022, 22, 236–250. [Google Scholar] [CrossRef] [PubMed]
  48. Miteva, D.; Peshevska-Sekulovska, M.; Snegarova, V.; Batselova, H.; Alexandrova, R.; Velikova, T. Mucosal COVID-19 Vaccines: Risks, Benefits and Control of the Pandemic. World J. Virol. 2022, 11, 221–236. [Google Scholar] [CrossRef] [PubMed]
  49. Freeman, D.; Lambe, S.; Yu, L.-M.; Freeman, J.; Chadwick, A.; Vaccari, C.; Waite, F.; Rosebrock, L.; Petit, A.; Vanderslott, S.; et al. Injection Fears and COVID-19 Vaccine Hesitancy. Psychol. Med. 2023, 53, 1185–1195. [Google Scholar] [CrossRef] [PubMed]
  50. Taddio, A.; Ipp, M.; Thivakaran, S.; Jamal, A.; Parikh, C.; Smart, S.; Sovran, J.; Stephens, D.; Katz, J. Survey of the Prevalence of Immunization Non-Compliance Due to Needle Fears in Children and Adults. Vaccine 2012, 30, 4807–4812. [Google Scholar] [CrossRef]
  51. Çuburu, N.; Kweon, M.-N.; Hervouet, C.; Cha, H.-R.; Pang, Y.-Y.S.; Holmgren, J.; Stadler, K.; Schiller, J.T.; Anjuère, F.; Czerkinsky, C. Sublingual Immunization with Nonreplicating Antigens Induces Antibody-Forming Cells and Cytotoxic T Cells in the Female Genital Tract Mucosa and Protects against Genital Papillomavirus Infection. J. Immunol. 2009, 183, 7851–7859. [Google Scholar] [CrossRef]
  52. Czerkinsky, C.; Holmgren, J. Mucosal Delivery Routes for Optimal Immunization: Targeting Immunity to the Right Tissues. In Mucosal Vaccines; Kozlowski, P.A., Ed.; Current Topics in Microbiology and Immunology; Springer: Berlin/Heidelberg, Germany, 2010; Volume 354, pp. 1–18. ISBN 978-3-642-23692-1. [Google Scholar]
  53. Smith, D.R.; Leggat, P.A. Pioneering Figures in Medicine: Albert Bruce Sabin—Inventor of the Oral Polio Vaccine. Kurume Med. J. 2005, 52, 111–116. [Google Scholar] [CrossRef]
  54. Kehagia, E.; Papakyriakopoulou, P.; Valsami, G. Advances in Intranasal Vaccine Delivery: A Promising Non-Invasive Route of Immunization. Vaccine 2023, 41, 3589–3603. [Google Scholar] [CrossRef]
  55. Zhou, B.; Meliopoulos, V.A.; Wang, W.; Lin, X.; Stucker, K.M.; Halpin, R.A.; Stockwell, T.B.; Schultz-Cherry, S.; Wentworth, D.E. Reversion of Cold-Adapted Live Attenuated Influenza Vaccine into a Pathogenic Virus. J. Virol. 2016, 90, 8454–8463. [Google Scholar] [CrossRef]
  56. Murphy, B.R.; Coelingh, K. Principles Underlying the Development and Use of Live Attenuated Cold-Adapted Influenza A and B Virus Vaccines. Viral Immunol. 2002, 15, 295–323. [Google Scholar] [CrossRef] [PubMed]
  57. Ilyushina, N.A.; Haynes, B.C.; Hoen, A.G.; Khalenkov, A.M.; Housman, M.L.; Brown, E.P.; Ackerman, M.E.; Treanor, J.J.; Luke, C.J.; Subbarao, K.; et al. Live Attenuated and Inactivated Influenza Vaccines in Children. J. Infect. Dis. 2015, 211, 352–360. [Google Scholar] [CrossRef] [PubMed]
  58. Slütter, B.; Pewe, L.L.; Lauer, P.; Harty, J.T. Cutting Edge: Rapid Boosting of Cross-Reactive Memory CD8 T Cells Broadens the Protective Capacity of the Flumist Vaccine. J. Immunol. 2013, 190, 3854–3858. [Google Scholar] [CrossRef]
  59. Zens, K.D.; Chen, J.K.; Farber, D.L. Vaccine-Generated Lung Tissue–Resident Memory T Cells Provide Heterosubtypic Protection to Influenza Infection. JCI Insight 2016, 1, e85832. [Google Scholar] [CrossRef]
  60. Gallichan, W.S.; Rosenthal, K.L. Long-Lived Cytotoxic T Lymphocyte Memory in Mucosal Tissues after Mucosal but Not Systemic Immunization. J. Exp. Med. 1996, 184, 1879–1890. [Google Scholar] [CrossRef] [PubMed]
  61. Pilapitiya, D.; Wheatley, A.K.; Tan, H.-X. Mucosal Vaccines for SARS-CoV-2: Triumph of Hope over Experience. eBioMedicine 2023, 92, 104585. [Google Scholar] [CrossRef]
  62. Trondsen, M.; Aqrawi, L.A.; Zhou, F.; Pedersen, G.; Trieu, M.C.; Zhou, P.; Cox, R.J. Induction of Local Secretory IgA and Multifunctional CD 4 + T-helper Cells Following Intranasal Immunization with a H5N1 Whole Inactivated Influenza Virus Vaccine in BALB/c Mice. Scand. J. Immunol. 2015, 81, 305–317. [Google Scholar] [CrossRef]
  63. Odevall, L.; Hong, D.; Digilio, L.; Sahastrabuddhe, S.; Mogasale, V.; Baik, Y.; Choi, S.; Kim, J.H.; Lynch, J. The Euvichol Story—Development and Licensure of a Safe, Effective and Affordable Oral Cholera Vaccine through Global Public Private Partnerships. Vaccine 2018, 36, 6606–6614. [Google Scholar] [CrossRef]
  64. Shaikh, H.; Lynch, J.; Kim, J.; Excler, J.-L. Current and Future Cholera Vaccines. Vaccine 2020, 38, A118–A126. [Google Scholar] [CrossRef]
  65. Lycke, N. Recent Progress in Mucosal Vaccine Development: Potential and Limitations. Nat. Rev. Immunol. 2012, 12, 592–605. [Google Scholar] [CrossRef] [PubMed]
  66. Chen, W.H.; Cohen, M.B.; Kirkpatrick, B.D.; Brady, R.C.; Galloway, D.; Gurwith, M.; Hall, R.H.; Kessler, R.A.; Lock, M.; Haney, D.; et al. Single-Dose Live Oral Cholera Vaccine CVD 103-HgR Protects against Human Experimental Infection with Vibrio cholerae O1 El Tor. Clin. Infect. Dis. 2016, 62, 1329–1335. [Google Scholar] [CrossRef]
  67. McCarty, J.M.; Lock, M.D.; Hunt, K.M.; Simon, J.K.; Gurwith, M. Safety and Immunogenicity of Single-Dose Live Oral Cholera Vaccine Strain CVD 103-HgR in Healthy Adults Age 18–45. Vaccine 2018, 36, 833–840. [Google Scholar] [CrossRef] [PubMed]
  68. Saluja, T.; Mogasale, V.V.; Excler, J.-L.; Kim, J.H.; Mogasale, V. An Overview of VaxchoraTM, a Live Attenuated Oral Cholera Vaccine. Hum. Vaccines Immunother. 2020, 16, 42–50. [Google Scholar] [CrossRef] [PubMed]
  69. Jahnmatz, M.; Richert, L.; al-Tawil, N.; Storsaeter, J.; Colin, C.; Bauduin, C.; Thalen, M.; Solovay, K.; Rubin, K.; Mielcarek, N.; et al. Safety and Immunogenicity of the Live Attenuated Intranasal Pertussis Vaccine BPZE1: A Phase 1b, Double-Blind, Randomised, Placebo-Controlled Dose-Escalation Study. Lancet Infect. Dis. 2020, 20, 1290–1301. [Google Scholar] [CrossRef]
  70. Shin, S.; Desai, S.N.; Sah, B.K.; Clemens, J.D. Oral Vaccines against Cholera. Clin. Infect. Dis. 2011, 52, 1343–1349. [Google Scholar] [CrossRef] [PubMed]
  71. Beltrán-Beck, B.; De La Fuente, J.; Garrido, J.M.; Aranaz, A.; Sevilla, I.; Villar, M.; Boadella, M.; Galindo, R.C.; Pérez De La Lastra, J.M.; Moreno-Cid, J.A.; et al. Oral Vaccination with Heat Inactivated Mycobacterium bovis Activates the Complement System to Protect against Tuberculosis. PLoS ONE 2014, 9, e98048. [Google Scholar] [CrossRef]
  72. Sukwa, N.; Mubanga, C.; Hatyoka, L.M.; Chilyabanyama, O.N.; Chibuye, M.; Mundia, S.; Munyinda, M.; Kamuti, E.; Siyambango, M.; Badiozzaman, S.; et al. Safety, Tolerability, and Immunogenicity of an Oral Inactivated ETEC Vaccine (ETVAX®) with dmLT Adjuvant in Healthy Adults and Children in Zambia: An Age Descending Randomised, Placebo-Controlled Trial. Vaccine 2023, 41, 6884–6894. [Google Scholar] [CrossRef]
  73. Walker, R.I.; Bourgeois, A.L. Oral Inactivated Whole Cell Vaccine for Mucosal Immunization: ETVAX Case Study. Front. Immunol. 2023, 14, 1125102. [Google Scholar] [CrossRef]
  74. Sur, D.; Lopez, A.L.; Kanungo, S.; Paisley, A.; Manna, B.; Ali, M.; Niyogi, S.K.; Park, J.K.; Sarkar, B.; Puri, M.K.; et al. Efficacy and Safety of a Modified Killed-Whole-Cell Oral Cholera Vaccine in India: An Interim Analysis of a Cluster-Randomised, Double-Blind, Placebo-Controlled Trial. Lancet 2009, 374, 1694–1702. [Google Scholar] [CrossRef]
  75. Amicizia, D.; Arata, L.; Zangrillo, F.; Panatto, D.; Gasparini, R. Overview of the Impact of Typhoid and Paratyphoid Fever. Utility of Ty21a Vaccine (Vivotif®). J. Prev. Med. Hyg. 2017, 58, E1–E8. [Google Scholar]
  76. US FDA. Vaccines Licensed for Use in the United States. Available online: https://www.fda.gov/vaccines-blood-biologics/vaccines/vaccines-licensed-use-united-states (accessed on 21 January 2024).
  77. Prequalified Vaccines|WHO—Prequalification of Medical Products (IVDs, Medicines, Vaccines and Immunization Devices, Vector Control). Available online: https://extranet.who.int/prequal/vaccines/prequalified-vaccines (accessed on 21 January 2024).
  78. Wilson, H.L.; Gerdts, V.; Babiuk, L.A. Mucosal Vaccine Development for Veterinary and Aquatic Diseases. In Mucosal Vaccines; Elsevier: Amsterdam, The Netherlands, 2020; pp. 811–829. ISBN 978-0-12-811924-2. [Google Scholar]
  79. Fairbrother, J.M.; Nadeau, É.; Bélanger, L.; Tremblay, C.-L.; Tremblay, D.; Brunelle, M.; Wolf, R.; Hellmann, K.; Hidalgo, Á. Immunogenicity and Protective Efficacy of a Single-Dose Live Non-Pathogenic Escherichia coli Oral Vaccine against F4-Positive Enterotoxigenic Escherichia Coli Challenge in Pigs. Vaccine 2017, 35, 353–360. [Google Scholar] [CrossRef]
  80. Curtiss, R.; Hassan, J.O. Nonrecombinant and Recombinant Avirulent Salmonella Vaccines for Poultry. Vet. Immunol. Immunopathol. 1996, 54, 365–372. [Google Scholar] [CrossRef] [PubMed]
  81. Hassan, J.O.; Curtiss, R. Development and Evaluation of an Experimental Vaccination Program Using a Live Avirulent Salmonella Typhimurium Strain to Protect Immunized Chickens against Challenge with Homologous and Heterologous Salmonella Serotypes. Infect. Immun. 1994, 62, 5519–5527. [Google Scholar] [CrossRef] [PubMed]
  82. Atterbury, R.J.; Morris, V.; Harrison, D.; Tucker, V.; Allen, V.M.; Davies, R.H.; Carrique-Mas, J.J. Effect of Delivery Method on the Efficacy of Salmonella Vaccination in Chickens. Vet. Rec. 2010, 167, 161–164. [Google Scholar] [CrossRef] [PubMed]
  83. Koeppel, K.N.; Geertsma, P.; Kuhn, B.F.; Van Schalkwyk, O.L.; Thompson, P.N. Antibody Response to Raboral VR-G® Oral Rabies Vaccine in Captive and Free-Ranging Black-Backed Jackals (Canis mesomelas). Onderstepoort J. Vet. Res. 2022, 89, a1975. [Google Scholar] [CrossRef] [PubMed]
  84. Johnson, S.R.; Slate, D.; Nelson, K.M.; Davis, A.J.; Mills, S.A.; Forbes, J.T.; VerCauteren, K.C.; Gilbert, A.T.; Chipman, R.B. Serological Responses of Raccoons and Striped Skunks to Ontario Rabies Vaccine Bait in West Virginia during 2012–2016. Viruses 2021, 13, 157. [Google Scholar] [CrossRef]
  85. Afkhami, S.; D’Agostino, M.R.; Vaseghi-Shanjani, M.; Lepard, M.; Yang, J.X.; Lai, R.; Choi, M.W.Y.; Chacon, A.; Zganiacz, A.; Franken, K.L.M.C.; et al. Intranasal Multivalent Adenoviral-Vectored Vaccine Protects against Replicating and Dormant M.Tb in Conventional and Humanized Mice. npj Vaccines 2023, 8, 25. [Google Scholar] [CrossRef]
  86. Wang, S.; Liang, B.; Wang, W.; Li, L.; Feng, N.; Zhao, Y.; Wang, T.; Yan, F.; Yang, S.; Xia, X. Viral Vectored Vaccines: Design, Development, Preventive and Therapeutic Applications in Human Diseases. Signal Transduct. Target. Ther. 2023, 8, 149. [Google Scholar] [CrossRef]
  87. Zhu, F.; Zhuang, C.; Chu, K.; Zhang, L.; Zhao, H.; Huang, S.; Su, Y.; Lin, H.; Yang, C.; Jiang, H.; et al. Safety and Immunogenicity of a Live-Attenuated Influenza Virus Vector-Based Intranasal SARS-CoV-2 Vaccine in Adults: Randomised, Double-Blind, Placebo-Controlled, Phase 1 and 2 Trials. Lancet Respir. Med. 2022, 10, 749–760. [Google Scholar] [CrossRef] [PubMed]
  88. Wang, D.; Song, J.; Wang, J.; Zhang, Z.; Shen, Q.; Wang, J.; Guo, H.; Wang, R.; Xie, C.; Lu, W.; et al. Efficient Gene Delivery Based on Guanidyl-Nucleic Acid Molecular Interactions. Adv. Funct. Mater. 2020, 30, 2004783. [Google Scholar] [CrossRef]
  89. Zhang, H.; Liu, Z.; Lihe, H.; Lu, L.; Zhang, Z.; Yang, S.; Meng, N.; Xiong, Y.; Fan, X.; Chen, Z.; et al. Intranasal G5-BGG/pDNA Vaccine Elicits Protective Systemic and Mucosal Immunity against SARS-CoV-2 by Transfecting Mucosal Dendritic Cells. Adv. Healthc. Mater. 2023, 2303261. [Google Scholar] [CrossRef] [PubMed]
  90. Kim, H.; Kirtane, A.R.; Kim, N.Y.; Rajesh, N.U.; Tang, C.; Ishida, K.; Hayward, A.M.; Langer, R.; Traverso, G. Gastrointestinal Delivery of an mRNA Vaccine Using Immunostimulatory Polymeric Nanoparticles. AAPS J. 2023, 25, 81. [Google Scholar] [CrossRef] [PubMed]
  91. Dénes, B.; Fuller, R.; Kelin, W.; Levin, T.; Gil, J.; Harewood, A.; Lőricz, M.; Wall, N.; Firek, A.; Langridge, W. A CTB-SARS-CoV-2-ACE-2 RBD Mucosal Vaccine Protects Against Coronavirus Infection. Vaccines 2023, 11, 1865. [Google Scholar] [CrossRef] [PubMed]
  92. Wolf, M.A.; Boehm, D.T.; DeJong, M.A.; Wong, T.Y.; Sen-Kilic, E.; Hall, J.M.; Blackwood, C.B.; Weaver, K.L.; Kelly, C.O.; Kisamore, C.A.; et al. Intranasal Immunization with Acellular Pertussis Vaccines Results in Long-Term Immunity to Bordetella pertussis in Mice. Infect. Immun. 2021, 89, e00607-20. [Google Scholar] [CrossRef]
  93. Jiang, W.; Wang, X.; Su, Y.; Cai, L.; Li, J.; Liang, J.; Gu, Q.; Sun, M.; Shi, L. Intranasal Immunization with a C-Di-GMP-Adjuvanted Acellular Pertussis Vaccine Provides Superior Immunity against Bordetella pertussis in a Mouse Model. Front. Immunol. 2022, 13, 878832. [Google Scholar] [CrossRef]
  94. Yount, K.S.; Hall, J.M.; Caution, K.; Shamseldin, M.M.; Guo, M.; Marion, K.; Fullen, A.R.; Huang, Y.; Maynard, J.A.; Quataert, S.A.; et al. Systemic Priming and Intranasal Booster with a BcfA-Adjuvanted Acellular Pertussis Vaccine Generates CD4+ IL-17+ Nasal Tissue Resident T Cells and Reduces B. pertussis Nasal Colonization. Front. Immunol. 2023, 14, 1181876. [Google Scholar] [CrossRef]
  95. Yousefi Avarvand, A.; Meshkat, Z.; Khademi, F.; Aryan, E.; Sankian, M.; Tafaghodi, M. Enhancement of the Immunogenicity of a Mycobacterium tuberculosis Fusion Protein Using ISCOMATRIX and PLUSCOM Nano-Adjuvants as Prophylactic Vaccine after Nasal Administration in Mice. Iran. J. Basic. Med. Sci. 2024, 27, 24–30. [Google Scholar] [CrossRef]
  96. Ning, H.; Zhang, W.; Kang, J.; Ding, T.; Liang, X.; Lu, Y.; Guo, C.; Sun, W.; Wang, H.; Bai, Y.; et al. Subunit Vaccine ESAT-6:C-Di-AMP Delivered by Intranasal Route Elicits Immune Responses and Protects against Mycobacterium tuberculosis Infection. Front. Cell. Infect. Microbiol. 2021, 11, 647220. [Google Scholar] [CrossRef]
  97. Gaspar, E.B.; Prudencio, C.R.; De Gaspari, E. Experimental Studies Using OMV in a New Platform of SARS-CoV-2 Vaccines. Hum. Vaccines Immunother. 2021, 17, 2965–2968. [Google Scholar] [CrossRef] [PubMed]
  98. Hozbor, D.F. Outer Membrane Vesicles: An Attractive Candidate for Pertussis Vaccines. Expert. Rev. Vaccines 2017, 16, 193–196. [Google Scholar] [CrossRef] [PubMed]
  99. Raeven, R.H.M.; Rockx-Brouwer, D.; Kanojia, G.; Van Der Maas, L.; Bindels, T.H.E.; Ten Have, R.; Van Riet, E.; Metz, B.; Kersten, G.F.A. Intranasal Immunization with Outer Membrane Vesicle Pertussis Vaccine Confers Broad Protection through Mucosal IgA and Th17 Responses. Sci. Rep. 2020, 10, 7396. [Google Scholar] [CrossRef] [PubMed]
  100. Keech, C.; Miller, V.E.; Rizzardi, B.; Hoyle, C.; Pryor, M.J.; Ferrand, J.; Solovay, K.; Thalen, M.; Noviello, S.; Goldstein, P.; et al. Immunogenicity and Safety of BPZE1, an Intranasal Live Attenuated Pertussis Vaccine, versus Tetanus–Diphtheria–Acellular Pertussis Vaccine: A Randomised, Double-Blind, Phase 2b Trial. Lancet 2023, 401, 843–855. [Google Scholar] [CrossRef]
  101. Lai, R.; Ogunsola, A.F.; Rakib, T.; Behar, S.M. Key Advances in Vaccine Development for Tuberculosis—Success and Challenges. npj Vaccines 2023, 8, 158. [Google Scholar] [CrossRef] [PubMed]
  102. Maeda, D.L.N.F.; Tian, D.; Yu, H.; Dar, N.; Rajasekaran, V.; Meng, S.; Mahsoub, H.M.; Sooryanarain, H.; Wang, B.; Heffron, C.L.; et al. Killed Whole-Genome Reduced-Bacteria Surface-Expressed Coronavirus Fusion Peptide Vaccines Protect against Disease in a Porcine Model. Proc. Natl. Acad. Sci. USA 2021, 118, e2025622118. [Google Scholar] [CrossRef]
  103. Correa, V.A.; Portilho, A.I.; De Gaspari, E. Vaccines, Adjuvants and Key Factors for Mucosal Immune Response. Immunology 2022, 167, 124–138. [Google Scholar] [CrossRef]
  104. Lawson, L.B.; Norton, E.B.; Clements, J.D. Defending the Mucosa: Adjuvant and Carrier Formulations for Mucosal Immunity. Curr. Opin. Immunol. 2011, 23, 414–420. [Google Scholar] [CrossRef]
  105. Sheng, Y.H.; Hasnain, S.Z. Mucus and Mucins: The Underappreciated Host Defence System. Front. Cell Infect. Microbiol. 2022, 12, 856962. [Google Scholar] [CrossRef]
  106. Neutra, M.R.; Mantis, N.J.; Kraehenbuhl, J.P. Collaboration of Epithelial Cells with Organized Mucosal Lymphoid Tissues. Nat. Immunol. 2001, 2, 1004–1009. [Google Scholar] [CrossRef]
  107. Yang, J.; Yan, H. Mucosal Epithelial Cells: The Initial Sentinels and Responders Controlling and Regulating Immune Responses to Viral Infections. Cell Mol. Immunol. 2021, 18, 1628–1630. [Google Scholar] [CrossRef]
  108. Ross, K.F.; Herzberg, M.C. Autonomous Immunity in Mucosal Epithelial Cells: Fortifying the Barrier against Infection. Microbes Infect. 2016, 18, 387–398. [Google Scholar] [CrossRef]
  109. Hammad, H.; Lambrecht, B.N. Barrier Epithelial Cells and the Control of Type 2 Immunity. Immunity 2015, 43, 29–40. [Google Scholar] [CrossRef]
  110. Bustamante-Marin, X.M.; Ostrowski, L.E. Cilia and Mucociliary Clearance. Cold Spring Harb. Perspect. Biol. 2017, 9, a028241. [Google Scholar] [CrossRef]
  111. Allaire, J.M.; Crowley, S.M.; Law, H.T.; Chang, S.-Y.; Ko, H.-J.; Vallance, B.A. The Intestinal Epithelium: Central Coordinator of Mucosal Immunity. Trends Immunol. 2018, 39, 677–696. [Google Scholar] [CrossRef] [PubMed]
  112. Groeger, S.; Meyle, J. Oral Mucosal Epithelial Cells. Front. Immunol. 2019, 10, 208. [Google Scholar] [CrossRef] [PubMed]
  113. France, M.M.; Turner, J.R. The Mucosal Barrier at a Glance. J. Cell Sci. 2017, 130, 307–314. [Google Scholar] [CrossRef]
  114. Linden, S.K.; Sutton, P.; Karlsson, N.G.; Korolik, V.; McGuckin, M.A. Mucins in the Mucosal Barrier to Infection. Mucosal Immunol. 2008, 1, 183–197. [Google Scholar] [CrossRef] [PubMed]
  115. Uhlig, F.; Hyland, N.P. Making Sense of Quorum Sensing at the Intestinal Mucosal Interface. Cells 2022, 11, 1734. [Google Scholar] [CrossRef] [PubMed]
  116. Benam, K.H.; Denney, L.; Ho, L.-P. How the Respiratory Epithelium Senses and Reacts to Influenza Virus. Am. J. Respir. Cell Mol. Biol. 2019, 60, 259–268. [Google Scholar] [CrossRef] [PubMed]
  117. Liu, S.; Hur, Y.H.; Cai, X.; Cong, Q.; Yang, Y.; Xu, C.; Bilate, A.M.; Gonzales, K.A.U.; Parigi, S.M.; Cowley, C.J.; et al. A Tissue Injury Sensing and Repair Pathway Distinct from Host Pathogen Defense. Cell 2023, 186, 2127–2143.e22. [Google Scholar] [CrossRef]
  118. Hewitt, R.J.; Lloyd, C.M. Regulation of Immune Responses by the Airway Epithelial Cell Landscape. Nat. Rev. Immunol. 2021, 21, 347–362. [Google Scholar] [CrossRef]
  119. Chakraborty, R.; Lo, D.D. Chronic Inflammation in Mucosal Tissues: Barrier Integrity, Inducible Lymphoid Tissues, and Immune Surveillance. Curr. Top. Microbiol. Immunol. 2020, 426, 45–63. [Google Scholar] [CrossRef]
  120. Ali, A.; Tan, H.; Kaiko, G.E. Role of the Intestinal Epithelium and Its Interaction With the Microbiota in Food Allergy. Front. Immunol. 2020, 11, 604054. [Google Scholar] [CrossRef]
  121. McGhee, J.R.; Fujihashi, K. Inside the Mucosal Immune System. PLoS Biol. 2012, 10, e1001397. [Google Scholar] [CrossRef]
  122. Krausgruber, T.; Fortelny, N.; Fife-Gernedl, V.; Senekowitsch, M.; Schuster, L.C.; Lercher, A.; Nemc, A.; Schmidl, C.; Rendeiro, A.F.; Bergthaler, A.; et al. Structural Cells Are Key Regulators of Organ-Specific Immune Responses. Nature 2020, 583, 296–302. [Google Scholar] [CrossRef]
  123. Hendel, S.K.; Kellermann, L.; Hausmann, A.; Bindslev, N.; Jensen, K.B.; Nielsen, O.H. Tuft Cells and Their Role in Intestinal Diseases. Front. Immunol. 2022, 13, 822867. [Google Scholar] [CrossRef]
  124. Olivares-Villagómez, D.; Van Kaer, L. Intestinal Intraepithelial Lymphocytes: Sentinels of the Mucosal Barrier. Trends Immunol. 2018, 39, 264–275. [Google Scholar] [CrossRef]
  125. Hoytema van Konijnenburg, D.P.; Mucida, D. Intraepithelial Lymphocytes. Curr. Biol. 2017, 27, R737–R739. [Google Scholar] [CrossRef]
  126. Buckley, A.; Turner, J.R. Cell Biology of Tight Junction Barrier Regulation and Mucosal Disease. Cold Spring Harb. Perspect. Biol. 2018, 10, a029314. [Google Scholar] [CrossRef]
  127. Sharkey, K.A.; Mawe, G.M. The Enteric Nervous System. Physiol. Rev. 2023, 103, 1487–1564. [Google Scholar] [CrossRef]
  128. Mohammad, N.S.; Nazli, R.; Zafar, H.; Fatima, S. Effects of Lipid Based Multiple Micronutrients Supplement on the Birth Outcome of Underweight Pre-Eclamptic Women: A Randomized Clinical Trial. Pak. J. Med. Sci. 2022, 38, 219–226. [Google Scholar] [CrossRef]
  129. Okumura, R.; Takeda, K. Maintenance of Intestinal Homeostasis by Mucosal Barriers. Inflamm. Regen. 2018, 38, 5. [Google Scholar] [CrossRef] [PubMed]
  130. Patel, H.; Yewale, C.; Rathi, M.N.; Misra, A. Mucosal Immunization: A Review of Strategies and Challenges. Crit. Rev. Ther. Drug Carr. Syst. 2014, 31, 273–303. [Google Scholar] [CrossRef] [PubMed]
  131. An, J.; Liu, Y.; Wang, Y.; Fan, R.; Hu, X.; Zhang, F.; Yang, J.; Chen, J. The Role of Intestinal Mucosal Barrier in Autoimmune Disease: A Potential Target. Front. Immunol. 2022, 13, 871713. [Google Scholar] [CrossRef]
  132. Brandtzaeg, P. Mucosal Immunity: Induction, Dissemination, and Effector Functions. Scand. J. Immunol. 2009, 70, 505–515. [Google Scholar] [CrossRef]
  133. Kabat, A.M.; Pott, J.; Maloy, K.J. The Mucosal Immune System and Its Regulation by Autophagy. Front. Immunol. 2016, 7, 240. [Google Scholar] [CrossRef]
  134. Nagura, H. Mucosal Defense Mechanism in Health and Disease. Role of the Mucosal Immune System. Acta Pathol. Jpn. 1992, 42, 387–400. [Google Scholar] [CrossRef] [PubMed]
  135. Boparai, J.K.; Sharma, P.K. Mini Review on Antimicrobial Peptides, Sources, Mechanism and Recent Applications. Protein Pept. Lett. 2020, 27, 4–16. [Google Scholar] [CrossRef]
  136. Zhang, R.; Xu, L.; Dong, C. Antimicrobial Peptides: An Overview of Their Structure, Function and Mechanism of Action. Protein Pept. Lett. 2022, 29, 641–650. [Google Scholar] [CrossRef]
  137. Przesmycki, F. Current view on the epidemiology of poliomyelitis. Postep. Hig. Med. Dosw. 1966, 20, 833–857. [Google Scholar]
  138. Luo, Y.; Song, Y. Mechanism of Antimicrobial Peptides: Antimicrobial, Anti-Inflammatory and Antibiofilm Activities. Int. J. Mol. Sci. 2021, 22, 11401. [Google Scholar] [CrossRef]
  139. Grond, K.; Kurtz, C.C.; Hatton, J.; Sonsalla, M.M.; Duddleston, K.N. Gut Microbiome Is Affected by Gut Region but Robust to Host Physiological Changes in Captive Active-Season Ground Squirrels. Anim. Microbiome 2021, 3, 56. [Google Scholar] [CrossRef]
  140. Juge, N. Relationship between Mucosa-Associated Gut Microbiota and Human Diseases. Biochem. Soc. Trans. 2022, 50, 1225–1236. [Google Scholar] [CrossRef]
  141. Shi, N.; Li, N.; Duan, X.; Niu, H. Interaction between the Gut Microbiome and Mucosal Immune System. Mil. Med. Res. 2017, 4, 14. [Google Scholar] [CrossRef] [PubMed]
  142. Zhang, L.; Wu, W.; Lee, Y.-K.; Xie, J.; Zhang, H. Spatial Heterogeneity and Co-Occurrence of Mucosal and Luminal Microbiome across Swine Intestinal Tract. Front. Microbiol. 2018, 9, 48. [Google Scholar] [CrossRef]
  143. Mei, Z.; Li, D. The Role of Probiotics in Vaginal Health. Front. Cell. Infect. Microbiol. 2022, 12, 963868. [Google Scholar] [CrossRef] [PubMed]
  144. Mitchell, C.; Marrazzo, J. Bacterial Vaginosis and the Cervicovaginal Immune Response. Am. J. Reprod. Immunol. 2014, 71, 555–563. [Google Scholar] [CrossRef] [PubMed]
  145. Plesniarski, A.; Siddik, A.B.; Su, R.-C. The Microbiome as a Key Regulator of Female Genital Tract Barrier Function. Front. Cell. Infect. Microbiol. 2021, 11, 790627. [Google Scholar] [CrossRef] [PubMed]
  146. Happel, A.-U.; Varsani, A.; Balle, C.; Passmore, J.-A.; Jaspan, H. The Vaginal Virome-Balancing Female Genital Tract Bacteriome, Mucosal Immunity, and Sexual and Reproductive Health Outcomes? Viruses 2020, 12, 832. [Google Scholar] [CrossRef]
  147. Pruski, P.; Lewis, H.V.; Lee, Y.S.; Marchesi, J.R.; Bennett, P.R.; Takats, Z.; MacIntyre, D.A. Assessment of Microbiota:Host Interactions at the Vaginal Mucosa Interface. Methods 2018, 149, 74–84. [Google Scholar] [CrossRef]
  148. Valenti, P.; Rosa, L.; Capobianco, D.; Lepanto, M.S.; Schiavi, E.; Cutone, A.; Paesano, R.; Mastromarino, P. Role of Lactobacilli and Lactoferrin in the Mucosal Cervicovaginal Defense. Front. Immunol. 2018, 9, 376. [Google Scholar] [CrossRef]
  149. Mahapatro, M.; Erkert, L.; Becker, C. Cytokine-Mediated Crosstalk between Immune Cells and Epithelial Cells in the Gut. Cells 2021, 10, 111. [Google Scholar] [CrossRef]
  150. Hinks, T.S.C.; Zhang, X.-W. MAIT Cell Activation and Functions. Front. Immunol. 2020, 11, 1014. [Google Scholar] [CrossRef]
  151. Mantis, N.J.; Rol, N.; Corthésy, B. Secretory IgA’s Complex Roles in Immunity and Mucosal Homeostasis in the Gut. Mucosal Immunol. 2011, 4, 603–611. [Google Scholar] [CrossRef] [PubMed]
  152. Cancro, M.P.; Tomayko, M.M. Memory B Cells and Plasma Cells: The Differentiative Continuum of Humoral Immunity. Immunol. Rev. 2021, 303, 72–82. [Google Scholar] [CrossRef] [PubMed]
  153. Li, Y.; Jin, L.; Chen, T. The Effects of Secretory IgA in the Mucosal Immune System. Biomed. Res. Int. 2020, 2020, 2032057. [Google Scholar] [CrossRef] [PubMed]
  154. Cerutti, A.; Chen, K.; Chorny, A. Immunoglobulin Responses at the Mucosal Interface. Annu. Rev. Immunol. 2011, 29, 273–293. [Google Scholar] [CrossRef] [PubMed]
  155. Woof, J.M.; Russell, M.W. Structure and Function Relationships in IgA. Mucosal Immunol. 2011, 4, 590–597. [Google Scholar] [CrossRef] [PubMed]
  156. Brandtzaeg, P.; Farstad, I.N.; Johansen, F.E.; Morton, H.C.; Norderhaug, I.N.; Yamanaka, T. The B-Cell System of Human Mucosae and Exocrine Glands. Immunol. Rev. 1999, 171, 45–87. [Google Scholar] [CrossRef] [PubMed]
  157. Craig, S.W.; Cebra, J.J. Peyer’s Patches: An Enriched Source of Precursors for IgA-Producing Immunocytes in the Rabbit. J. Exp. Med. 1971, 134, 188–200. [Google Scholar] [CrossRef]
  158. Kaetzel, C.S. Cooperativity among Secretory IgA, the Polymeric Immunoglobulin Receptor, and the Gut Microbiota Promotes Host-Microbial Mutualism. Immunol. Lett. 2014, 162, 10–21. [Google Scholar] [CrossRef] [PubMed]
  159. Farache, J.; Koren, I.; Milo, I.; Gurevich, I.; Kim, K.-W.; Zigmond, E.; Furtado, G.C.; Lira, S.A.; Shakhar, G. Luminal Bacteria Recruit CD103+ Dendritic Cells into the Intestinal Epithelium to Sample Bacterial Antigens for Presentation. Immunity 2013, 38, 581–595. [Google Scholar] [CrossRef] [PubMed]
  160. Mabbott, N.A.; Donaldson, D.S.; Ohno, H.; Williams, I.R.; Mahajan, A. Microfold (M) Cells: Important Immunosurveillance Posts in the Intestinal Epithelium. Mucosal Immunol. 2013, 6, 666–677. [Google Scholar] [CrossRef]
  161. Kujala, P.; Raymond, C.R.; Romeijn, M.; Godsave, S.F.; van Kasteren, S.I.; Wille, H.; Prusiner, S.B.; Mabbott, N.A.; Peters, P.J. Prion Uptake in the Gut: Identification of the First Uptake and Replication Sites. PLoS Pathog. 2011, 7, e1002449. [Google Scholar] [CrossRef] [PubMed]
  162. McDole, J.R.; Wheeler, L.W.; McDonald, K.G.; Wang, B.; Konjufca, V.; Knoop, K.A.; Newberry, R.D.; Miller, M.J. Goblet Cells Deliver Luminal Antigen to CD103+ Dendritic Cells in the Small Intestine. Nature 2012, 483, 345–349. [Google Scholar] [CrossRef]
  163. Reboldi, A.; Arnon, T.I.; Rodda, L.B.; Atakilit, A.; Sheppard, D.; Cyster, J.G. IgA Production Requires B Cell Interaction with Subepithelial Dendritic Cells in Peyer’s Patches. Science 2016, 352, aaf4822. [Google Scholar] [CrossRef]
  164. Cerutti, A. The Regulation of IgA Class Switching. Nat. Rev. Immunol. 2008, 8, 421–434. [Google Scholar] [CrossRef]
  165. Stavnezer, J.; Kang, J. The Surprising Discovery That TGF Beta Specifically Induces the IgA Class Switch. J. Immunol. 2009, 182, 5–7. [Google Scholar] [CrossRef]
  166. Pabst, O. New Concepts in the Generation and Functions of IgA. Nat. Rev. Immunol. 2012, 12, 821–832. [Google Scholar] [CrossRef]
  167. Kubinak, J.L.; Petersen, C.; Stephens, W.Z.; Soto, R.; Bake, E.; O’Connell, R.M.; Round, J.L. MyD88 Signaling in T Cells Directs IgA-Mediated Control of the Microbiota to Promote Health. Cell Host Microbe 2015, 17, 153–163. [Google Scholar] [CrossRef]
  168. Lycke, N.; Eriksen, L.; Holmgren, J. Protection against Cholera Toxin after Oral Immunization Is Thymus-Dependent and Associated with Intestinal Production of Neutralizing IgA Antitoxin. Scand. J. Immunol. 1987, 25, 413–419. [Google Scholar] [CrossRef] [PubMed]
  169. Boyaka, P.N. Inducing Mucosal IgA: A Challenge for Vaccine Adjuvants and Delivery Systems. J. Immunol. 2017, 199, 9–16. [Google Scholar] [CrossRef]
  170. Duchemin, M.; Tudor, D.; Cottignies-Calamarte, A.; Bomsel, M. Antibody-Dependent Cellular Phagocytosis of HIV-1-Infected Cells Is Efficiently Triggered by IgA Targeting HIV-1 Envelope Subunit Gp41. Front. Immunol. 2020, 11, 1141. [Google Scholar] [CrossRef]
  171. Chen, X.; Sun, H.-Y.; Lee, C.Y.; Rostad, C.A.; Trost, J.; Abreu, R.B.; Carlock, M.A.; Wilson, J.R.; Gansebom, S.; Ross, T.M.; et al. Functional Antibody-Dependent Cell Mediated Cytotoxicity (ADCC) Responses to Vaccine and Circulating Influenza Strains Following Vaccination. Virology 2022, 569, 44–55. [Google Scholar] [CrossRef]
  172. Dickinson, B.L.; Badizadegan, K.; Wu, Z.; Ahouse, J.C.; Zhu, X.; Simister, N.E.; Blumberg, R.S.; Lencer, W.I. Bidirectional FcRn-Dependent IgG Transport in a Polarized Human Intestinal Epithelial Cell Line. J. Clin. Invest. 1999, 104, 903–911. [Google Scholar] [CrossRef]
  173. Bemark, M.; Angeletti, D. Know Your Enemy or Find Your Friend?-Induction of IgA at Mucosal Surfaces. Immunol. Rev. 2021, 303, 83–102. [Google Scholar] [CrossRef]
  174. Lamm, M.E.; Nedrud, J.G.; Kaetzel, C.S.; Mazanec, M.B. IgA and Mucosal Defense. APMIS 1995, 103, 241–246. [Google Scholar] [CrossRef]
  175. Iwasaki, A. Exploiting Mucosal Immunity for Antiviral Vaccines. Annu. Rev. Immunol. 2016, 34, 575–608. [Google Scholar] [CrossRef]
  176. Kryukova, N.; Baranova, I.; Abramova, N.; Khromova, E.; Pachomov, D.; Svitich, O.; Chuchalin, A.; Kostinov, M. Mucosal Immunity in Health Care Workers’ Respiratory Tracts in the Post-COVID-19 Period. Sci. Rep. 2023, 13, 7162. [Google Scholar] [CrossRef]
  177. Castro-Dopico, T.; Clatworthy, M.R. IgG and Fcγ Receptors in Intestinal Immunity and Inflammation. Front. Immunol. 2019, 10, 805. [Google Scholar] [CrossRef] [PubMed]
  178. Russell, M.W.; Moldoveanu, Z.; Ogra, P.L.; Mestecky, J. Mucosal Immunity in COVID-19: A Neglected but Critical Aspect of SARS-CoV-2 Infection. Front. Immunol. 2020, 11, 611337. [Google Scholar] [CrossRef] [PubMed]
  179. Kammona, O.; Kiparissides, C. Recent Advances in Nanocarrier-Based Mucosal Delivery of Biomolecules. J. Control. Release 2012, 161, 781–794. [Google Scholar] [CrossRef]
  180. Tiwari, S.; Agrawal, G.P.; Vyas, S.P. Molecular Basis of the Mucosal Immune System: From Fundamental Concepts to Advances in Liposome-Based Vaccines. Nanomedicine 2010, 5, 1617–1640. [Google Scholar] [CrossRef]
  181. Corr, S.C.; Gahan, C.C.G.M.; Hill, C. M-Cells: Origin, Morphology and Role in Mucosal Immunity and Microbial Pathogenesis. FEMS Immunol. Med. Microbiol. 2008, 52, 2–12. [Google Scholar] [CrossRef]
  182. Kumar, B.V.; Ma, W.; Miron, M.; Granot, T.; Guyer, R.S.; Carpenter, D.J.; Senda, T.; Sun, X.; Ho, S.-H.; Lerner, H.; et al. Human Tissue-Resident Memory T Cells Are Defined by Core Transcriptional and Functional Signatures in Lymphoid and Mucosal Sites. Cell Rep. 2017, 20, 2921–2934. [Google Scholar] [CrossRef]
  183. Netea, M.G.; Joosten, L.A.B.; Latz, E.; Mills, K.H.G.; Natoli, G.; Stunnenberg, H.G.; O’Neill, L.A.J.; Xavier, R.J. Trained Immunity: A Program of Innate Immune Memory in Health and Disease. Science 2016, 352, aaf1098. [Google Scholar] [CrossRef] [PubMed]
  184. Sui, Y.; Berzofsky, J.A. Myeloid Cell-Mediated Trained Innate Immunity in Mucosal AIDS Vaccine Development. Front. Immunol. 2020, 11, 315. [Google Scholar] [CrossRef]
  185. D’Agostino, M.R.; Lai, R.; Afkhami, S.; Khera, A.; Yao, Y.; Vaseghi-Shanjani, M.; Zganiacz, A.; Jeyanathan, M.; Xing, Z. Airway Macrophages Mediate Mucosal Vaccine–Induced Trained Innate Immunity against Mycobacterium tuberculosis in Early Stages of Infection. J. Immunol. 2020, 205, 2750–2762. [Google Scholar] [CrossRef]
  186. Xing, Z.; Afkhami, S.; Bavananthasivam, J.; Fritz, D.K.; D’Agostino, M.R.; Vaseghi-Shanjani, M.; Yao, Y.; Jeyanathan, M. Innate Immune Memory of Tissue-Resident Macrophages and Trained Innate Immunity: Re-Vamping Vaccine Concept and Strategies. J. Leukoc. Biol. 2020, 108, 825–834. [Google Scholar] [CrossRef]
  187. Boyaka, P.N.; McGhee, J.R.; Czerkinsky, C.; Mestecky, J. Mucosal Vaccines: An Overview. In Mucosal Immunology; Elsevier: Amsterdam, The Netherlands, 2005; pp. 855–874. ISBN 978-0-12-491543-5. [Google Scholar]
  188. Russell, M.W.; Mestecky, J. Mucosal Vaccines. In Mucosal Immunology; Elsevier: Amsterdam, The Netherlands, 2015; pp. 1039–1046. ISBN 978-0-12-415847-4. [Google Scholar]
  189. Malkevitch, N.V.; Robert-Guroff, M. A Call for Replicating Vector Prime-Protein Boost Strategies in HIV Vaccine Design. Expert Rev. Vaccines 2004, 3, S105–S117. [Google Scholar] [CrossRef]
  190. Terrinoni, M.; Nordqvist, S.L.; Löfstrand, M.; Nilsson, F.; Källgård, S.; Sharma, T.; Lebens, M.R.; Holmgren, J. A Thermostable, Dry Formulation Inactivated Hikojima Whole Cell/Cholera Toxin B Subunit Oral Cholera Vaccine. Vaccine 2023, 41, 3347–3357. [Google Scholar] [CrossRef] [PubMed]
  191. Tam, H.H.; Melo, M.B.; Kang, M.; Pelet, J.M.; Ruda, V.M.; Foley, M.H.; Hu, J.K.; Kumari, S.; Crampton, J.; Baldeon, A.D.; et al. Sustained Antigen Availability during Germinal Center Initiation Enhances Antibody Responses to Vaccination. Proc. Natl. Acad. Sci. USA 2016, 113, E6639–E6648. [Google Scholar] [CrossRef] [PubMed]
  192. Krammer, F.; Palese, P. Universal Influenza Virus Vaccines That Target the Conserved Hemagglutinin Stalk and Conserved Sites in the Head Domain. J. Infect. Dis. 2019, 219, S62–S67. [Google Scholar] [CrossRef] [PubMed]
  193. Edgar, J.E.; Trezise, S.; Anthony, R.M.; Krammer, F.; Palese, P.; Ravetch, J.V.; Bournazos, S. Antibodies Elicited in Humans upon Chimeric Hemagglutinin-Based Influenza Virus Vaccination Confer FcγR-Dependent Protection in Vivo. Proc. Natl. Acad. Sci. USA 2023, 120, e2314905120. [Google Scholar] [CrossRef]
  194. Isakova-Sivak, I.; Korenkov, D.; Smolonogina, T.; Kotomina, T.; Donina, S.; Matyushenko, V.; Mezhenskaya, D.; Krammer, F.; Rudenko, L. Broadly Protective Anti-Hemagglutinin Stalk Antibodies Induced by Live Attenuated Influenza Vaccine Expressing Chimeric Hemagglutinin. Virology 2018, 518, 313–323. [Google Scholar] [CrossRef]
  195. Krammer, F.; Pica, N.; Hai, R.; Margine, I.; Palese, P. Chimeric Hemagglutinin Influenza Virus Vaccine Constructs Elicit Broadly Protective Stalk-Specific Antibodies. J. Virol. 2013, 87, 6542–6550. [Google Scholar] [CrossRef]
  196. Rollier, C.S.; Reyes-Sandoval, A.; Cottingham, M.G.; Ewer, K.; Hill, A.V.S. Viral Vectors as Vaccine Platforms: Deployment in Sight. Curr. Opin. Immunol. 2011, 23, 377–382. [Google Scholar] [CrossRef] [PubMed]
  197. Sandbrink, J.B.; Koblentz, G.D. Biosecurity Risks Associated with Vaccine Platform Technologies. Vaccine 2022, 40, 2514–2523. [Google Scholar] [CrossRef]
  198. Shiver, J.W.; Emini, E.A. Recent Advances in the Development of HIV-1 Vaccines Using Replication-Incompetent Adenovirus Vectors. Annu. Rev. Med. 2004, 55, 355–372. [Google Scholar] [CrossRef]
  199. Wands, A.M.; Fujita, A.; McCombs, J.E.; Cervin, J.; Dedic, B.; Rodriguez, A.C.; Nischan, N.; Bond, M.R.; Mettlen, M.; Trudgian, D.C.; et al. Fucosylation and Protein Glycosylation Create Functional Receptors for Cholera Toxin. eLife 2015, 4, e09545. [Google Scholar] [CrossRef]
  200. Gustafsson, T.; Hua, Y.-J.; Dahlgren, M.W.; Livingston, M.; Johansson-Lindbom, B.; Yrlid, U. Direct Interaction between Cholera Toxin and Dendritic Cells Is Required for Oral Adjuvant Activity. Eur. J. Immunol. 2013, 43, 1779–1788. [Google Scholar] [CrossRef]
  201. Holmgren, J.; Lönnroth, I.; Svennerholm, L. Tissue Receptor for Cholera Exotoxin: Postulated Structure from Studies with GM1 Ganglioside and Related Glycolipids. Infect. Immun. 1973, 8, 208–214. [Google Scholar] [CrossRef] [PubMed]
  202. Sethi, A.; Wands, A.M.; Mettlen, M.; Krishnamurthy, S.; Wu, H.; Kohler, J.J. Cell Type and Receptor Identity Regulate Cholera Toxin Subunit B (CTB) Internalization. Interface Focus 2019, 9, 20180076. [Google Scholar] [CrossRef] [PubMed]
  203. Jertborn, M.; Nordström, I.; Kilander, A.; Czerkinsky, C.; Holmgren, J. Local and Systemic Immune Responses to Rectal Administration of Recombinant Cholera Toxin B Subunit in Humans. Infect. Immun. 2001, 69, 4125–4128. [Google Scholar] [CrossRef]
  204. Harris, J.B. Cholera: Immunity and Prospects in Vaccine Development. J. Infect. Dis. 2018, 218, S141–S146. [Google Scholar] [CrossRef]
  205. Bergquist, C.; Johansson, E.L.; Lagergård, T.; Holmgren, J.; Rudin, A. Intranasal Vaccination of Humans with Recombinant Cholera Toxin B Subunit Induces Systemic and Local Antibody Responses in the Upper Respiratory Tract and the Vagina. Infect. Immun. 1997, 65, 2676–2684. [Google Scholar] [CrossRef] [PubMed]
  206. Clemens, J.D.; Sack, D.A.; Harris, J.R.; Chakraborty, J.; Neogy, P.K.; Stanton, B.; Huda, N.; Khan, M.U.; Kay, B.A.; Khan, M.R. Cross-Protection by B Subunit-Whole Cell Cholera Vaccine against Diarrhea Associated with Heat-Labile Toxin-Producing Enterotoxigenic Escherichia coli: Results of a Large-Scale Field Trial. J. Infect. Dis. 1988, 158, 372–377. [Google Scholar] [CrossRef]
  207. Scerpella, E.G.; Sanchez, J.L.; Mathewson, I.I.I.; Torres-Cordero, J.V.; Sadoff, J.C.; Svennerholm, A.M.; DuPont, H.L.; Taylor, D.N.; Ericsson, C.D. Safety, Immunogenicity, and Protective Efficacy of the Whole-Cell/Recombinant B Subunit (WC/rBS) Oral Cholera Vaccine Against Travelers’ Diarrhea. J. Travel Med. 1995, 2, 22–27. [Google Scholar] [CrossRef]
  208. Antonio-Herrera, L.; Badillo-Godinez, O.; Medina-Contreras, O.; Tepale-Segura, A.; García-Lozano, A.; Gutierrez-Xicotencatl, L.; Soldevila, G.; Esquivel-Guadarrama, F.R.; Idoyaga, J.; Bonifaz, L.C. The Nontoxic Cholera B Subunit Is a Potent Adjuvant for Intradermal DC-Targeted Vaccination. Front. Immunol. 2018, 9, 2212. [Google Scholar] [CrossRef]
  209. Yang, X.; Wang, X.; Song, Y.; Zhou, P.; Li, D.; Zhang, C.; Jin, X.; Huang, Z.; Zhou, D. Chimpanzee Adenoviral Vector Prime-Boost Regimen Elicits Potent Immune Responses against Ebola Virus in Mice and Rhesus Macaques. Emerg. Microbes Infect. 2019, 8, 1086–1097. [Google Scholar] [CrossRef]
  210. Zhou, D.; Wu, T.-L.; Lasaro, M.O.; Latimer, B.P.; Parzych, E.M.; Bian, A.; Li, Y.; Li, H.; Erikson, J.; Xiang, Z.; et al. A Universal Influenza A Vaccine Based on Adenovirus Expressing Matrix-2 Ectodomain and Nucleoprotein Protects Mice from Lethal Challenge. Mol. Ther. 2010, 18, 2182–2189. [Google Scholar] [CrossRef]
  211. Sallard, E.; Zhang, W.; Aydin, M.; Schröer, K.; Ehrhardt, A. The Adenovirus Vector Platform: Novel Insights into Rational Vector Design and Lessons Learned from the COVID-19 Vaccine. Viruses 2023, 15, 204. [Google Scholar] [CrossRef] [PubMed]
  212. Hartman, Z.C.; Appledorn, D.M.; Amalfitano, A. Adenovirus Vector Induced Innate Immune Responses: Impact upon Efficacy and Toxicity in Gene Therapy and Vaccine Applications. Virus Res. 2008, 132, 1–14. [Google Scholar] [CrossRef] [PubMed]
  213. Cao, K.; Wang, X.; Peng, H.; Ding, L.; Wang, X.; Hu, Y.; Dong, L.; Yang, T.; Hong, X.; Xing, M.; et al. A Single Vaccine Protects against SARS-CoV-2 and Influenza Virus in Mice. J. Virol. 2022, 96, e0157821. [Google Scholar] [CrossRef] [PubMed]
  214. Cheng, T.; Wang, X.; Song, Y.; Tang, X.; Zhang, C.; Zhang, H.; Jin, X.; Zhou, D. Chimpanzee Adenovirus Vector-Based Avian Influenza Vaccine Completely Protects Mice against Lethal Challenge of H5N1. Vaccine 2016, 34, 4875–4883. [Google Scholar] [CrossRef] [PubMed]
  215. Xu, K.; Song, Y.; Dai, L.; Zhang, Y.; Lu, X.; Xie, Y.; Zhang, H.; Cheng, T.; Wang, Q.; Huang, Q.; et al. Recombinant Chimpanzee Adenovirus Vaccine AdC7-M/E Protects against Zika Virus Infection and Testis Damage. J. Virol. 2018, 92, e01722-17. [Google Scholar] [CrossRef] [PubMed]
  216. Xu, F.; Wu, S.; Yi, L.; Peng, S.; Wang, F.; Si, W.; Hou, L.; Zhu, T. Safety, Mucosal and Systemic Immunopotency of an Aerosolized Adenovirus-Vectored Vaccine against SARS-CoV-2 in Rhesus Macaques. Emerg. Microbes Infect. 2022, 11, 438–441. [Google Scholar] [CrossRef] [PubMed]
  217. Tang, X.; Zhang, H.; Song, Y.; Zhou, D.; Wang, J. Hemagglutinin-Targeting Artificial MicroRNAs Expressed by Adenovirus Protect Mice From Different Clades of H5N1 Infection. Mol. Ther. Nucleic Acids 2016, 5, e311. [Google Scholar] [CrossRef]
  218. Sakurai, F.; Tachibana, M.; Mizuguchi, H. Adenovirus Vector-Based Vaccine for Infectious Diseases. Drug Metab. Pharmacokinet. 2022, 42, 100432. [Google Scholar] [CrossRef]
  219. Gebre, M.S.; Brito, L.A.; Tostanoski, L.H.; Edwards, D.K.; Carfi, A.; Barouch, D.H. Novel Approaches for Vaccine Development. Cell 2021, 184, 1589–1603. [Google Scholar] [CrossRef] [PubMed]
  220. Zhu, F.-C.; Li, Y.-H.; Guan, X.-H.; Hou, L.-H.; Wang, W.-J.; Li, J.-X.; Wu, S.-P.; Wang, B.-S.; Wang, Z.; Wang, L.; et al. Safety, Tolerability, and Immunogenicity of a Recombinant Adenovirus Type-5 Vectored COVID-19 Vaccine: A Dose-Escalation, Open-Label, Non-Randomised, First-in-Human Trial. Lancet 2020, 395, 1845–1854. [Google Scholar] [CrossRef] [PubMed]
  221. Gray, G.E.; Allen, M.; Moodie, Z.; Churchyard, G.; Bekker, L.-G.; Nchabeleng, M.; Mlisana, K.; Metch, B.; de Bruyn, G.; Latka, M.H.; et al. Safety and Efficacy of the HVTN 503/Phambili Study of a Clade-B-Based HIV-1 Vaccine in South Africa: A Double-Blind, Randomised, Placebo-Controlled Test-of-Concept Phase 2b Study. Lancet Infect. Dis. 2011, 11, 507–515. [Google Scholar] [CrossRef] [PubMed]
  222. Buchbinder, S.P.; Mehrotra, D.V.; Duerr, A.; Fitzgerald, D.W.; Mogg, R.; Li, D.; Gilbert, P.B.; Lama, J.R.; Marmor, M.; Del Rio, C.; et al. Efficacy Assessment of a Cell-Mediated Immunity HIV-1 Vaccine (the Step Study): A Double-Blind, Randomised, Placebo-Controlled, Test-of-Concept Trial. Lancet 2008, 372, 1881–1893. [Google Scholar] [CrossRef] [PubMed]
  223. Buchbinder, S.P.; McElrath, M.J.; Dieffenbach, C.; Corey, L. Use of Adenovirus Type-5 Vectored Vaccines: A Cautionary Tale. Lancet 2020, 396, e68–e69. [Google Scholar] [CrossRef] [PubMed]
  224. Duerr, A.; Huang, Y.; Buchbinder, S.; Coombs, R.W.; Sanchez, J.; del Rio, C.; Casapia, M.; Santiago, S.; Gilbert, P.; Corey, L.; et al. Extended Follow-up Confirms Early Vaccine-Enhanced Risk of HIV Acquisition and Demonstrates Waning Effect over Time among Participants in a Randomized Trial of Recombinant Adenovirus HIV Vaccine (Step Study). J. Infect. Dis. 2012, 206, 258–266. [Google Scholar] [CrossRef]
  225. Moodie, Z.; Metch, B.; Bekker, L.-G.; Churchyard, G.; Nchabeleng, M.; Mlisana, K.; Laher, F.; Roux, S.; Mngadi, K.; Innes, C.; et al. Continued Follow-Up of Phambili Phase 2b Randomized HIV-1 Vaccine Trial Participants Supports Increased HIV-1 Acquisition among Vaccinated Men. PLoS ONE 2015, 10, e0137666. [Google Scholar] [CrossRef]
  226. Frahm, N.; DeCamp, A.C.; Friedrich, D.P.; Carter, D.K.; Defawe, O.D.; Kublin, J.G.; Casimiro, D.R.; Duerr, A.; Robertson, M.N.; Buchbinder, S.P.; et al. Human Adenovirus-Specific T Cells Modulate HIV-Specific T Cell Responses to an Ad5-Vectored HIV-1 Vaccine. J. Clin. Investig. 2012, 122, 359–367. [Google Scholar] [CrossRef]
  227. Perreau, M.; Pantaleo, G.; Kremer, E.J. Activation of a Dendritic Cell-T Cell Axis by Ad5 Immune Complexes Creates an Improved Environment for Replication of HIV in T Cells. J. Exp. Med. 2008, 205, 2717–2725. [Google Scholar] [CrossRef]
  228. Auclair, S.; Liu, F.; Niu, Q.; Hou, W.; Churchyard, G.; Morgan, C.; Frahm, N.; Nitayaphan, S.; Pitisuthithum, P.; Rerks-Ngarm, S.; et al. Distinct Susceptibility of HIV Vaccine Vector-Induced CD4 T Cells to HIV Infection. PLoS Pathog. 2018, 14, e1006888. [Google Scholar] [CrossRef] [PubMed]
  229. Qureshi, H.; Ma, Z.-M.; Huang, Y.; Hodge, G.; Thomas, M.A.; DiPasquale, J.; DeSilva, V.; Fritts, L.; Bett, A.J.; Casimiro, D.R.; et al. Low-Dose Penile SIVmac251 Exposure of Rhesus Macaques Infected with Adenovirus Type 5 (Ad5) and Then Immunized with a Replication-Defective Ad5-Based SIV Gag/Pol/Nef Vaccine Recapitulates the Results of the Phase IIb Step Trial of a Similar HIV-1 Vaccine. J. Virol. 2012, 86, 2239–2250. [Google Scholar] [CrossRef]
  230. Westwood, A.; Healey, G.D.; Williamson, E.D.; Eyles, J.E. Activation of Dendritic Cells by Microparticles Containing Bacillus anthracis Protective Antigen. Vaccine 2005, 23, 3857–3863. [Google Scholar] [CrossRef]
  231. Johansen, P.; Storni, T.; Rettig, L.; Qiu, Z.; Der-Sarkissian, A.; Smith, K.A.; Manolova, V.; Lang, K.S.; Senti, G.; Müllhaupt, B.; et al. Antigen Kinetics Determines Immune Reactivity. Proc. Natl. Acad. Sci. USA 2008, 105, 5189–5194. [Google Scholar] [CrossRef]
  232. Gonzalez-Cruz, P.; Gill, H.S. Demystifying Particle-Based Oral Vaccines. Expert. Opin. Drug Deliv. 2021, 18, 1455–1472. [Google Scholar] [CrossRef] [PubMed]
  233. Sharp, F.A.; Ruane, D.; Claass, B.; Creagh, E.; Harris, J.; Malyala, P.; Singh, M.; O’Hagan, D.T.; Pétrilli, V.; Tschopp, J.; et al. Uptake of Particulate Vaccine Adjuvants by Dendritic Cells Activates the NALP3 Inflammasome. Proc. Natl. Acad. Sci. USA 2009, 106, 870–875. [Google Scholar] [CrossRef] [PubMed]
  234. Alameh, M.-G.; Tombácz, I.; Bettini, E.; Lederer, K.; Sittplangkoon, C.; Wilmore, J.R.; Gaudette, B.T.; Soliman, O.Y.; Pine, M.; Hicks, P.; et al. Lipid Nanoparticles Enhance the Efficacy of mRNA and Protein Subunit Vaccines by Inducing Robust T Follicular Helper Cell and Humoral Responses. Immunity 2021, 54, 2877–2892. [Google Scholar] [CrossRef] [PubMed]
  235. Ghaffari Marandi, B.H.; Zolfaghari, M.R.; Kazemi, R.; Motamedi, M.J.; Amani, J. Immunization against Vibrio Cholerae, ETEC, and EHEC with Chitosan Nanoparticle Containing LSC Chimeric Protein. Microb. Pathog. 2019, 134, 103600. [Google Scholar] [CrossRef] [PubMed]
  236. Saeed, M.I.; Omar, A.R.; Hussein, M.Z.; Elkhidir, I.M.; Sekawi, Z. Development of Enhanced Antibody Response toward Dual Delivery of Nano-Adjuvant Adsorbed Human Enterovirus-71 Vaccine Encapsulated Carrier. Hum. Vaccin. Immunother. 2015, 11, 2414–2424. [Google Scholar] [CrossRef]
  237. Chen, S.C.; Jones, D.H.; Fynan, E.F.; Farrar, G.H.; Clegg, J.C.; Greenberg, H.B.; Herrmann, J.E. Protective Immunity Induced by Oral Immunization with a Rotavirus DNA Vaccine Encapsulated in Microparticles. J. Virol. 1998, 72, 5757–5761. [Google Scholar] [CrossRef] [PubMed]
  238. Joyce, M.G.; King, H.A.D.; Elakhal-Naouar, I.; Ahmed, A.; Peachman, K.K.; Macedo Cincotta, C.; Subra, C.; Chen, R.E.; Thomas, P.V.; Chen, W.-H.; et al. A SARS-CoV-2 Ferritin Nanoparticle Vaccine Elicits Protective Immune Responses in Nonhuman Primates. Sci. Transl. Med. 2022, 14, eabi5735. [Google Scholar] [CrossRef] [PubMed]
  239. Houser, K.V.; Chen, G.L.; Carter, C.; Crank, M.C.; Nguyen, T.A.; Burgos Florez, M.C.; Berkowitz, N.M.; Mendoza, F.; Hendel, C.S.; Gordon, I.J.; et al. Safety and Immunogenicity of a Ferritin Nanoparticle H2 Influenza Vaccine in Healthy Adults: A Phase 1 Trial. Nat. Med. 2022, 28, 383–391. [Google Scholar] [CrossRef]
  240. Rodrigues, M.Q.; Alves, P.M.; Roldão, A. Functionalizing Ferritin Nanoparticles for Vaccine Development. Pharmaceutics 2021, 13, 1621. [Google Scholar] [CrossRef]
  241. Lamb, Y.N. BNT162b2 mRNA COVID-19 Vaccine: First Approval. Drugs 2021, 81, 495–501. [Google Scholar] [CrossRef]
  242. Szabó, G.T.; Mahiny, A.J.; Vlatkovic, I. COVID-19 mRNA Vaccines: Platforms and Current Developments. Mol. Ther. 2022, 30, 1850–1868. [Google Scholar] [CrossRef]
  243. Shi, W.; Liu, J.; Huang, Y.; Qiao, L. Papillomavirus Pseudovirus: A Novel Vaccine to Induce Mucosal and Systemic Cytotoxic T-Lymphocyte Responses. J. Virol. 2001, 75, 10139–10148. [Google Scholar] [CrossRef]
  244. Yang, L.; Song, Y.; Li, X.; Huang, X.; Liu, J.; Ding, H.; Zhu, P.; Zhou, P. HIV-1 Virus-like Particles Produced by Stably Transfected Drosophila S2 Cells: A Desirable Vaccine Component. J. Virol. 2012, 86, 7662–7676. [Google Scholar] [CrossRef]
  245. Plana-Duran, J.; Bastons, M.; Rodriguez, M.J.; Climent, I.; Cortés, E.; Vela, C.; Casal, I. Oral Immunization of Rabbits with VP60 Particles Confers Protection against Rabbit Hemorrhagic Disease. Arch. Virol. 1996, 141, 1423–1436. [Google Scholar] [CrossRef]
  246. Huang, X.; Zhu, Q.; Huang, X.; Yang, L.; Song, Y.; Zhu, P.; Zhou, P. In Vivo Electroporation in DNA-VLP Prime-Boost Preferentially Enhances HIV-1 Envelope-Specific IgG2a, Neutralizing Antibody and CD8 T Cell Responses. Vaccine 2017, 35, 2042–2051. [Google Scholar] [CrossRef] [PubMed]
  247. Akhtar, M.; Chowdhury, M.I.; Bhuiyan, T.R.; Kaim, J.; Ahmed, T.; Rafique, T.A.; Khan, A.; Rahman, S.I.A.; Khanam, F.; Begum, Y.A.; et al. Evaluation of the Safety and Immunogenicity of the Oral Inactivated Multivalent Enterotoxigenic Escherichia coli Vaccine ETVAX in Bangladeshi Adults in a Double-Blind, Randomized, Placebo-Controlled Phase I Trial Using Electrochemiluminescence and ELISA Assays for Immunogenicity Analyses. Vaccine 2019, 37, 5645–5656. [Google Scholar] [CrossRef] [PubMed]
  248. 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.; et al. 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] [PubMed]
  249. Rakhra, K.; Abraham, W.; Wang, C.; Moynihan, K.D.; Li, N.; Donahue, N.; Baldeon, A.D.; Irvine, D.J. Exploiting Albumin as a Mucosal Vaccine Chaperone for Robust Generation of Lung-Resident Memory T Cells. Sci. Immunol. 2021, 6, eabd8003. [Google Scholar] [CrossRef]
  250. Bastola, R.; Noh, G.; Keum, T.; Bashyal, S.; Seo, J.-E.; Choi, J.; Oh, Y.; Cho, Y.; Lee, S. Vaccine Adjuvants: Smart Components to Boost the Immune System. Arch. Pharm. Res. 2017, 40, 1238–1248. [Google Scholar] [CrossRef]
  251. Tandrup Schmidt, S.; Foged, C.; Korsholm, K.S.; Rades, T.; Christensen, D. Liposome-Based Adjuvants for Subunit Vaccines: Formulation Strategies for Subunit Antigens and Immunostimulators. Pharmaceutics 2016, 8, 7. [Google Scholar] [CrossRef]
  252. Cox, J.C.; Coulter, A.R. Adjuvants--a Classification and Review of Their Modes of Action. Vaccine 1997, 15, 248–256. [Google Scholar] [CrossRef]
  253. Lee, S.; Nguyen, M.T. Recent Advances of Vaccine Adjuvants for Infectious Diseases. Immune Netw. 2015, 15, 51–57. [Google Scholar] [CrossRef]
  254. Verma, S.K.; Mahajan, P.; Singh, N.K.; Gupta, A.; Aggarwal, R.; Rappuoli, R.; Johri, A.K. New-Age Vaccine Adjuvants, Their Development, and Future Perspective. Front. Immunol. 2023, 14, 1043109. [Google Scholar] [CrossRef]
  255. Riese, P.; Schulze, K.; Ebensen, T.; Prochnow, B.; Guzmán, C.A. Vaccine Adjuvants: Key Tools for Innovative Vaccine Design. Curr. Top. Med. Chem. 2013, 13, 2562–2580. [Google Scholar] [CrossRef]
  256. Rhee, J.H.; Lee, S.E.; Kim, S.Y. Mucosal Vaccine Adjuvants Update. Clin. Exp. Vaccine Res. 2012, 1, 50–63. [Google Scholar] [CrossRef]
  257. Corthésy, B.; Bioley, G. Lipid-Based Particles: Versatile Delivery Systems for Mucosal Vaccination against Infection. Front. Immunol. 2018, 9, 431. [Google Scholar] [CrossRef]
  258. Petrovsky, N. Comparative Safety of Vaccine Adjuvants: A Summary of Current Evidence and Future Needs. Drug Saf. 2015, 38, 1059–1074. [Google Scholar] [CrossRef]
  259. Savelkoul, H.F.J.; Ferro, V.A.; Strioga, M.M.; Schijns, V.E.J.C. Choice and Design of Adjuvants for Parenteral and Mucosal Vaccines. Vaccines 2015, 3, 148–171. [Google Scholar] [CrossRef]
  260. Wang, P. Natural and Synthetic Saponins as Vaccine Adjuvants. Vaccines 2021, 9, 222. [Google Scholar] [CrossRef]
  261. Zhu, D.; Tuo, W. QS-21: A Potent Vaccine Adjuvant. Nat. Prod. Chem. Res. 2016, 3, e113. [Google Scholar] [CrossRef]
  262. Garçon, N.; Van Mechelen, M. Recent Clinical Experience with Vaccines Using MPL- and QS-21-Containing Adjuvant Systems. Expert Rev. Vaccines 2011, 10, 471–486. [Google Scholar] [CrossRef]
  263. O’Hagan, D.T.; Friedland, L.R.; Hanon, E.; Didierlaurent, A.M. Towards an Evidence Based Approach for the Development of Adjuvanted Vaccines. Curr. Opin. Immunol. 2017, 47, 93–102. [Google Scholar] [CrossRef]
  264. Sun, H.-X.; Xie, Y.; Ye, Y.-P. ISCOMs and ISCOMATRIX. Vaccine 2009, 27, 4388–4401. [Google Scholar] [CrossRef]
  265. Pedersen, G.K.; Höschler, K.; Øie Solbak, S.M.; Bredholt, G.; Pathirana, R.D.; Afsar, A.; Breakwell, L.; Nøstbakken, J.K.; Raae, A.J.; Brokstad, K.A.; et al. Serum IgG Titres, but Not Avidity, Correlates with Neutralizing Antibody Response after H5N1 Vaccination. Vaccine 2014, 32, 4550–4557. [Google Scholar] [CrossRef]
  266. Cibulski, S.; Rivera-Patron, M.; Suárez, N.; Pirez, M.; Rossi, S.; Yendo, A.C.; de Costa, F.; Gosmann, G.; Fett-Neto, A.; Roehe, P.M.; et al. Leaf Saponins of Quillaja Brasiliensis Enhance Long-Term Specific Immune Responses and Promote Dose-Sparing Effect in BVDV Experimental Vaccines. Vaccine 2018, 36, 55–65. [Google Scholar] [CrossRef]
  267. Pleguezuelos, O.; Dille, J.; de Groen, S.; Oftung, F.; Niesters, H.G.M.; Islam, M.A.; Næss, L.M.; Hungnes, O.; Aldarij, N.; Idema, D.L.; et al. Immunogenicity, Safety, and Efficacy of a Standalone Universal Influenza Vaccine, FLU-v, in Healthy Adults: A Randomized Clinical Trial. Ann. Intern. Med. 2020, 172, 453–462. [Google Scholar] [CrossRef]
  268. Manning, J.E.; Oliveira, F.; Coutinho-Abreu, I.V.; Herbert, S.; Meneses, C.; Kamhawi, S.; Baus, H.A.; Han, A.; Czajkowski, L.; Rosas, L.A.; et al. Safety and Immunogenicity of a Mosquito Saliva Peptide-Based Vaccine: A Randomised, Placebo-Controlled, Double-Blind, Phase 1 Trial. Lancet 2020, 395, 1998–2007. [Google Scholar] [CrossRef]
  269. Yang, J.-X.; Tseng, J.-C.; Yu, G.-Y.; Luo, Y.; Huang, C.-Y.F.; Hong, Y.-R.; Chuang, T.-H. Recent Advances in the Development of Toll-like Receptor Agonist-Based Vaccine Adjuvants for Infectious Diseases. Pharmaceutics 2022, 14, 423. [Google Scholar] [CrossRef]
  270. Fan, J.; Jin, S.; Gilmartin, L.; Toth, I.; Hussein, W.M.; Stephenson, R.J. Advances in Infectious Disease Vaccine Adjuvants. Vaccines 2022, 10, 1120. [Google Scholar] [CrossRef]
  271. Saxena, M.; Sabado, R.L.; La Mar, M.; Mohri, H.; Salazar, A.M.; Dong, H.; Correa Da Rosa, J.; Markowitz, M.; Bhardwaj, N.; Miller, E. Poly-ICLC, a TLR3 Agonist, Induces Transient Innate Immune Responses in Patients With Treated HIV-Infection: A Randomized Double-Blinded Placebo Controlled Trial. Front. Immunol. 2019, 10, 725. [Google Scholar] [CrossRef]
  272. Overton, E.T.; Goepfert, P.A.; Cunningham, P.; Carter, W.A.; Horvath, J.; Young, D.; Strayer, D.R. Intranasal Seasonal Influenza Vaccine and a TLR-3 Agonist, Rintatolimod, Induced Cross-Reactive IgA Antibody Formation against Avian H5N1 and H7N9 Influenza HA in Humans. Vaccine 2014, 32, 5490–5495. [Google Scholar] [CrossRef] [PubMed]
  273. Mallory, R.M.; Formica, N.; Pfeiffer, S.; Wilkinson, B.; Marcheschi, A.; Albert, G.; McFall, H.; Robinson, M.; Plested, J.S.; Zhu, M.; et al. Safety and Immunogenicity Following a Homologous Booster Dose of a SARS-CoV-2 Recombinant Spike Protein Vaccine (NVX-CoV2373): A Secondary Analysis of a Randomised, Placebo-Controlled, Phase 2 Trial. Lancet Infect. Dis. 2022, 22, 1565–1576. [Google Scholar] [CrossRef]
  274. Brekke, K.; Sommerfelt, M.; Ökvist, M.; Dyrhol-Riise, A.M.; Kvale, D. The Therapeutic HIV Env C5/Gp41 Vaccine Candidate Vacc-C5 Induces Specific T Cell Regulation in a Phase I/II Clinical Study. BMC Infect. Dis. 2017, 17, 228. [Google Scholar] [CrossRef]
  275. Petrina, M.; Martin, J.; Basta, S. Granulocyte Macrophage Colony-Stimulating Factor Has Come of Age: From a Vaccine Adjuvant to Antiviral Immunotherapy. Cytokine Growth Factor Rev. 2021, 59, 101–110. [Google Scholar] [CrossRef]
  276. Launay, O.; Grabar, S.; Bloch, F.; Desaint, C.; Jegou, D.; Lallemand, C.; Erickson, R.; Lebon, P.; Tovey, M.G. Effect of Sublingual Administration of Interferon-Alpha on the Immune Response to Influenza Vaccination in Institutionalized Elderly Individuals. Vaccine 2008, 26, 4073–4079. [Google Scholar] [CrossRef]
  277. Woltman, A.M.; Ter Borg, M.J.; Binda, R.S.; Sprengers, D.; von Blomberg, B.M.E.; Scheper, R.J.; Hayashi, K.; Nishi, N.; Boonstra, A.; van der Molen, R.; et al. Alpha-Galactosylceramide in Chronic Hepatitis B Infection: Results from a Randomized Placebo-Controlled Phase I/II Trial. Antivir. Ther. 2009, 14, 809–818. [Google Scholar] [CrossRef]
  278. Atmar, R.L.; Bernstein, D.I.; Harro, C.D.; Al-Ibrahim, M.S.; Chen, W.H.; Ferreira, J.; Estes, M.K.; Graham, D.Y.; Opekun, A.R.; Richardson, C.; et al. Norovirus Vaccine against Experimental Human Norwalk Virus Illness. N. Engl. J. Med. 2011, 365, 2178–2187. [Google Scholar] [CrossRef]
  279. Freytag, L.C.; Clements, J.D. Mucosal Adjuvants: New Developments and Challenges. In Mucosal Immunology; Elsevier: Amsterdam, The Netherlands, 2015; pp. 1183–1199. ISBN 978-0-12-415847-4. [Google Scholar]
  280. Freytag, L.C.; Clements, J.D. Mucosal Adjuvants. Vaccine 2005, 23, 1804–1813. [Google Scholar] [CrossRef]
  281. Chang, S.-Y.; Cha, H.-R.; Igarashi, O.; Rennert, P.D.; Kissenpfennig, A.; Malissen, B.; Nanno, M.; Kiyono, H.; Kweon, M.-N. Cutting Edge: Langerin+ Dendritic Cells in the Mesenteric Lymph Node Set the Stage for Skin and Gut Immune System Cross-Talk. J. Immunol. 2008, 180, 4361–4365. [Google Scholar] [CrossRef]
  282. Anosova, N.G.; Chabot, S.; Shreedhar, V.; Borawski, J.A.; Dickinson, B.L.; Neutra, M.R. Cholera Toxin, E. coli Heat-Labile Toxin, and Non-Toxic Derivatives Induce Dendritic Cell Migration into the Follicle-Associated Epithelium of Peyer’s Patches. Mucosal Immunol. 2008, 1, 59–67. [Google Scholar] [CrossRef]
  283. Datta, S.K.; Sabet, M.; Nguyen, K.P.L.; Valdez, P.A.; Gonzalez-Navajas, J.M.; Islam, S.; Mihajlov, I.; Fierer, J.; Insel, P.A.; Webster, N.J.; et al. Mucosal Adjuvant Activity of Cholera Toxin Requires Th17 Cells and Protects against Inhalation Anthrax. Proc. Natl. Acad. Sci. USA 2010, 107, 10638–10643. [Google Scholar] [CrossRef]
  284. Lu, Y.-J.; Yadav, P.; Clements, J.D.; Forte, S.; Srivastava, A.; Thompson, C.M.; Seid, R.; Look, J.; Alderson, M.; Tate, A.; et al. Options for Inactivation, Adjuvant, and Route of Topical Administration of a Killed, Unencapsulated Pneumococcal Whole-Cell Vaccine. Clin. Vaccine Immunol. 2010, 17, 1005–1012. [Google Scholar] [CrossRef]
  285. Shakya, A.K.; Chowdhury, M.Y.E.; Tao, W.; Gill, H.S. Mucosal Vaccine Delivery: Current State and a Pediatric Perspective. J. Control. Release 2016, 240, 394–413. [Google Scholar] [CrossRef]
  286. Bode, C.; Zhao, G.; Steinhagen, F.; Kinjo, T.; Klinman, D.M. CpG DNA as a Vaccine Adjuvant. Expert. Rev. Vaccines 2011, 10, 499–511. [Google Scholar] [CrossRef]
  287. Shi, S.; Zhu, H.; Xia, X.; Liang, Z.; Ma, X.; Sun, B. Vaccine Adjuvants: Understanding the Structure and Mechanism of Adjuvanticity. Vaccine 2019, 37, 3167–3178. [Google Scholar] [CrossRef]
  288. Gursel, M.; Klinman, D.M. Use of CpG Oligonucleotides as Mucosal Adjuvants. In Mucosal Immunology; Elsevier: Amsterdam, The Netherlands, 2015; pp. 1201–1209. ISBN 978-0-12-415847-4. [Google Scholar]
  289. Campbell, J.D. Development of the CpG Adjuvant 1018: A Case Study. Methods Mol. Biol. 2017, 1494, 15–27. [Google Scholar] [CrossRef]
  290. Zhang, Q.; Zhang, X.; Zhu, Y.; Sun, P.; Zhang, L.; Ma, J.; Zhang, Y.; Zeng, L.; Nie, X.; Gao, Y.; et al. Recognition of Cyclic Dinucleotides and Folates by Human SLC19A1. Nature 2022, 612, 170–176. [Google Scholar] [CrossRef]
  291. Luteijn, R.D.; Zaver, S.A.; Gowen, B.G.; Wyman, S.K.; Garelis, N.E.; Onia, L.; McWhirter, S.M.; Katibah, G.E.; Corn, J.E.; Woodward, J.J.; et al. SLC19A1 Transports Immunoreactive Cyclic Dinucleotides. Nature 2019, 573, 434–438. [Google Scholar] [CrossRef]
  292. Morehouse, B.R.; Govande, A.A.; Millman, A.; Keszei, A.F.A.; Lowey, B.; Ofir, G.; Shao, S.; Sorek, R.; Kranzusch, P.J. STING Cyclic Dinucleotide Sensing Originated in Bacteria. Nature 2020, 586, 429–433. [Google Scholar] [CrossRef]
  293. Cluff, C.W. Monophosphoryl Lipid A (MPL) as an Adjuvant for Anti-Cancer Vaccines: Clinical Results. Adv. Exp. Med. Biol. 2010, 667, 111–123. [Google Scholar] [CrossRef] [PubMed]
  294. Haghshenas, M.R.; Mousavi, T.; Kheradmand, M.; Afshari, M.; Moosazadeh, M. Efficacy of Human Papillomavirus L1 Protein Vaccines (Cervarix and Gardasil) in Reducing the Risk of Cervical Intraepithelial Neoplasia: A Meta-Analysis. Int. J. Prev. Med. 2017, 8, 44. [Google Scholar] [CrossRef] [PubMed]
  295. Fabrizi, F.; Cerutti, R.; Nardelli, L.; Tripodi, F.; Messa, P. HBV Vaccination with Fendrix Is Effective and Safe in Pre-Dialysis CKD Population. Clin. Res. Hepatol. Gastroenterol. 2020, 44, 49–56. [Google Scholar] [CrossRef] [PubMed]
  296. Coccia, M.; Collignon, C.; Hervé, C.; Chalon, A.; Welsby, I.; Detienne, S.; van Helden, M.J.; Dutta, S.; Genito, C.J.; Waters, N.C.; et al. Cellular and Molecular Synergy in AS01-Adjuvanted Vaccines Results in an Early IFNγ Response Promoting Vaccine Immunogenicity. NPJ Vaccines 2017, 2, 25. [Google Scholar] [CrossRef] [PubMed]
  297. Cappello, V.; Marchetti, L.; Parlanti, P.; Landi, S.; Tonazzini, I.; Cecchini, M.; Piazza, V.; Gemmi, M. Ultrastructural Characterization of the Lower Motor System in a Mouse Model of Krabbe Disease. Sci. Rep. 2016, 6, 1. [Google Scholar] [CrossRef] [PubMed]
  298. Miao, E.A.; Andersen-Nissen, E.; Warren, S.E.; Aderem, A. TLR5 and Ipaf: Dual Sensors of Bacterial Flagellin in the Innate Immune System. Semin. Immunopathol. 2007, 29, 275–288. [Google Scholar] [CrossRef] [PubMed]
  299. Cui, B.; Liu, X.; Fang, Y.; Zhou, P.; Zhang, Y.; Wang, Y. Flagellin as a Vaccine Adjuvant. Expert Rev. Vaccines 2018, 17, 335–349. [Google Scholar] [CrossRef] [PubMed]
  300. Hajam, I.A.; Dar, P.A.; Shahnawaz, I.; Jaume, J.C.; Lee, J.H. Bacterial Flagellin-a Potent Immunomodulatory Agent. Exp. Mol. Med. 2017, 49, e373. [Google Scholar] [CrossRef]
  301. Kayamuro, H.; Yoshioka, Y.; Abe, Y.; Arita, S.; Katayama, K.; Nomura, T.; Yoshikawa, T.; Kubota-Koketsu, R.; Ikuta, K.; Okamoto, S.; et al. Interleukin-1 Family Cytokines as Mucosal Vaccine Adjuvants for Induction of Protective Immunity against Influenza Virus. J. Virol. 2010, 84, 12703–12712. [Google Scholar] [CrossRef] [PubMed]
  302. Tovey, M.G.; Lallemand, C. Safety, Tolerability, and Immunogenicity of Interferons. Pharmaceuticals 2010, 3, 1162–1186. [Google Scholar] [CrossRef] [PubMed]
  303. Mohan, T.; Zhu, W.; Wang, Y.; Wang, B.-Z. Applications of Chemokines as Adjuvants for Vaccine Immunotherapy. Immunobiology 2018, 223, 477–485. [Google Scholar] [CrossRef] [PubMed]
  304. Gao, Y.; Guo, Y. Research Progress in the Development of Natural-Product-Based Mucosal Vaccine Adjuvants. Front. Immunol. 2023, 14, 1152855. [Google Scholar] [CrossRef] [PubMed]
  305. Khademi, F.; Taheri, R.-A.; Yousefi Avarvand, A.; Vaez, H.; Momtazi-Borojeni, A.A.; Soleimanpour, S. Are Chitosan Natural Polymers Suitable as Adjuvant/Delivery System for Anti-Tuberculosis Vaccines? Microb. Pathog. 2018, 121, 218–223. [Google Scholar] [CrossRef] [PubMed]
  306. Petrovsky, N.; Cooper, P.D. AdvaxTM, a Novel Microcrystalline Polysaccharide Particle Engineered from Delta Inulin, Provides Robust Adjuvant Potency Together with Tolerability and Safety. Vaccine 2015, 33, 5920–5926. [Google Scholar] [CrossRef]
  307. Sheikh-Mohamed, S.; Isho, B.; Chao, G.Y.C.; Zuo, M.; Cohen, C.; Lustig, Y.; Nahass, G.R.; Salomon-Shulman, R.E.; Blacker, G.; Fazel-Zarandi, M.; et al. Systemic and Mucosal IgA Responses Are Variably Induced in Response to SARS-CoV-2 mRNA Vaccination and Are Associated with Protection against Subsequent Infection. Mucosal Immunol. 2022, 15, 799–808. [Google Scholar] [CrossRef]
  308. Huang, N.; Pérez, P.; Kato, T.; Mikami, Y.; Okuda, K.; Gilmore, R.C.; Conde, C.D.; Gasmi, B.; Stein, S.; Beach, M.; et al. SARS-CoV-2 Infection of the Oral Cavity and Saliva. Nat. Med. 2021, 27, 892–903. [Google Scholar] [CrossRef]
  309. Mostaghimi, D.; Valdez, C.N.; Larson, H.T.; Kalinich, C.C.; Iwasaki, A. Prevention of Host-to-Host Transmission by SARS-CoV-2 Vaccines. Lancet Infect. Dis. 2022, 22, e52–e58. [Google Scholar] [CrossRef]
  310. Bain, C.C.; Cerovic, V. Interactions of the Microbiota with the Mucosal Immune System. Immunology 2020, 159, 1–3. [Google Scholar] [CrossRef]
  311. Lin, D.; Yang, L.; Wen, L.; Lu, H.; Chen, Q.; Wang, Z. Crosstalk between the Oral Microbiota, Mucosal Immunity, and the Epithelial Barrier Regulates Oral Mucosal Disease Pathogenesis. Mucosal Immunol. 2021, 14, 1247–1258. [Google Scholar] [CrossRef]
  312. Kayama, H.; Okumura, R.; Takeda, K. Interaction Between the Microbiota, Epithelia, and Immune Cells in the Intestine. Annu. Rev. Immunol. 2020, 38, 23–48. [Google Scholar] [CrossRef]
  313. Roubidoux, E.K.; Schultz-Cherry, S. Animal Models Utilized for the Development of Influenza Virus Vaccines. Vaccines 2021, 9, 787. [Google Scholar] [CrossRef]
  314. Nguyen, T.-Q.; Rollon, R.; Choi, Y.-K. Animal Models for Influenza Research: Strengths and Weaknesses. Viruses 2021, 13, 1011. [Google Scholar] [CrossRef]
  315. Song, Y.; Wang, X.; Zhang, H.; Tang, X.; Li, M.; Yao, J.; Jin, X.; Ertl, H.C.J.; Zhou, D. Repeated Low-Dose Influenza Virus Infection Causes Severe Disease in Mice: A Model for Vaccine Evaluation. J. Virol. 2015, 89, 7841–7851. [Google Scholar] [CrossRef]
  316. Izurieta, H.S. Adverse Events Reported Following Live, Cold-Adapted, Intranasal Influenza Vaccine. JAMA 2005, 294, 2720. [Google Scholar] [CrossRef]
  317. Glezen, W. Asthma, Influenza, and Vaccination. J. Allergy Clin. Immunol. 2006, 118, 1199–1206. [Google Scholar] [CrossRef]
  318. Mosley, J.F.; Smith, L.L.; Brantley, P.; Locke, D.; Como, M. Vaxchora: The First FDA-Approved Cholera Vaccination in the United States. Pharm. Ther. 2017, 42, 638–640. [Google Scholar]
  319. Luquero, F.J.; Grout, L.; Ciglenecki, I.; Sakoba, K.; Traore, B.; Heile, M.; Dialo, A.A.; Itama, C.; Serafini, M.; Legros, D.; et al. First Outbreak Response Using an Oral Cholera Vaccine in Africa: Vaccine Coverage, Acceptability and Surveillance of Adverse Events, Guinea, 2012. PLoS Negl. Trop. Dis. 2013, 7, e2465. [Google Scholar] [CrossRef] [PubMed]
  320. McNeil, M.M.; Paradowska-Stankiewicz, I.; Miller, E.R.; Marquez, P.L.; Seshadri, S.; Collins, L.C.; Cano, M.V. Adverse Events Following Adenovirus Type 4 and Type 7 Vaccine, Live, Oral in the Vaccine Adverse Event Reporting System (VAERS), United States, October 2011–July 2018. Vaccine 2019, 37, 6760–6767. [Google Scholar] [CrossRef] [PubMed]
  321. Moro, P.L.; Perez-Vilar, S.; Lewis, P.; Bryant-Genevier, M.; Kamiya, H.; Cano, M. Safety Surveillance of Diphtheria and Tetanus Toxoids and Acellular Pertussis (DTaP) Vaccines. Pediatrics 2018, 142, e20174171. [Google Scholar] [CrossRef] [PubMed]
  322. Song, B.J.; Katial, R.K. Update on Side Effects from Common Vaccines. Curr. Allergy Asthma Rep. 2004, 4, 447–453. [Google Scholar] [CrossRef] [PubMed]
  323. Platt, L.R.; Estivariz, C.F.; Sutter, R.W. Vaccine-Associated Paralytic Poliomyelitis: A Review of the Epidemiology and Estimation of the Global Burden. J. Infect. Dis. 2014, 210, S380–S389. [Google Scholar] [CrossRef] [PubMed]
  324. Lemiale, F.; Kong, W.; Akyürek, L.M.; Ling, X.; Huang, Y.; Chakrabarti, B.K.; Eckhaus, M.; Nabel, G.J. Enhanced Mucosal Immunoglobulin A Response of Intranasal Adenoviral Vector Human Immunodeficiency Virus Vaccine and Localization in the Central Nervous System. J. Virol. 2003, 77, 10078–10087. [Google Scholar] [CrossRef] [PubMed]
  325. Van Ginkel, F.W.; Jackson, R.J.; Yuki, Y.; McGhee, J.R. Cutting Edge: The Mucosal Adjuvant Cholera Toxin Redirects Vaccine Proteins into Olfactory Tissues. J. Immunol. 2000, 165, 4778–4782. [Google Scholar] [CrossRef] [PubMed]
  326. 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]
  327. Banerjee, S. Safety and Efficacy of Low Dose Escherichia coli Enterotoxin Adjuvant for Urease Based Oral Immunisation against Helicobacter pylori in Healthy Volunteers. Gut 2002, 51, 634–640. [Google Scholar] [CrossRef] [PubMed]
  328. Marie, C.; Ali, A.; Chandwe, K.; Petri, W.A.; Kelly, P. Pathophysiology of Environmental Enteric Dysfunction and Its Impact on Oral Vaccine Efficacy. Mucosal Immunol. 2018, 11, 1290–1298. [Google Scholar] [CrossRef]
  329. Mori, F.; Barni, S.; Liccioli, G.; Novembre, E. Oral Immunotherapy (OIT): A Personalized Medicine. Medicina 2019, 55, 684. [Google Scholar] [CrossRef]
  330. Thomas, H.C.; Parrott, M.V. The Induction of Tolerance to a Soluble Protein Antigen by Oral Administration. Immunology 1974, 27, 631–639. [Google Scholar]
  331. Hartwell, B.L.; Pickens, C.J.; Leon, M.; Northrup, L.; Christopher, M.A.; Griffin, J.D.; Martinez-Becerra, F.; Berkland, C. Soluble Antigen Arrays Disarm Antigen-Specific B Cells to Promote Lasting Immune Tolerance in Experimental Autoimmune Encephalomyelitis. J. Autoimmun. 2018, 93, 76–88. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mucosa structures and immunity mechanisms. Mucosal immune responses are composed of multiple immune effector mechanisms with a complex network of innate and adaptive immune components and play critical roles in defending against invasive pathogens at host mucosal barriers. The first defense line of mucosal barriers is epithelial cells, which have tight cell junctions serving as physical barriers, together with microbiota and antimicrobial peptides in the lumen as biological and biochemical barriers, respectively. Ciliated cells in the respiratory tract epithelium are important for mucociliary clearance by propelling pathogens and particles out of the airways. Mucosal epithelium can be defined into different types: nonkeratinized epithelium, columnar epithelium, and squamous epithelium (the squamous epithelial cells in the respiratory tract and ends of the GI tract are omitted from the figure) according to their structures and functions. Pathogen surveillance is mediated by M cells (antigen transport) and DCs (antigen processing and transport) at mucosal barriers. Antigens are captured and processed by DCs and macrophages, leading to the activation and maturation of those antigen-presenting cells (APCs). Antigen-stimulated DCs later initiate downstream immune responses by migrating to lymph nodes to present antigens on MHC molecules to T-cells, which further mediate expansion of naïve T-cells into differentiated T helper subsets (Th1, Th2, Th17, Treg) via regulation of transcription factors and secretion of lineage-defining cytokines. Activated T helper subsets then perform their functions, including upregulation of polymeric Ig receptor (pIgR) expression and promotion of B-cell differentiation into plasma cells, with IgA class switching to produce secretory IgA (SIgA) to intercept the pathogens at the mucosa. Soluble factors (BAFF, APRIL) secreted by DCs, innate lymphoid cells (ILCs), and epithelial cells could also enhance IgA production by T-cell independent class switching. IgG, derived from local B-cells or from blood, is also present in mucosal tissues to directly neutralize pathogens or mediate cytotoxicity by recognition of Fc receptor expressed on nature killer (NK) cells. In addition, antigen-specific cytotoxic T-cells (CTLs) also enter the mucosa to kill those infected cells. (Figure created with BioRender.com).
Figure 1. Mucosa structures and immunity mechanisms. Mucosal immune responses are composed of multiple immune effector mechanisms with a complex network of innate and adaptive immune components and play critical roles in defending against invasive pathogens at host mucosal barriers. The first defense line of mucosal barriers is epithelial cells, which have tight cell junctions serving as physical barriers, together with microbiota and antimicrobial peptides in the lumen as biological and biochemical barriers, respectively. Ciliated cells in the respiratory tract epithelium are important for mucociliary clearance by propelling pathogens and particles out of the airways. Mucosal epithelium can be defined into different types: nonkeratinized epithelium, columnar epithelium, and squamous epithelium (the squamous epithelial cells in the respiratory tract and ends of the GI tract are omitted from the figure) according to their structures and functions. Pathogen surveillance is mediated by M cells (antigen transport) and DCs (antigen processing and transport) at mucosal barriers. Antigens are captured and processed by DCs and macrophages, leading to the activation and maturation of those antigen-presenting cells (APCs). Antigen-stimulated DCs later initiate downstream immune responses by migrating to lymph nodes to present antigens on MHC molecules to T-cells, which further mediate expansion of naïve T-cells into differentiated T helper subsets (Th1, Th2, Th17, Treg) via regulation of transcription factors and secretion of lineage-defining cytokines. Activated T helper subsets then perform their functions, including upregulation of polymeric Ig receptor (pIgR) expression and promotion of B-cell differentiation into plasma cells, with IgA class switching to produce secretory IgA (SIgA) to intercept the pathogens at the mucosa. Soluble factors (BAFF, APRIL) secreted by DCs, innate lymphoid cells (ILCs), and epithelial cells could also enhance IgA production by T-cell independent class switching. IgG, derived from local B-cells or from blood, is also present in mucosal tissues to directly neutralize pathogens or mediate cytotoxicity by recognition of Fc receptor expressed on nature killer (NK) cells. In addition, antigen-specific cytotoxic T-cells (CTLs) also enter the mucosa to kill those infected cells. (Figure created with BioRender.com).
Vaccines 12 00191 g001
Figure 2. Different types of mucosal epithelium structures and functions. The oral cavity, esophagus, vagina, and anus have nonkeratinized epithelium composed of several layers of cells. The columnar epithelium is found in the stomach, intestines, and respiratory tract, except for the alveoli and gallbladder, which are composed of tall, narrow cells that are tightly packed together. Squamous epithelium is a type of epithelium found in the lungs, blood vessels, and body cavities, which are composed of flat, scale-like cells that are tightly packed together. (Figure created with BioRender.com).
Figure 2. Different types of mucosal epithelium structures and functions. The oral cavity, esophagus, vagina, and anus have nonkeratinized epithelium composed of several layers of cells. The columnar epithelium is found in the stomach, intestines, and respiratory tract, except for the alveoli and gallbladder, which are composed of tall, narrow cells that are tightly packed together. Squamous epithelium is a type of epithelium found in the lungs, blood vessels, and body cavities, which are composed of flat, scale-like cells that are tightly packed together. (Figure created with BioRender.com).
Vaccines 12 00191 g002
Table 2. List of vaccine adjuvants licensed or in clinical trials.
Table 2. List of vaccine adjuvants licensed or in clinical trials.
Adjuvant NameType/CompositionImmune Stimulation FunctionDevelopment StagesDisease TypeReferences
Cholera Toxin (CT) *Recombinant detoxified bacterial A–B subunitGangliosides binding; cAMP stimulation; DC activation; mucosal IgA and Th productionlicensed for human use (Dukoral®)Vibrio cholerae[41,103,255,256]
AlumAluminum mineral salt; Insoluble particulates of hydroxide, phosphate, or hydroxyphosphate sulfate saltsAntigen adsorption and controlled release; Humoral immunity modulations; Induction of Th2 immune responses and inflammationlicensed for human use (Daptacel®, Twinrix®, Gardasil®, Bexsero®, Prevnar 20®)HAV, HBV, HPV, Diphtheria, Tetanus, Haemophilus influenza, Meningococcal, Pneumococcal[254,255]
MF59, AS03 *Oil-in-water (squalene in water) emulsionAPC stimulation and activation; Modulation of humoral and cellular immune responseslicensed for human use (Fluad®, Pandemrix®, Arepanrix™)Influenza (pandemic and seasonal)[254,255,257,258,259]
CpG ODN (e.g., 1018 ISS)OligonucleotideSoluble TLR9 ligand; Increase both humoral and cellular immune responses; Th1-specific cytokine expression and CD8+ T-cell activitylicensed for human use (Heplisav-B™, CorbeVax™)HBV, SARS-CoV-2[103,254,258]
MPL *, AS04 *Non-toxic derivative of LPS (MPL); Alum-adsorbed MPL (AS04)TLR4 agonist; Increase APC maturation; Induction of Th1 type immunity and improve both humoral and cellular immune responseslicensed for human use (Cervarix™, Fendrix™)HBV, HPV[103,254,255,258,259]
VLP *, Virosomes *, Liposomes (IRIV) *Liposomes, viral proteins, and nanoparticlesPAMP signals; Improved APC antigen uptake to promote T-cell and B-cell activationlicensed for human use (Epaxal®, Engerix®-B, Gardasil®, Cervarix™)HBV, HPV, HAV, Influenza[254,255,257,259]
Alhydroxiquim-IIAlum adsorbed TLR7/8 agonistStimulate TLR7 and TLR8; Lymph node-specific delivery and antigen releaselicensed for human use (COVAXIN®)SARS-CoV-2[254]
QS-21Purified plant extract containing water-soluble triterpene glycosides (saponins)Activation of strong humoral and cellular immune responses; stimulation of cytotoxic T-cell responselicensed for human use (Shingrix®, Arexvy®, Mosquirix®)Varicella-zoster virus, RSV, Malaria[260,261]
AS01 *Liposomes containing MPLA and saponin QS-21TLR4 ligand; Induction of Th1 type and CD8+ T-cell mediated immunity; Augment of antibody responsesPhase IIIMalaria, (RTS,S) and herpes zoster vaccine (HZ/su)[262]
EGVac systemBacterial polysaccharide and DNAStimulation of both B-cell and T-cell immune responsesPhase IIHPV[258,263]
Saponin complexes *, ISCOM *, QS-21 *, Matrix-M *Nanocomplex of lipid, pure saponins, and cholesterolInduction of humoral immunity; Th1, Th2 and CD8+ T-cell responsesPhase IInfluenza[258,259,264,265,266]
ISA51Oil-in-water emulsion stabilized with non-ionic surfactantAugments antibody responses and T-cell activityPhase IIInfluenza (seasonal)[267,268]
VAX2012Q, VAX125TLR5 agonist, bacterial flagellin fused antigenPromotion of antibody titer and T-cell immunityPhase IIInfluenza[258,269]
Ampligen® *, rintatolimod, PIKAPoly I:C, double-stranded (ds)-RNA polymer analog, TLR3 agonistEnhanced humoral responses and Th1; Th17 responsesPhase I and Phase IIIInfluenza, Rabies, HIV-1[254,270,271,272]
Matrix-M *Protein nanoparticleImproved antibody responses and T-cell activitiesPhase IISARS-CoV-2[273]
GM-CSF *Granulocyte-macrophage colony-stimulating factor, cytokineEnhanced mucosal IgA responses; activation of DCs and T-cellsPhase I and Phase IIHIV-1, SARS-CoV-2[274,275]
IFN-α *Type 1 interferon (IFN)Induction of secretory IgA (sIgA) response in salivaPhase IInfluenza[276]
CCL3 *ChemokineSecretion of mucosal IgA and CTLPre-clinicalHIV-1[270]
α-GalCer *GlycosphingolipidsActivation of pattern recognition receptor; induction of IgA responses and Th1/CTL responsesPhase IIHBV[277]
Chitosan *Natural-product-based particulate-polysaccharides,Enhanced mucosal T-cell responses; induction of proinflammatory cytokines and secretory IgA responses and Th2 responses; PRR activation; upregulation of co-stimulatory molecules; activation of complement pathwaysPhase IINorovirus[278]
*: Adjuvants developed for mucosal vaccines.
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

Song, Y.; Mehl, F.; Zeichner, S.L. Vaccine Strategies to Elicit Mucosal Immunity. Vaccines 2024, 12, 191. https://doi.org/10.3390/vaccines12020191

AMA Style

Song Y, Mehl F, Zeichner SL. Vaccine Strategies to Elicit Mucosal Immunity. Vaccines. 2024; 12(2):191. https://doi.org/10.3390/vaccines12020191

Chicago/Turabian Style

Song, Yufeng, Frances Mehl, and Steven L. Zeichner. 2024. "Vaccine Strategies to Elicit Mucosal Immunity" Vaccines 12, no. 2: 191. https://doi.org/10.3390/vaccines12020191

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