Friday 11 November 2022

Engineering Antiviral Vaccines

La terra è un bel posto peccato ci siano gli umani...

The earth is a beautiful place, too bad there are humans...

Source:

https://pubs.acs.org/doi/10.1021/acsnano.0c06109

Engineering Antiviral Vaccines

Xingwu Zhou,● Xing Jiang,● Moyuan Qu,● George E. Aninwene, II, Vadim Jucaud, James J. Moon,

Zhen Gu, Wujin Sun,* and Ali Khademhosseini*


Cite This: ACS Nano 2020, 14, 12370−12389 Read Online


ACCESS Metrics & More Article Recommendations


ABSTRACT: Despite the vital role of vaccines in fighting viral pathogens, effective

vaccines are still unavailable for many infectious diseases. The importance of vaccines

cannot be overstated during the outbreak of a pandemic, such as the coronavirus disease

2019 (COVID-19) pandemic. The understanding of genomics, structural biology, and

innate/adaptive immunity have expanded the toolkits available for current vaccine

development. However, sudden outbreaks and the requirement of population-level

immunization still pose great challenges in today’s vaccine designs. Well-established

vaccine development protocols from previous experiences are in place to guide the

pipelines of vaccine development for emerging viral diseases. Nevertheless, vaccine

development may follow different paradigms during a pandemic. For example, multiple

vaccine candidates must be pushed into clinical trials simultaneously, and manufacturing

capability must be scaled up in early stages. Factors from essential features of safety,

efficacy, manufacturing, and distributions to administration approaches are taken into

consideration based on advances in materials science and engineering technologies. In

this review, we present recent advances in vaccine development by focusing on vaccine discovery, formulation, and delivery

devices enabled by alternative administration approaches. We hope to shed light on developing better solutions for faster and

better vaccine development strategies through the use of biomaterials, biomolecular engineering, nanotechnology, and

microfabrication techniques.

KEYWORDS: COVID-19, pandemics, infectious disease, vaccine, immunotherapy, drug discovery, drug delivery, biomedical devices


T


he appearance of severe acute respiratory syndrome-

coronavirus 2 (SARS-CoV-2) has caused unprece-

dented global disruption to health, economy, and

social stability.1 SARS-CoV-2 causes coronavirus disease 2019

(COVID-19) in which individuals can suffer from mild to

severe symptoms once infected.2 Global scientific communities

are working together to fight the crisis caused by this

pandemic. The initial efforts to combat this pandemic included

identifying infected patients and ramping up clinical trials for

repurposing existing drugs.3 However, the development of

effective vaccines to prevent SARS-CoV-2 infection is believed

by many to be the most effective long-term solution.4

The pathway from identifying a virus to having a vaccine is

lengthy and expensive. After billions of dollars spent and

multiple years of trials, researchers generally face high rates of

failure in the traditional vaccine development paradigms

(Figure 1a).4 Despite rapid responses from the scientific

community and the pharmaceutical industry, vaccines have

historically not been ready even after an epidemic becomes

manageable, such as severe acute respiratory syndrome

(SARS), Zika, and Ebola.5 Considering economic factors,

funding for vaccine development may be reallocated after a

substantial drop in infection cases is observed, as in the 2015−


2016 Zika and SARS epidemics. The 2014−2016 Ebola

outbreak in Africa resulted in over 28,000 cases and 11,000

deaths.6 After continuous research funding support, in

December 2019, the FDA granted the first approval to Ervebo

(a single-dose, live attenuated virus vaccine from Merck) for

Ebola virus prevention.7 Benefiting from egg- and cell-based

platforms, vaccine approval for the most recent 2009 H1N1

pandemic was relatively fast following the peak of the outbreak,

and the vaccine was ultimately incorporated in the seasonal

influenza vaccine.8 Since the initiation and ending of epidemics

are highly unpredictable, the rapid deployment of vaccine

development may still be too slow for a sudden outbreak.4,5

To tackle a pandemic, a different vaccine development

process is needed to rapidly produce safe and effective vaccines

against novel viruses. This process is characterized by

overlapping phases, multiple potential vaccine candidates,


Received: July 21, 2020

Accepted: September 18, 2020

Published: October 1, 2020


Review


www.acsnano.org


© 2020 American Chemical Society


12370


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


This article is made available via the ACS COVID-19 subset for unrestricted RESEARCH re-use

and analyses in any form or by any means with acknowledgement of the original source.

These permissions are granted for the duration of the World Health Organization (WHO)

declaration of COVID-19 as a global pandemic.


Downloaded via 95.145.163.132 on November 9, 2022 at 06:37:05 (UTC).

See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



and early ramp-up of manufacturing capacities (Figure 1a).4

After the initial release of the genetic sequence of SARS-CoV-

2, researchers worldwide started to develop vaccines for the

prevention of COVID-19. The first dose was administered into

humans by Moderna Inc. on March 16, 2020.9 This method

utilized lipid nanoparticles (LNPs) with a formulated mRNA

vaccine.9 This work was fast-tracked at an unprecedented pace,

even though mRNA-based vaccines have never been licensed

before.9 A snapshot of the landscape of COVID-19 vaccines

(Figure 1b) shows that around 31 vaccines have entered

clinical trials as of August 25, 2020.10 A high diversity of

vaccine candidates is seen in these forerunners, which includes

six mRNA-based, five inactivated virus-based, five nonreplicat-

ing viral vector-based, one replicating viral vector-based, four

DNA-based, and one cell culture-based vaccines as well as nine

protein-based vaccines (Figure 1b,c). The incomplete under-

standing of the SARS-CoV-2 interactions with the human

immune system became one of the major obstacles in vaccine

development. A recent report on a second COVID-19

infection in the same patient further raises the question of

protective immunity and immune memory. For instance, it is

still unknown how durable the immune memory response

induced by SARS-CoV-2 could be or what is the threshold of

antibodies titers that can protect the patients against

reinfection.11,12 In addition, protective immunity against

SARS-CoV-2 should also be comprehensively studied, where

the balance of cellular and humoral immunity, the ratio of

effector to memory T cells, the maintenance of memory B

cells, and functional features of activated T cells should be


characterized.11,12 These aspects could ultimately facilitate

vaccine designs and evaluations.

In this review, we aim to provide engineering insights on

vaccine development by covering vaccine discovery, vaccine

formulations in terms of different materials (lipid, polymer, and

inorganic particles), and vaccine delivery devices. We will

present how some of the existing challenges could potentially

be addressed by nano/micro/macroscale engineering ap-

proaches. We will also discuss some emerging technologies

that may contribute to future vaccine development pipelines.


VACCINE DISCOVERY

After a virus is identified, different categories of vaccines can be

developed, which includes virus-, viral vector-, nucleic acid-,

protein/peptide-, and/or cell-based vaccines (Figure 1c).13

Virus-based vaccines commonly require patient-derived

viruses, while other types can utilize the viral genomic

sequences to accelerate vaccine development. Previously

approved vaccines mostly fall into the virus-based and the

recombinant protein-based categories, but not into the nucleic

acid-based one (Figure 2a).14 The discovery process of each

vaccine type is distinct with corresponding pros and cons

(Figure 2b−d).

Virus-Based Vaccines. Virus-based vaccines, including the

weakened and inactivated types, are still the most widely used

to date. The effective prevention of the polio epidemic by

vaccination in the 20th century represented a landmark success

in medical history.18 This vaccination regimen included the use

of the inactivated polio vaccine (IPV) and oral polio vaccine


Figure 1. Summary of vaccine development paradigms and major types of vaccines. (a) Paradigms of vaccine development in a traditional

condition vs during a pandemic.5 (b) Snapshot of the vaccine landscape for COVID-19 with the different types of vaccine candidates in

various clinical stages as of August 25, 2020.10 (c) Schematic of major types of vaccines, including virus-based (weakened and inactivated),

viral vector-based (replicating and nonreplicating), nucleic acid-based (DNA and mRNA), protein and peptide-based, and cell-based.13


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12371



(OPV).18 The IPV was made by inactivating lab-cultured

poliovirus by formalin. The OPV was weakened live polio-

virus.18 Weakened viruses are highly immunogenic and are

commonly developed by the deletion of viral genes. Never-

theless, mutations could cause virulence reversion as observed

for vaccine-derived polioviruses that have been responsible for

recent polio outbreaks.19 This represents an obstacle for polio

eradication and, more importantly, a concern for the weakened

virus-based vaccines in general. Further genetic manipulation

strategies are therefore required to prevent this issue. For

instance, modifying the 5′ untranslated region (UTR), 2C

coding region, and 3D polymerase region resulted in a

genetically stabilized OPV strain that was less likely to regain

virulence and exhibited limited viral adaptability.20


On the contrary, inactivated virus vaccines generally do not

have virulence reversion issues. They also do not rely on

extensive genetic manipulations, but the inactivation process

may cause the loss of antigenicity of viral immunogens. This

vaccine type has been quickly adapted to cope with COVID-

19, where patient-derived viruses were inactivated with β-

propiolactone, followed with purification by chromatography,

and formulated with Al(OH)3 as the adjuvant.15 The vaccine

protected nonhuman primates from SARS-CoV-2 challenges

with no notable side effects.15 Advantages of these virus-based

methods include well-developed methodologies, well-estab-

lished regulation policies, and existing manufacturing facilities

from previous practices. Nevertheless, the requirement of virus

seeding and culture along with unknown safety issues


Figure 2. Vaccine discovery processes for different types of vaccines. (a) Timeline of previously approved antiviral vaccines.14 (b) Virus-

based vaccines require the seeding and culture of specific viruses derived from patients. By manipulating the genetic sequences of the virus

or inactivating the virus directly with chemicals, weakened and inactivated virus vaccines can be produced.15 (c) Viral vector-based and

nucleic acid-based vaccines are independent of virus culture and rely on the genetic sequence of the virus and/or the selection of

immunogenic sequences of the virus. By selecting commonly developed vectors, virus-specific sequences can be inserted.16 Nucleic acid

vaccines require further formulations for optimal efficacy. (d) Structure-based understanding of native viral proteins can validate the

expressed recombinant proteins and predict the peptide sequences desirable as vaccines. Protein/peptide vaccines require further

formulations for optimal efficacy.17


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12372



necessitates alternative practices for the development of other

vaccines.

Viral Vector-Based Vaccines. Viral vector-based vaccines

work by inserting viral antigen-encoding DNA sequences into

host cells, which leads to the expression of viral antigens by the

host cells.16 The rationale is based on the capability of viruses

to infect cells efficiently through viral genome integration

mechanisms.16 This method could also be problematic since

the genome of the host could be altered and cause other

diseases. The selection of viruses is therefore crucial in

balancing the efficacy (efficient infectivity) and safety (limited

pathogenicity). Another common obstacle is the host’s pre-

existing immunity to certain vectors that would mitigate the

efficacy of a vaccine. The selected viral vectors could be further

genetically modified to be nonreplicable by disrupting genes

responsible for replication. Adenoviruses (Ad) are the most

extensively used vectors for vaccines because of several

advantages: infectious to various cell types, efficient transgene

expression, high in vitro growth, lack of genome integration,

and genetic stability.21 For instance, Johnson & Johnson (J &

J) deployed the Ad26 vector to deliver gene encoding surface

proteins of the SARS-CoV-2, and CanSino Biological used the

Ad5 vector.22 The spike (S) glycoprotein of SARS-CoV-2 was

cloned with the tissue plasminogen activator signal peptide

gene in the early region 1 (E1) and early region 3 (E3) deleted

Ad5 vector. The E1 and E3 deletions render the Ad5 vector

nonreplicating. Even though rapid humoral and T cell

responses were induced by the vaccine, pre-existing anti-Ad5

immunity partially diminished the immune responses among

healthy adults.22 To address the pre-existing immunity to Ad,

chimpanzee Ad (ChAdOX1) represents an alternative solution

that provided broad protective immunity against the Middle

East Respiratory Syndrome Coronavirus (MERS-CoV).23 An

investigational vaccine for SARS-CoV-2 from the University of

Oxford is based on this vector.23 Their preclinical study

showed the prevention of SARS-CoV-2 pneumonia in rhesus

macaques with no evidence of subsequent immune-enhanced

diseases.23 However, viral loads were more significantly

reduced in the lower respiratory tract but not in the nasal

swabs. In general, because of previous vaccine programs, the

manufacturing capacity and protocols for this method are

already well-established; therefore, less effort may be needed to

shift toward the large-scale production of vector-based

vaccines. However, both the pre-existing immunity toward

certain vectors and previously demonstrated safety concerns

over the increased risk of human immunodeficient virus type 1

(HIV-1) acquisition should be carefully evaluated.22

Nucleic Acid-Based Vaccines. DNA- and mRNA-based

vaccines have the greatest potential for rapid development

because of their synthetic nature that circumvents the need for

cell culture or fermentation of viruses. The synthetic nature of

this method also results in a high safety profile and rapid

manufacturing ability.24 However, this method generally

requires further delivery formulations or better delivery devices

to boost vaccine efficacy.24 In contrast to RNA, the increased

stability of DNA decreases the need for frozen storage and low-

temperature transportation.24 One of the forerunners among

COVID-19 vaccines is DNA-based (Inovio).10 Similar DNA-

based vaccines have had promising results for the prevention of

MERS.25 In the discovery phase of DNA-based vaccines,

strategies such as codon/RNA optimization and the addition

of highly efficient immunoglobulin leader sequences are

needed to enhance the magnitude and breadth of the immune


response.26 For instance, Muthumani et al. designed a

consensus DNA sequence for the S protein by analyzing S

protein sequences, where sequences from clades A and B can

be both involved.26 To enhance the in vivo expression, the

immunoglobulin E (IgE) leader sequence was added to the

immunogen sequence. The insert was further incorporated into

the pVax1 vector, which allowed a high-level transient

expression of proteins of interest in mammalian cells.26 The

strategies of adding the IgE leader sequence and the use of the

pVax1 vector could be quickly adapted to the different virus

sequences as well.

Due to their high potency, rapid development, and low cost

for manufacturing, mRNA vaccines have emerged as a

promising alternative vaccination development strategy for

various infectious diseases and cancers.27 No mRNA-based

vaccine is currently FDA-approved. Several strategies have

been proposed to solve the common drawbacks of mRNA-

based vaccines, which include mRNA instability and high

innate immunogenicity.28 The 5′ and 3′ UTR of mRNA has

significant regulatory influences on both stability and trans-

latability.28 Half-life and expression can be enhanced when

these sequences come from viral genes: A 5′ cap is essential for

efficient protein expression, and a poly(A) tail affects mRNA

translation and stability.29 Furthermore, the replacement of

rare codons with frequently used synonymous codons also

enhances protein expression.30 Additional optimization strat-

egies include increasing G:C content and introducing modified

nucleosides.31,32 In spite of the innate immunostimulatory

effect of exogenous or unpurified mRNA, further investigations

are required to assess the advantages of their use for vaccines

specifically.33 Innate immune sensing mechanisms of mRNA

have resulted in the inhibition of antigen expression and a

decreased immune response.34 Therefore, further formulations

for vaccine delivery could potentially reduce innate mRNA-

associated immunogenicity and favor immune activation.

Subunit Protein/Peptide Vaccines. Unlike nucleic acid-

based vaccines, which produce viral protein after admin-

istration, the direct application of viral protein antigens may be

a more straightforward method to trigger immunity. This

approach currently represents the largest category of COVID-

19 vaccines under preclinical investigation.4,10 Full-length

proteins are advantageous because they can induce the

development of antibodies against multiple epitopes. Impor-

tantly, there is a higher probability that the developed

antibodies could bind to the native conformation of the viral

protein. However, this also increases the chance of inducing

nonspecific cross-reactive antibodies.35 In addition to the high

cost of recombinant protein technologies, the recombinant

protein products may not fully match all the molecular features

of the corresponding viral proteins, which may lead to

ineffective induction of broad neutralizing antibodies

(bNAb).35 Optimizing peptide sequences in the discovery

phase can improve the stabilization of epitopes associated with

bNAb while reducing the antigenicity of non-NAb epitopes.17

Given that only specific amino acid sequences of the full-

length protein antigens are responsible for effective immune

responses, minimally immunogenic peptides, mimicking B and

T cell epitopes, have been proposed for vaccines.17 Early

approaches used short peptide sequences (8−10 amino acids)

that could potentially be presented by major histocompatibility

complex (MHC) I molecules. Importantly, the memory

immune response of cytotoxic T cells could not be efficiently

induced with such T cell epitope-mimicking peptides.36 One


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12373



proposed rationale is that in vivo direct loading of short

peptides occurs for all cells expressing MHC molecules,

including T and B cells. Unlike DCs, they are incapable of

generating an effective immune response.37 These early

findings benefitted the design of synthetic long peptide

(SLP) vaccines where both helper T cells (TH) peptides and

toll-like receptor (TLR) ligand or peptides could be hybridized

to enhance the immune response. For example, Jackson et al.

demonstrated one single SLP vaccine that can target TLR 2

can serve as self-adjuvant and contain both MHC I- and II-

specific peptides (T cell epitopes).38 SLPs cannot directly bind

with MHC I and have to be internalized by DCs for further

presentation. Furthermore, in silico computer-based ap-

proaches can be used to benefit the discovery of peptide


vaccines by enabling quicker screening and more accurate

predictions of peptide sequences with high immunogenicity

and binding affinities to MHC I and II molecules.39 These

approaches have been used to accelerate the development

pipeline of vaccines, and particularly for emerging viral diseases

such as COVID-19. These computational approaches can

predict the binding affinity of specific peptides sequences to

either MHC I and II molecules or B cell receptors/antibodies.

In this regard, theoretical B and T cell epitopes specific to

SARS-CoV-2 were synthesized to generate a multiepitope

protein.39 The antigenicity, allergenicity, physiochemical

parameter, and secondary and tertiary structures of these

multiepitope proteins were determined in silico. However, the

processing in the endosomal/lysosomal compartments of the


Figure 3. Action mechanisms of vaccine formulations. (a) Vaccines generate humoral and cellular immunity within lymph nodes (LNs): I.

DCs can process the antigens and present the peptide fragments via both MHC class I and class II molecules. II. B cells can directly

recognize the antigens via BCRs and present the antigenic peptide fragments by MHC class II to helper T cells (CD4+). Stimulated B cells

can subsequently initiate a humoral immune response. III. Cytotoxic T cells (CD8+) can recognize the antigenic peptide fragments presented

by MHC class I through TCRs and trigger the cellular immune response.12,45 (b) Intracellular response of DCs to antigen presentation for

different types of vaccines through PRRs. I. Vaccine formulations can be effectively internalized by cells, followed by II. Endosomal release.

Before III. MHC loading of antigen peptide, peptide vaccine undergoes enzymatic processing, DNA vaccine undergoes transcription and

translation, and mRNA vaccine undergoes translation.24,27,37


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12374



designed multiepitope SARS-CoV-2 antigens remains to be

addressed.

Cell-Based Vaccines. DCs are the most potent antigen-

presenting cells (APCs) at the interface of innate and adaptive

immunity to sense the “danger signals”, process the antigens,

and present the nonself-antigens.40 Unlike other types of

vaccines, DC vaccines are manipulated ex vivo to be antigen-

specific, and thus they are ready to initiate immune responses

after being reinfused in vivo. The DC vaccine is quality

controllable and natively able to traffic efficiently in vivo.

Different methodologies have been developed to produce

modified DC vaccines ex vivo, including loading DCs with

immunogenic peptides or proteins (subunit DCs), viral

transduction to express immunogenic peptides,41 or using

mRNA-based DCs.42 Different strategies may produce DC

vaccines with different potency, for instance, genetically

modified DCs may be more efficacious in immune activation

than subunit DC vaccines.41 In addition, supplementary

immunomodulatory genes could be incorporated to enhance

the potency. However, the process of producing cell-based

vaccines is labor-intensive and expensive.43 Thus, this method

could be the most challenging to be scaled up as a universal

vaccine for a global pandemic.43

VACCINE FORMULATIONS

Action Mechanisms. Vaccines are commonly adminis-

tered intramuscularly, making it less efficient in interacting

with the abundant immune cells present within the skin.

Alternative strategies, including oral or intranasal deliveries, are

also widely studied because of their ease of use.44 The

effectiveness of alternative strategies is based on mucosal

immunity.44 Nonetheless, different strategies rely on the

activation of the adaptive immune system, which involves the

interactions between T cells and APCs (DCs and B cells).12 As

shown in Figure 3a, APCs can internalize exogenous antigens

and process them into antigenic peptide fragments. The


peptide fragments can subsequently be loaded onto the MHC

II molecules and displayed on the surface of the APCs.12,45

CD4+ helper T cells could recognize peptides presented by

MHC class II through T cell receptors (TCRs), which could

induce the production of cytokines, such as IL-2, that are

essential to the normal function of immune cells.12,45 B cells

can directly capture antigens through B-cell receptors (BCRs).

The antigens can be internalized and subsequently processed

into antigenic peptides for presentation by MHC class II to

CD4+ helper T cells. This process is required for the clonal

expansion and differentiation of B cells into IgG antibody-

secreting plasma cells and memory B cells, which is classified as

humoral immunity.12,45 Generally, only endogenous antigens

could be degraded into peptides and loaded onto MHC class I

molecules for presentation to CD8+ T cells.12 Cross-

presentation is the ability of certain APCs to uptake, process,

and present extracellular antigens in an MHC class I-

dependent manner to CD8+ T cells. Cross-presentation can

be achieved by several cell types, including DCs (the primary

cell type), neutrophils, macrophages, and endothelial cells.12

The MHC class I presented peptide fragments can activate

CD8+ cytotoxic T cells through TCR recognition. Antigen-

specific CD8+ cytotoxic T cells can then recognize these

peptide fragments presented by target cells and kill them,

which refers to cellular immunity.12,45 Lastly, DCs are also

capable of detecting “danger signals” via an array of receptors

called pattern recognition receptors (PRRs). PRRs are capable

of recognizing pathogenic molecules as part of the innate

immunity for a response to virus infection by identifying

pathogen-associated molecular patterns (PAMPs) from viruses

or microbes. The understanding of innate immunity could

benefit the development of adjuvants to enhance the immune

response further. Together, the understanding of the action

mechanisms facilitates the design of vaccine formulations to

boost their efficacy.46

Requirement of Formulations. Based on the current

understanding of eliciting optimal immune responses, materi-


Figure 4. Materials-based vaccine formulations for important immunological functions. (a) Vaccine formulations that protect the active

“cargo” and enhance cellular uptake may be based on lipids, polymeric materials, and/or inorganic particles. (b) Vaccine formulations that

allow LN targeting by forming extra small-sized albumin-based vaccine complexes, tuning the net charge to be negative and incorporating

targeting ligands. (c) Vaccine formulations that promote immunostimulatory effects by using microbial wall-derived polysaccharides as

nanocarriers, co-delivering adjuvants with peptide antigens, and exploring innate immunostimulatory lipids.


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12375



als-based nanoengineering approaches could improve the

efficacy of vaccines. In this regard, different methodologies

could be applied based on different types of vaccines, with

either distinct or shared requirements. Virus and viral vector-

based vaccines generally do not require further formulations,

and their efficacy and safety profiles are largely dependent on

the purification and inactivation strategies, the optimization of

culturing of viral particles, and the choice of vectors.47 They

have efficient innate advantages in delivery due to their small

sizes, and they are easily recognized by the immune system.

Although inactivated or weakened virus-based vaccines still

make up most of the effective vaccines today, they are not

effective for all pathogens.47 Similarly, the potency of DC

vaccines is mostly dependent on how DCs are manipulated ex

vivo. Strategies in further potentiating cell-based immuno-

therapies, such as the decoration with nanoparticles (NPs),48

are mainly developed for cancer immunotherapies, but they

may be less suitable for large-scale manufacturing that is

needed for the production of antiviral vaccines.

In contrast, synthetic vaccines (nucleic acid-based and

protein/peptide-based ones) generally require additional

formulation designs for optimal efficacy.49 These synthetic

molecules all have essential secondary or higher-order

structures that could be disrupted in physiological con-

ditions.12 Various endogenous enzymes (such as nuclease or


protease) can digest these synthetic molecules before reaching

immunocompetent cells, such as APCs, or immunocompetent

sites, such as secondary lymphoid organs.12 Furthermore,

physiochemical features of these synthetic molecules may pose

challenges for cells to efficiently uptake and process them. For

instance, DNA and mRNA are too large to directly diffuse

across the cellular membrane and their dense negative charge

repels them from cell membranes, which further prevents

efficient cellular uptake.27 Even after penetrating the

membrane, the vaccine formulations need to reach the

endosomal compartments either to be processed into peptides

(protein/peptide-based vaccines), released for further tran-

scription (DNA-based vaccines), or translation (mRNA-based

vaccines). The resulting peptide fragments can be further

loaded onto MHC molecules and exported to the membrane

surface for antigen presentation to T cells (Figure 3b).50

We refer vaccine formulations to engineering approaches

that could facilitate the in vivo delivery of vaccines and enhance

immune responses.51 We categorize the vaccine formulations

in terms of commonly used materials: lipids, polymeric

materials, and inorganic particles (Figure 4). Commonly

pursued feature goals include protection of cargo, improving

loading capacity, and enhancing cellular uptake (Figure 4a).52

In addition, the ideal microenvironment for initiating and

amplifying immune responses is secondary lymphoid organs


Table 1. List of the Representative Formulations Used for Vaccines

materials formulations cargos (vaccine types) functions refs


lipids


ionizable lipid mRNA (glycoprotein 100 (gp100) and tyrosinase-

related protein 2 (TRP 2)) and adjuvants

lipopolysaccharide (LPS)


stabilization, enhanced cellular uptake 54


covalently cross-linked lipid

bilayers


protein/adjuvants (Monophosphoryl lipid A

(MPLA))


stabilization, sustained delivery, co-delivery 55


net negative charged formulations

by tuning cationic lipids and

mRNA ratio


mRNA (influenza virus hemagglutinin (HA) and

OVA)


LN targeting, stabilization 56


nanodiscs via sHDL peptide neoantigens/adjuvants (CpG); dual

adjuvants (CpG and MPLA)


co-delivery, enhanced uptake 57, 58


lipids with cyclic amino headgroup mRNA (OVA and E7) immunostimulatory function 59


polymers


modified dendrimer mRNA (H1N1 influenza, Ebola viruses, and

Toxoplasma gondii parasite)


enhanced loading, cellular uptake, and

endosomal release


60


stearic acid conjugated low MW

PEI


peptide (TRP 2) and mRNA (HIV-1) lowered charge density, higher

biocompatibility


61, 62


cyclodextrin conjugated PEI mRNA (HIV) lowered charge density, higher

biocompatibility


63


PLGA NPs protein (OVA) controlled antigen release 64

PEI/chitosan/poly lysine coated

PLGA/PLA


protein (HBV surface antigen) enhanced loading and cellular uptake,

intrinsic immunostimulatory function


65−67


PBAE and PLGA plasmid DNA (Luciferase) enhanced loading and lower cytotoxicity 68

ionizable amino polyester mRNA (luciferase) spleen targeting 69

oligopeptide-modified PBAE mRNA (enhanced green fluorescent protein) APC targeting 70

albumin and albumin-binding

lipids


peptide (OVA) /adjuvant (CpG) LN targeting 71


polysaccharide from microbial cell

wall and PEI


mRNA (OVA) immunostimulatory, enhanced loading and

endosomal release


72


inorganic

particles


aluminum hydroxide phosphoserine conjugated protein antigen (HIV

trimer)


immunostimulatory, sustained release 73


chemically functionalized Au NPs

with hydrophobicity


− intrinsic immunostimulatory function 74


aspect ratio optimized and surface

modified Au NRs


plasmid DNA (HIV-1) enhanced cellular uptake, intrinsic

immunostimulatory function


75, 76


sized optimized Au NPs peptide (OVA) and adjuvant (CpG) enhanced DC activation 77

mesoporous Si NPs adjuvant; protein (E2 of diarrhea virus) intrinsic immunostimulatory function 78, 79

antigen density tunable/Size

controllable QDs


peptide (self-antigen from multiple sclerosis) LN homing and higher density of antigens

leads to more effective tolerance, intrinsic

antiviral effects


80, 81


oxidized multiwalled CNTs protein (NY-ESO-1) and adjuvant (CpG) co-delivery, enhanced uptake 82


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12376



where APCs, T cells, and B cells are in close proximity for cell-

to-cell interactions. Therefore, targeting DC or LN can be one

type of approach to pursue (Figure 4b).53 To further boost

immune responses, immunostimulatory components or

adjuvants are usually required, and coformulation approaches

have been widely explored for developing immunostimulatory

vaccine formulations (Figure 4c). Relevant studies have been

summarized in Table 1.

Lipids. Lipid molecules are suitable for vaccine formula-

tions because of their high safety profile and tunable

physiochemical features.55,59 These advantages also make

them the leading platform in developing vaccines quickly.

Currently, there are two mRNA vaccine forerunners in

development for COVID-19 that are based on lipid

formulations (Moderna and BioNTech).10 Lipids are a class

of amphiphiles that contain hydrophilic head groups and

hydrophobic tails. Interactions between those molecules can

drive the formation of spherical vesicles to encapsulate vaccine

formulations either with one (unilamellar) or more (multi-

lamellar) lipids bilayers.55 Lipid molecules have also been

explored to achieve distinct biological functions.59 For

instance, cationic lipids are commonly used for efficient

cellular uptake and increase the loading of negatively charged

mRNA and DNA.56 DOTMA, DOTAP, and DOPE are

common examples.56 However, cationic lipids are commonly

associated with high toxicity and neutralization by anionic

serum proteins, which reduce the delivery efficacy.83 There-

fore, ionizable lipids that can change their charge state at

different pH conditions can be utilized to maintain efficacy, to

reduce toxicity, and to facilitate the endosomal release of

cargos.84 Some common strategies to improve the efficacy of

LNPs include enhancing cellular uptake and promoting

endosomal escape by incorporating cholesterol, polyethylene

glycol (PEG) lipids, or helper lipids DOPE.56 Hydrophobic

cholesterol can fill the gaps between lipid tails to stabilize the

vesicle. PEG lipids can shield the vesicles and protect them

from clearance when systemically delivered.56 Helper lipids

have unsaturated bonds to form an unstable hexagonal lamellar

phase that enhances the endosomal escape.85 In addition, good

biocompatibility of lipids makes it possible for expedited FDA

approvals.86 Therefore, LNPs can be optimized in vaccine

formulations to stabilize and enhance uptake of cargo, to

improve immunostimulatory functions for antigens, and for

targeted delivery. Oberli et al. utilized an ionizable lipid, a

phospholipid, cholesterol, and a PEG anchored lipid as LNP

formulations for mRNA vaccine delivery, which resulted in

stable nanoformulations, efficient antigen presentation, and

elicitation of a strong T cell response.54 However, the in vivo

potency of LNP-based vaccines could be limited by their

suboptimal physiological stability, which can cause the fast

release of antigens before initiating any immune response.

Moon et al. developed a hyperstabilized lipid-based vaccine

formulation by covalently cross-linking two lipid molecule

layers between their head groups using a mild chemical

condition that is compatible with loaded antigens.55 The

resulting vesicles are stabilized multilamellar structures that

could trap high levels of protein antigens. Combined with

PEGylation as a well-established stabilization strategy, the

hyperstabilized NPs can sustain the release of antigens up to 30

days in the presence of serum, which significantly enhanced the

humoral immune response.87 By adding the lipid-like mono-

phosphoryl lipid A (MPLA) as adjuvants, the vaccine


formulation can further enhance T cell-mediated humoral

responses.

In addition, LNPs formulations can be optimized for LN

targeting, thereby enhancing the immune response in situ. One

notable study by Kranz et al. demonstrated that a designed

mRNA-lipoplexes (RNA-LPX) could target DCs in vivo simply

by using lipid molecules.56 They established a library of RNA

lipoplexes by tuning the lipid to RNA ratios. Cationic lipids

DOTMA or DOTAP and the helper lipid DOPE or cholesterol

were used. The charge ratio, size, ζ potential, colloidal

properties, and RNA stability also varied with the shifting of

lipid to RNA ratios. Interestingly and unexpectedly, all

negatively charged particles displayed selective targeting to

the spleen. The negatively charged formulation (1.3:2

lipid:RNA) displayed effective targeting of DCs and formed

monodisperse NPs (∼300 nm) that protected encapsulated

mRNA. The authors demonstrated the in vivo efficacy using

RNA-LPX encoding for the influenza virus hemagglutinin

(HA), where DCs, NK, B, and T cells were activated by a

single intravenous injection of HA-LPX. They also observed an

enhanced IFN-α production by vaccination, which is a typical

hallmark of APCs sensing for antiviral environment develop-

ment induced by RNA virus infections through TLR3 and

TLR7.88

To enhance immunostimulatory effects, a common strategy

is to co-deliver adjuvants to stimulate immune responses. Kuai

et al. designed an antigen and adjuvant co-delivery nanodisc

platform based on synthetic high-density lipoprotein

(sHDL).89 One of the striking features of the nanodisc is its

extra-small size: ∼10 nm in diameter, which contributed to the

enhanced trafficking to LNs.57 Exploiting the endogenous role

of HDL as a cholesterol carrier, adjuvant CpG was modified

with cholesterol to efficiently insert into preformed nanodiscs.

This strategy also allowed the co-loading of more than one

adjuvant, which promoted a synergistic immune activation.58

Together, these features contributed to the sustained antigen

presentation by DCs. Recently, there is a growing interest in

exploring innate immunostimulatory functions of lipid

molecules. To expand the diversity of lipid molecules, Miao

et al. developed a library of lipids through a one-step three-

component reaction that included amines, isocyanides, and

alkyl ketones.59 This strategy significantly increased the

diversity of synthesized lipids compared to traditional two-

component reactions.59 They showed that the immunostimu-

latory effect of lipids was dependent on the head groups. The

identified lipids with cyclic amino head groups displayed

immune cell activation efficacy through the intracellular

stimulator of interferon genes (STING) pathway. The lipids

could be further formulated with mRNA to form LNPs as

mRNA vaccine formulations to efficiently deliver the

oligonucleotide products. Another advantage is that the

formulation could activate the mRNA-independent intra-

cellular STING pathway rather than TLRs, which could

reduce systemic toxicity. In addition, inhibition of the antigen

expression related to the TLRs binding by exogenous mRNA

was prevented to favor an optimal immune response.27

Polymers. Aside from the lipids system, polymers have

been widely explored for the delivery of different vaccine

candidates. Specifically, cationic and ionizable polymers are

promising in nucleic acid-based vaccine formulations.

Positively charged polymers can condense nucleic acids as

nanocomplexes for efficient cellular uptake and endosomal

release. In addition, early studies showed that cationic


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12377



materials tend to induce an inflammatory response.90 One

commonly applied polymer category is polyamines, such as

polyethylene imine (PEI) and its derivatives.91 Chahal et al.

utilized modified dendrimer for antigen-encoding mRNA

replicon complexation to form monodisperse NPs.60 This

delivery system, with self-amplifying mRNA, enabled immuni-

zation in mice, which protected them from various virus

challenges, including H1N1 influenza, Ebola virus, and the

Toxoplasma gondii parasite.60 However, their high molecular

weight (MW) and branched polymer structure (needed to

allow a massive loading of vaccine therapeutics) also lead to

severe cytotoxicity, which limits wider applications.92 Zhao et

al. developed a low MW PEI with modifications (stearic acid

and 2 kDa PEI conjugates) to form self-assembled micelles for

delivering mRNA encoding HIV-1 group-specific antigen

(gag).62 Li et al. used 2 kDa PEI conjugated with cyclodextrin

for intranasal delivery of HIV antigen-encoding mRNA that

was able to stimulate HIV-specific immune response.63

Cyclodextrin modification provided mucosal permeation and

mitigated cytotoxicity by lowering the charge density. Other

commonly used polymers include polyester-based, such as

poly(lactic-co-glycolic acid) (PLGA) and poly(lactic acid)

(PLA), because of their biocompatibility and biodegradabil-

ity.93 Several studies have demonstrated the delivery of protein

antigens in PLGA NPs with enhanced immune responses as

compared to soluble formulations.94,95 In addition, such solid

polymer particles can control the kinetics of antigens released

from the polymer core based on different biodegradation

profiles.64,96 However, nonmodified PLGA is generally

negatively charged, and as a result, it has a limited cellular

uptake efficiency and loading of anionic antigens.66 Therefore,

various coating or modification strategies (such as PEI,65,97

chitosan,66 and poly-L-lysine (PLL) coatings) have been

demonstrated to enable the delivery of vaccines and promote

interactions with negatively charged cellular membranes.67

Notably, the cationic polymeric delivery carrier itself could

induce T helper cell responses through TLR-4 dependent IL-

12 secretion.67,98 Poly(β-amino esters) (PBAE) is also

extensively used in delivering plasmids DNA or mRNA

vaccines because of its nontoxic and biodegradable nature.99

Fields et al. developed NPs formulations for plasmid DNA

delivery by combining PLGA and PBAE to reconcile the

drawbacks from each component: the low DNA loading

efficiency of PLGA and the cytotoxicity of PBAE.68 Other

polymeric materials have also been explored, such as

polysaccharides100 and block copolymers.101

In terms of targeted delivery to immune cells, versatility in

polymer chemistry shows great potential in developing tissue-

selective formulations. For instance, a large library of PBAE-

based polymers could be easily synthesized and screened for

desirable target polymers.69 Kowalski et al. designed an

ionizable amino polyester (APE) for mRNA delivery through

the ring opening of lactones with amino alcohols.69 Tissue-

selective delivery was achieved by different APE candidates,

including APCs within the spleen.69 Oligopeptides have been

used to modify PBAE to allow the tuning of the charge status

of NPs.102 Initially, this was done to optimize cargo loading

and mitigate cytotoxicity. Later, Fornaguera et al. demon-

strated that specific oligopeptide modifications lead to APCs

targeting within the spleen and exhibited an efficient

transfection of mRNA.70 Inspirations also come from using

native LN trafficking molecules. Clinically, visualization of

sentinel LNs is achieved by injecting dyes that bind to albumin


because of the LN trafficking capability of albumin particles.103

Inspired by the endogenous role of albumin, Liu et al. designed

a vaccine by conjugating the adjuvants or antigens to the

albumin-binding lipid tail.71 The vaccines displayed a notable

increased accumulation in the LNs, which decreased the

systemic distribution of the vaccines and prolonged the vaccine

bioavailability. These features limited the systemic toxicity

while increasing the potency of vaccination for OVA antigen,

HIV antigen, and human papillomavirus (HPV) antigen.71

Some polymeric delivery carriers have been demonstrated to

have immunostimulatory effects, such as polymers with

cationic charge and hydrophobic domains.104,105 In addition,

inspired by microbes’ immunostimulatory effect, Son et al.

designed sugar capsules that were derived from microbial cell

walls for vaccine NPs coating.72 The coating process was

achieved by using silicate NPs (Si NPs) as a template, followed

by its removal after the mRNA was loaded and the sugar

capsules were entirely coated. Notably, the removal of Si NPs

renders relatively flexible nanostructures that are beneficial for

lymphatic drainage and accumulation.72 The mRNA cargos

can be efficiently loaded by PEI coating due to electrostatic

interactions. The PEI coating was also expected to facilitate the

endosomal release for mRNA translation and antigen

presentation. The sugar capsule itself functioned as a strong

DC activator by eliciting the production of a variety of pro-

inflammatory factors, including IL-6, TNF-α, and IL-12 p40,

without the use of exogenous adjuvants.72

Inorganic Particles. Inorganic materials have been used in

vaccine formulations for a long time. Back in 1926, an

aluminum compound (alum) was discovered to be able to

precipitate the toxoid to enhance the immunogenicity.106 Alum

is still the most widely used adjuvant today. In spite of over 80

years of development, the underlying mechanisms for alum as

an adjuvant are still controversial.107 Popular explanations

include the “depot theory” for extended antigen release106 and

enhanced uptake by APCs through antigen absorption on

alum.108 Various immunity signaling and activation pathways

have also been identified to contribute to the function of the

adjuvants.109 Recently, Moyer et al. utilized the interaction

between alum and phosphoserine (pSer) to engineer a short

peptide-immunogen conjugate to significantly prolong the

immunogen release from pSer-immunogen:alum complexes.73

The complexes could form NPs for efficient trafficking to LNs

and initiated antigen processing and presentation by APCs.73

Site-specific pSer modification on the base of the HIV trimer

could enable the immunogen (HIV trimer) to be displayed

with a specific orientation, which could prevent the induction

of undesired NAbs.73 Apart from alum, other inorganic

materials have been proposed that may stimulate vaccine-

specific immune responses, such as silica crystal110 and calcium

phosphate.109

In addition to the traditional use as immunostimulatory

adjuvants, a wide range of biocompatible inorganic particles

have also been developed as vaccine formulations,111 including

quantum dots (QDs),80 gold nanoparticles (Au NPs),77

carbon nanotubes (CNTs),82 and Si NPs.79 The physical

properties of formulations, such as surface charge, hydro-

phobicity, and aspect ratios, are closely associated with the

efficacy of vaccines.112 Inorganic materials have a high degree

of tunability to allow them to fulfill the requirements of vaccine

formulations. Easy preparation and good storage stability are

additional advantages of inorganic particles.113 For example,

Au NPs have been chemically functionalized to investigate the


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12378



relationship between surface hydrophobicity and the associated

immune activity.74 It is also expected that other functional

groups may bring a pathway-dependent immune activation

that is specifically desirable for antiviral defense. Xu et al.

described surface-engineered Au nanorods (NRs) for plasmid

DNA vaccine delivery.75 With certain surface modifications

and aspect ratios, cellular toxicity of Au NRs can be lowered,

internalization can be enhanced, and intrinsic adjuvant features

can be utilized.76 In vivo vaccination with HIV-1 envelope

glycoprotein (Env) plasmids demonstrated enhanced cellular

and humoral immunity.75 Zhou et al. synthesized a series of Au

NPs (15 to 80 nm) for co-delivery of antigens (OVA peptides)

and adjuvants (CpG) by surface conjugation.77 Particles of 60

and 80 nm outperformed other sizes (15, 30, and 40 nm) in

activating DCs; this was most likely due to a higher payload of

antigens. Notably, they promoted DC homing to the LNs and

induced potent cellular immune response.77 High tunability is

also achievable in other organic particles, such as adjuvant-

based mesoporous Si NPs,78 mesoporous Si NP-based


antigen/adjuvant co-delivery systems,79 antigen density tuna-

ble QDs,80 size-controlled antiviral QDs,81 and surface

oxidized antigen/adjuvant-based multiwall CNTs.82

VACCINE DELIVERY DEVICES

If the vaccine formulations are considered as utilizing

nanoscale engineering, then vaccine delivery devices utilize

micro/macroscale engineering to improve vaccine efficiency

and efficacy. Beyond the optimization of the vaccine itself and

designing appropriate delivery carriers, the efficacy of a vaccine

also depends on pharmacokinetics (PK) of the antigens after

administration.114 When the body is exposed to an exogenous

antigen, both humoral and cellular immune responses are

initiated to neutralize it, which also leads to immunological

memory.12 During a virus infection, the replication process

typically takes one to several weeks to sustainably stimulate the

immune system with a continuous supplement of antigens.115

In contrast, typical vaccines show a rapid clearance and can

only be detected in LNs within days.114 Although still poorly


Figure 5. Summary and characteristics of various vaccine delivery devices. (a) Drug delivery devices developed for vaccine administration: I.

Traditional vaccine-containing solutions for bolus injection. II. Minimally invasive microneedle devices (including fast-dissolving, sustained-

released, and liquid-eluting hollow microneedles). III. Injectable drug delivery devices (such as microparticles, solid scaffolds, and hydrogel-

based). IV. Other devices for enhancing vaccine efficacy (such as electroporation and iontophoresis). (b) Schematic for representative PKs

of vaccine and administration features for different vaccine delivery devices: I. Traditional bolus injection is invasive and may require

multiple doses. II. Fast dissolving or liquid eluting MNs can be self-administrable and minimally invasive but still require multiple doses.125

III. Sustained released MNs can be both minimally invasive, with only one dose necessary,122 while scaffold and gel vaccines could be

injectable to avoid implantation and also only require one dose. IV. Pulse-released microparticles can mimic a multiple-dose regimen with

only one injection.127,128 (c) Schematic of representative mechanisms for vaccine delivery devices. I. Scaffold vaccines: The scaffold could

sustainably release recruiting factors to home DCs, and encapsulated antigens could further be processed by DCs in situ.129 II. Gel vaccines:

Hydrogel microenvironment can be engineered to support encapsulated cell vaccines (such as DC vaccines) by incorporating cell adhesive

peptides and supplementary factors.136


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12379



understood, the PK of a vaccine, which refers to the dynamics

of the antigens presence during immunization, has a profound

impact on immune response.87 One study indicated that

increasing the dosage of vaccines over 2 weeks at both priming

and boost stages induced a durable immune response and

increased antibody production compared to traditional bolus

immunization.87 Also, computational models suggested that

antigens can be better captured in germinal centers (GCs) if

antigen availability is extended.116 In addition to the

advantages of extended antigen availability, traditional

vaccination strategies often fail to yield the desired optimized

immune responses in a single dose. Furthermore, vaccine

regimes are often designed to be multiple shots, thus achieving

an even longer immune protection by rechallenging the

immune system with the same antigens in secondary

“boosting” shots administered after the primary shots.117

Therefore, vaccine delivery devices have been widely

investigated, to develop single-dose vaccines, minimal

invasively administered vaccines, and self-administrable

vaccines (Figure 5a,b). However, its potential can only be

realized if various design principles are met (Table 2).

Microneedles Vaccines. Transdermal microneedles

(MNs) patches have been proposed as a promising drug

delivery device for immunotherapies144−147 based on the

minimally invasive nature of MNs and the highly active

immune environment of the skin. Additionally, since MNs

patch-based vaccines can be easily administered without the

need for significant medical training, these vaccines can be

easily deployed both in areas with limited medical resources,

such as developing countries and during times of significant

medical crisis, as seen during an outbreak or pandemic. Early

studies include the use of vaccine-coated metal MNs118 and

fast-dissolving polymer-based MNs.119 Recently, the dissolving

MNs (carboxymethyl cellulose-based) were used for adjuvants

and Ad5-based vaccine co-delivery in an attempt to address

COVID-19.147 A 1 month storage of MNs vaccine at 4 °C

maintained the immunogenicity of Ad, where no significant

loss in antibody responses was shown in mice studies.147

Similar dissolving MNs are also being investigated to deliver

the SARS-CoV-2 subunit spike protein (S1).120 Potent virus-

specific antibody responses were elicited in mice as early as 2

weeks after the immunization with MNs. The dissolvable MNs

for influenza vaccine have been previously evaluated in a phase

I clinical trial compared to the traditional intramuscular

hypodermic injection of a single-dose vaccine.121 The MN

patch was well-tolerated and immunogenic after a single-dose

vaccination. In addition, self-administration exhibited efficacies

comparable to the administration by medical workers. The

dissolving feature leaves no sharp wastes, and self-admin-

istration could significantly relieve the pressure on medical

resources, which could be especially critical in coping with a

pandemic. Beyond these milestones, advances in micro-

fabrication, materials engineering, and vaccine formulation

optimization have provided additional opportunities to

enhance vaccine efficacies.148

Recent advances in microfabrication, such as 3D printing

and stereolithography, can allow for more sophisticated MNs

designs.125 Liquid-based vaccine formulations are still the most

widely used.125 However, liquid-based vaccine formulations are

incompatible with most solid-based MNs forms. Inspired by

the capability of snake-fangs to efficiently inject venom, Bae et

al. used an exposure lithography-based strategy to produce

MNs patches with snake-fang architectures for efficient liquid


Table 2. Summary of Typical Designs, Advantages, and Limitations of Different Vaccine Delivery Devices

delivery devices designs advantages limitations refs


MNs vaccine (minimally invasive; minimal require-

ment on medical professionals and medical

equipment)


fast-dissolving

MNs


fast-dissolving polymers as MNs matrix fast and enhanced transdermal delivery; could be cold-chain free multiple administrations; vaccine for-

mulations may get denatured from

MNs fabrication


118−121


sustained-re-

lease MNs


controlled biodegradable materials as

MNs matrix


potential to be single shot; protect the vaccine formulations in the

physiological environment; could be cold-chain free


vaccine formulations may get dena-

tured from MNs fabrication


122−124


liquid-eluting

MNs


hollow MNs with different architectures compatible with liquid vaccine formulation; fast and enhanced

transdermal delivery


multiple administrations; requirement

of cold-chain transport


125


injectable materials assisted vaccine (potential for

single-shot vaccines to reduce the requirement of

medical resources)


pulse-released

microparticle


materials with different biodegradation

profiles


optimal release for a potent immune responses; release kinetic can

be optimized by materials engineering


vaccine formulations may get dena-

tured during the fabrication


126−128


scaffold vac-

cine


solid scaffolds that could tune the release

of different immunomodulatory factors


continuously modulate the immune response in situ injectability need to be achieved to

avoid surgical implantation


129−134


hydrogel vac-

cine


biomimicking gellable materials with

tunable biophysical properties and high

cytocompatibility


compatible with various types of vaccines, including cell-based

vaccines; simple fabrication process and the water environment

help protect the cargos


in situ gelation may be weak to achieve

sustained release


135−143


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12380



formulations delivery.125 The design was compatible with most

existing vaccine formulations, and the patch can be applied

with the pressure of the thumb and achieve an efficient

transdermal delivery.125 Nevertheless, solid-based vaccines,

including MNs-based ones, also attracted significant attention

because of the decreased cost of vaccine manufacturing,

transport, and storage when compared to the liquid-based

formulations.149 One of the key challenges faced by these MNs

patches is the loss of vaccine activity during the drying or

heating process of MNs patch fabrication,150 where molecular

structures of vaccines may be disrupted.151 Different strategies

can be adopted, such as the use of lyoprotective polymers as

MNs matrices, which include silk fibroin,122 gelatin, dextran,

and PVP,152 or adding sugar molecules as stabilizing excipients,

including trehalose and sucrose.147,152 In the vaccine discovery

phase, the introduction of certain mutations can improve the

thermal stability of vaccines as well,148 which can be further

applied to specific MNs formulations. Additionally, a hyper-

stabilized vaccine formulation, such as the interbilayer-cross-

linked multilamellar vesicles, helped protect vaccines during

the drying process of MNs fabrication and have outperformed

other traditional delivery strategies in inducing a potent

antigen-specific immune response.123

MNs also provide advantages in tuning optimal release

profiles of vaccines. As mentioned earlier, the sustained

availability of vaccines is a desirable feature for immunization.

Vaccines targeting different infectious diseases may vary in

terms of optimal release profiles based on the unique features

of each virus. Initially, MNs for vaccination were mainly an

upgrade to traditional bolus delivery by using a fast-dissolving

MNs matrix,119 which requires multiple doses for sufficient

immunity. With the exploration of the desired PK of vaccines,

MNs can play a much more important role. For example, a silk

matrix with crystallized structures could be used as an MNs

matrix to sustainably release loaded vaccines over 2 weeks and

to elicit potent immune responses with one single admin-

istration.122 In addition, biodegradable polymers with known

degradation kinetics can be used to encapsulate the vaccines

and tune antigen release profiles. These polymers include

biodegradable cationic PBAE123 and PLGA.124

Injectable Material-Assisted Vaccines. Macroscale

biomaterials strategies have recently boosted the field of

immunotherapy.153 There are many advantages in controlled

release and encapsulation of various cargos, ranging from small

molecules to proteins and cells.153 These varied cargo elements

can help to meet the need for diverse types of vaccine

candidates. In terms of controlling PKs of vaccines, both

sustained and pulsatile releases could be possibly achieved.153

In addition, materials engineering can enable devices that do

not require surgically invasive administrations.153 Together,

they could minimize the requirement of medical resources

without mitigating the vaccine efficacy.

PLGA has also been developed for a single-shot vaccine

based on the tunability of its biodegradation profile.126 In the

case of polio vaccines, two to three administrations over several

months are required. Tzeng et al. demonstrated that IPV could

be encapsulated within PLGA microspheres stably and released

through two bursts that were one-month apart.126 They

examined the effect of adding cationic excipients (Eudragit E,

PLL, and branched PEI) to stabilize IPV antigens, protecting

them from denaturation by acidic byproducts, and modulate

degradation profiles of PLGA microspheres.126 Notably, a

single shot of PLGA microsphere-based vaccines could elicit


the production of NAb as potent as two bolus injections.126

Combined with the advances in microfabrication and additive

manufacturing, McHugh et al. fabricated microscale (∼400

μm) hollow PLGA-based microparticles that were inject-

able.127 The microparticles can be filled with vaccine

formulations, and the degradation rate of the microparticle

shell (a lactic/glycolic copolymer) can be tuned to achieve

controlled release spanning from a few days to a few

months.127 Such a design can accomplish the goals of general

transdermal vaccine injection and achieve temporally con-

trolled pulsed delivery of antigens in a single injection.127

Recently, a similar PLGA-based microparticles platform was

applied to release a STING agonist (cyclic guanosine

monophosphate-adenosine monophosphate (cGAMP)) in a

pulsatile manner over a long period (∼16 days).128 One single

injection of the pulsatile drug-releasing microparticles induced

potent immunity as effective as three separate injections. These

engineered platforms are versatile in terms of encapsulating

different cargos required by vaccine formulations and tuning

release kinetics correspondingly. However, even though the

use of FDA-approved PLGA could facilitate the clinical

translation to a certain degree, several design limitations

need to be overcome for translating PLGA-based drug delivery

devices into clinical applications. For instance, the pulsatile

release was achieved with hollow microparticles that have thick

PLGA shells to encapsulate the drugs in the hollow core for

controlled release. This design significantly sacrificed their

loading efficiency due to the limited volume of the core.

Clinically, patients need a high drug dosage, and this

formulation needs further optimization to meet the dosage

requirement. In addition, PLGA can create an acidic

microenvironment during biodegradation, damaging the

encapsulated drugs.126 Furthermore, the encapsulated vaccines

need to survive the fabrication process; brief heating requires

thermally stable vaccines.127,128 All these aspects could

compromise their applications for a broader collection of

drugs or vaccines.

Traditionally, ex vivo modification of antigen-specific DC

vaccines is a multistep process associated with high costs and

high levels of complexity.43 It was proposed that in situ DC

recruitment, activation, and further trafficking to LNs would be

more desirable compared to the ex vivo modification strategy

(Figure 5c).129 The early study by Ali et al. demonstrated the

use of macroporous scaffold vaccines based on PLGA, which

presented cytokines (GM-CSF), danger signals (CpG), and

antigens in a spatiotemporally defined manner. GM-CSF can

be sustainably released from the scaffolds over 20 days for

effective DC recruitment.129 By co-loading the CpG through

PEI-mediated electrostatic interactions, CpG was immobilized

on the scaffolds for local internalization by the recruited DCs.

This process mimicked an infection and was able to

continuously recruit and activate DCs.129 This specific scaffold

vaccine was surgically implanted; nevertheless, it encouraged

further development of injectable vaccines that work similarly.

Kim et al. demonstrated the use of mesoporous silica rods

(MSRs) with a high-aspect-ratio that could be injected with a

syringe needle.130 They could self-assemble in vivo as

macroporous scaffolds to initiate potent immune responses.130

The DCs could be recruited into the scaffolds where the

inflammatory signals (GM-CSF), adjuvants (CpG), and

antigens (OVA) were sustainably released from the scaffolds

for in situ DC activation and maturation. It was also

demonstrated that locally injected scaffold vaccines based on


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12381



MSRs could elicit strong systemic immune responses with the

increase in levels of antigen-specific B cells, helper T cells, and

cytotoxic T cells.130 A PEI coating was later added to the

design to enable antigen loading through adsorption for

enhanced immunogenicity.131 This specific system was

recently utilized for synthetic peptide vaccines and small

antigens.132 This strategy was able to bypass multiple

immunization requirements but sustained a more robust

antibody response, compared to traditional bolus or alum

vaccine formulations.132

Scaffold-based vaccines have displayed efficacy in inducing

systemic immune response through in situ immune modu-

lations. In vivo self-assembled scaffolds provide a good

alternative to bypass surgically implantation; however, the

self-assembling process is commonly undefined in terms of

how long it could take for self-assembling.133 Materials

engineering facilitates the fabrication of scaffolds maintaining

injectability. Bencherif et al. described a strategy for fabricating

injectable scaffolds with shape memory based on cryotropic

gelation.134 The cryogels were based on alginate methacrylate,

and they were capable of being reversibly deformed, over 90%

strain, after being chemically cross-linked in a frozen condition

(−20 °C).134 The ice crystals hindered the cross-linking in

frozen spots and therefore formed the macroporous pores that

not only enabled the shape memory for the gel but also

allowed the DCs infiltration.134

Hydrogels have been extensively explored in the field of

immunotherapy.153 They have advantages in terms of control-

lable delivery of multiple cargos, in situ immunomodulation,

and mild fabrication process for various vaccine formulations

(Figure 5c). Early pioneering work includes the loading of ex

vivo antigen-primed DC vaccines within Ca2+ cross-linked

alginate gel by Hori et al.135 Activated DCs have a short

lifespan, and continuous immune responses may be suppressed

if no sufficient host DCs and T cells could be recruited and

primed in vivo.135 Therefore, cell-based vaccines delivered by

gels are expected to improve cell survival and prolong the

presence of DCs.136 Both host DCs and cytotoxic T cells can

be recruited to the injection site by cytokines and chemokines

secreted from the injected DCs. In addition, the enhanced

efficacy of transplanted cells can be enhanced by supplement-

ing the gels with other immunomodulatory molecules, such as

cytokines (IL-2) or adjuvants (CpG).137 The increased

persistence of transplanted cells could be achieved by

designing gels mimicking ECM by the incorporation of cell-

adhesive peptides, such as RGD.154

Apart from cell-based vaccines, other types of vaccines could

also be incorporated into gel systems, where the release profiles

can be tuned. Roy et al. utilized the in situ cross-linking

mechanism of tetrafunctional poly(ethylene oxide) amine and

poly(ethylene oxide) succinimidyl glutarate to encapsulate

plasmid DNA.138 The gelation did not interfere with the

supercoiled structure of plasmid DNA, and the gelling ability

can be tuned by adjusting the concentrations of these two

components. Notably, gene expression levels were comparable

to bare DNA, but the duration of expression was significantly

longer. Other in situ gel systems explored for vaccination

include temperature-responsive gels139 and Michael-addition-

type hydrogels.140 Singh et al. engineered dextran and PEG-

based hydrogels that could be cross-linked in situ for antigen/

adjuvant delivery.141,142 Umeki et al. designed an immunosti-

mulatory CpG DNA-based hydrogel for antigen delivery.143

To achieve a sustained release of antigens, they proposed to


cationize the antigens by ethylenediamine conjugation where

antigens could form complexes with a negatively charged DNA

hydrogel matrix. The innate immunostimulatory function of

gel and prolonged antigen presence leads to an enhanced

antigen-specific immune response with less toxicity.143 In situ

gelling strategies or physically cross-linked gels from these

studies generally take minutes to hours, which renders limited

control on cargos release and material properties during the

gelling period.143 Therefore, the development of gel systems

that are both injectable and display defined cargo release/

biodegradation profiles could be desirable for vaccine

administration.

Other Delivery Devices. Many more investigations are

currently ongoing to engineer alternative drug delivery devices.

For instance, to promote the uptake of DNA vaccine by cells,

electroporation is one of the traditional and effective strategies

to enhance the uptake of plasmids.155 Different parameters,

such as electrical features (voltage, current, and frequency) and

length of the electrode (intradermal or intramuscular targets),

need to be optimized to ensure efficacy and safety and to

minimize pain. This process has recently demonstrated

promising clinical phase I results for MERS.25 The same

electroporation device (CELLECTRA) is also being inves-

tigated for a DNA vaccine of SARS-CoV-2 from Inovio.156 The

S protein of SARS-CoV-2 encoding DNA vaccine demon-

strated potent antigen-specific T cell responses and the

production of neutralization antibodies in preclinical mice

and guinea pigs studies.156 In order to incorporate the

advantages of MNs, electroactive MNs devices have also

been developed for the electroporation of DNA vaccine in a

minimally invasive manner.157 Similar devices using iontopho-

resis can also enhance the transdermal delivery of vaccines.158

Recently, Tadros et al. designed millimeter-scale star-shaped

particles, termed STAR particles, based on aluminum oxide for

topical vaccine delivery,159 and the minimally invasive STAR

particles elicited an immune response comparable to intra-

muscular injection.159 Even though STAR particles are simple

to manufacture and well-tolerated by the immune system, a

much higher dosage was administered compared to intra-

muscular injection. In addition, potential human-to-human

variability in administration and drug leftovers on the skin may

result in differential efficacy and undesirable toxicity.

In terms of reducing the demands for medical resources,

vaccines that could be intranasally or orally administered are

desired if their efficacy is comparable to transdermal delivery.44

The main strategy for intranasal or oral vaccine development is

currently limited to weakened viral vaccines mimicking natural

infections. Nonliving vaccines are safer but less immunogenic

and thus may benefit greatly from drug delivery formulations

and devices.160 A variety of drug delivery devices have been

proposed for oral vaccination delivery. For instance, the

micromotor-driven microparticles for oral delivery of anti-

gens.161,162 The micromotor vaccine can efficiently enter the

intestine to enhance the antigen uptake. Aran et al. designed a

needle-free oral microjet where the high-pressure liquid jet of

the vaccine was produced by a self-contained gas-producing

reaction.163 It was shown that a buccal immunity response was

enhanced, and this method was able to elicit a potent humoral

immune response both locally and systemically.163 Abramson

et al. engineered a self-orienting system for oral delivery that

could attach to the gastric wall by autonomously positioning to

maximize the delivery efficiency.164 Innovations in drug


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12382



delivery devices will significantly transform traditional

vaccination regimes.

SUMMARY AND FUTURE DIRECTIONS

The devastating disruptions caused by the sudden outbreak of

COVID-19 has urged vaccine development at an unprece-

dented speed.4 Within 70 days of the first release of the genetic

information on SARS-CoV-2, the first dose of a vaccine was

administered to human participants.9 As of August 25, 2020,

31 vaccine candidates are in various stages of clinical trials, and

around 142 candidates are in preclinical trials.10 Apart from the

high diversity of the vaccine candidates, many candidates are

based on different technological platforms, such as nano-

technology delivered mRNA vaccine, electroporation device-

enabled DNA vaccine, and MN-delivered vaccine. The

COVID-19 pandemic has accelerated the deployment of

different vaccine formulations and delivery devices. It is

reasonable to expect that more efforts will be dedicated to

further optimize the vaccine designs to help fight future

outbreaks.

Here, we reviewed recent research efforts in the field of

vaccine discovery, vaccine formulation development, and

vaccine delivery device designs. Because of the deployment

of various engineering technologies, it is expected that future

vaccines will not only need to demonstrate safety and efficacy

but also to be relatively convenient and simple to use for better

coping with population-level vaccinations or sudden outbreaks

worldwide. Nanoengineered formulations could enhance the

immune response by delivering the required components

together and maximizing their uptake by immunocompetent

cells. To expedite the future preparedness of vaccines for the

occurrence of a pandemic, synthetic vaccines have demon-

strated great potential for rapid clinical applications because

they are molecularly defined and independent of the time-

consuming culture of viruses, which can accelerate the initial

vaccine development. However, synthetic vaccines are also the

ones that need to be formulated the most, which constitutes

the major part of their time costs. To accelerate the

development of synthetic vaccines, the nanoengineered

formulations need to be simple and robust to manufacture in

terms of chemistry or the selection of raw materials.165,166 In

terms of vaccine delivery devices, self-administrable vaccines,

single-shot vaccines, and solid MN-based vaccines could

further enhance the availability of vaccines to a broader

population. For instance, self-administrable vaccines can be

mailed to patients at home without the need of medical

experts.165 Single-shot vaccines can be much more convenient

compared to the traditional immunization regimens that

require multiple shots. Furthermore, solid MN-based vaccines

that are cold chain free could simplify the vaccine trans-

portation and distribution process. All of these aspects could

be enabled by engineering approaches to accelerate the

development of vaccines at a pandemic speed.165,166 Although

combining the advantages of vaccine formulations and vaccine

delivery devices is promising, micro/macro-engineering still

needs to develop solutions that do not require “harsh”

conditions to be compatible with nanoengineered formula-

tions.167 Expectedly, platform technologies that can be quickly

adapted to address emerging viral diseases could speed up the

pace of vaccine development.

However, several challenges remain before these goals can

be fully achieved. First, a differently identified virus is unknown

in terms of its interactions with our immune system and which


type of immune responses are most desirable to trigger

prolonged immunity. An improved understanding of our

immune system and its interactions with viruses are the basis

for designing an effective vaccine, which also guides the steps

for designing corresponding vaccine administration regimens.

Second, innovative formulations or delivery devices for

vaccines are mostly in the preclinical stage, which have not

been approved by the FDA yet. For example, MN-based

vaccines showed encouraging results from a phase I trial of

dissolvable MN-based influenza vaccine and indicated a

promise of using it as an alternative strategy for vaccination.121

However, patient compliance with this approach and its

efficacy needs to be tested in a larger population, considering

only 100 human subjects are currently involved. Also, the

efficacy of vaccines by self-administration should be further

validated in comparison to those performed by medical

experts. In addition, the proposed formulations and delivery

devices should be easy to manufacture with minimal batch-to-

batch variability to simplify the quality control process. The

raw materials used could prioritize the FDA-approved ones to

simplify the regulatory process, and the required chemistry is

preferred to be simple and robust. Even though advanced

manufacturing technologies excel in reducing batch-to-batch

variability, a high cost may be associated. The improved

efficacy of a vaccine from these formulations and delivery

devices should not be complicated by the fabrication process

so that it can facilitate the clinical translation and regulatory

approvals.165

Additional challenges for vaccine development also exist

during a pandemic. Multiple vaccine candidates flush into

different stages of clinical trials that could burden the

regulatory process by the FDA for selecting promising

candidates and deciding final ones.4 Furthermore, human

trials are starting even before the most suitable animal models

for safety and efficacy are defined.168 Based on previous

successful preclinical and early stage clinical data for a similar

platform, animal studies are even skipped directly for human

administrations.10 Other engineering approaches such as

organs-on-a-chip169−171 or screening using organoids172 may

serve as a promising tool in future clinical trials for efficacy and

toxicity determination. Ethically controversial human challenge

trials by deliberately infecting humans with a virus have been

proposed and pursued to accelerate the process of vaccine

development.173 Therefore, potentially we should have vaccine

clinical trials specifically designed for different scenarios by

incorporating engineering advances and regulatory guidelines

readily outlined for unusual trials. In conclusion, human health

and societal stability are constantly challenged by sudden virus

outbreaks, and the rapid development of effective vaccines is

believed to be an ultimate solution. We expect to observe faster

and more efficient vaccine development paradigms in the

future by combining the advances in different engineering

fields.


AUTHOR INFORMATION

Corresponding Authors

Ali Khademhosseini −Department of Bioengineering, Center for

Minimally Invasive Therapeutics, Department of Chemical and

Biomolecular Engineering, California NanoSystems Institute,

Jonsson Comprehensive Cancer Center, and Department of

Radiological Sciences, University of California, Los Angeles, Los

Angeles, California 90095, United States; Terasaki Institute for


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12383



Biomedical Innovation, Los Angeles, California 90064, United

States; Email: khademh@terasaki.org

Wujin Sun −Department of Bioengineering, Center for

Minimally Invasive Therapeutics, and California NanoSystems

Institute, University of California, Los Angeles, Los Angeles,

California 90095, United States; Terasaki Institute for

Biomedical Innovation, Los Angeles, California 90064, United

States; orcid.org/0000-0002-3167-111X; Email: wsun@

terasaki.org

Authors

Xingwu Zhou −Department of Bioengineering and Center for

Minimally Invasive Therapeutics, University of California, Los

Angeles, Los Angeles, California 90095, United States;

Department of Pharmaceutical Sciences, University of Michigan,

Ann Arbor, Michigan 48109, United States

Xing Jiang −School of Nursing, Nanjing University of Chinese

Medicine, Nanjing 210023, China

Moyuan Qu −The Affiliated Stomatology Hospital, Zhejiang

University School of Medicine, Key Laboratory of Oral

Biomedical Research of Zhejiang Province, Zhejiang University

School of Stomatology, Hangzhou 310006, China

George E. Aninwene, II −Department of Bioengineering and

Center for Minimally Invasive Therapeutics, University of

California, Los Angeles, Los Angeles, California 90095, United

States

Vadim Jucaud −Terasaki Institute for Biomedical Innovation,

Los Angeles, California 90064, United States

James J. Moon −Department of Pharmaceutical Sciences,

University of Michigan, Ann Arbor, Michigan 48109, United

States

Zhen Gu −Department of Bioengineering, Center for Minimally

Invasive Therapeutics, California NanoSystems Institute, and

Jonsson Comprehensive Cancer Center, University of California,

Los Angeles, Los Angeles, California 90095, United States

Complete contact information is available at:

https://pubs.acs.org/10.1021/acsnano.0c06109


Author Contributions

●These authors contributed equally to this work.

Notes

The authors declare no competing financial interest.

ACKNOWLEDGMENTS

This work was supported by National Institute of Health

(CA214411, GM126831, and AR069564).

VOCABULARY

Coronavirus disease 2019 (COVID-19), a disease caused by

severe acute respiratory syndrome coronavirus 2 (SARS-CoV-

2) and has been declared as a pandemic by WHO since March

11, 2020; antiviral vaccine, a biological preventive preparation

strategy to train the immune systems to fight against viral

pathogens; vaccine formulations, biomaterial engineering

strategies aiming to boost the efficacy or reduce side effects

of vaccines through co-delivering adjuvants, targeting immune

cells, or protecting vaccines activities; vaccine delivery

devices, tools for administering the vaccine formulations, and

they can be used by medical workers or the patients

themselves; pattern recognition receptors (PRRs), proteins

that can recognize pathogenic molecules as part of innate

immunity in response to virus infection, including identifying


pathogen-associated molecular patterns from a virus; adju-

vants, supplementary agents that could be co-delivered with

vaccines to boost the immune response, minimize the dosage

of antigens, and maintain longer immunity; lymph nodes,

round or bean-shaped clusters of cells containing abundant

immune cells that can initiate the robust immune response

toward viral infections


REFERENCES

(1) Bedford, J.; Enria, D.; Giesecke, J.; Heymann, D. L.; Ihekweazu,

C.; Kobinger, G.; Lane, H. C.; Memish, Z.; Oh, M.-d.; Sall, A. A.;

Schuchat, A.; Ungchusak, K.; Wieler, L. H. COVID-19: Towards

Controlling of a Pandemic. Lancet 2020, 395, 1015−1018.

(2) Berlin, D. A.; Gulick, R. M.; Martinez, F. J. Severe Covid-19. N.

Engl. J. Med. 2020, DOI: 10.1056/NEJMcp2009575.

(3) Rome, B. N.; Avorn, J. Drug Evaluation during the Covid-19

Pandemic. N. Engl. J. Med. 2020, 382, 2282−2284.

(4) Lurie, N.; Saville, M.; Hatchett, R.; Halton, J. Developing Covid-

19 Vaccines at Pandemic Speed. N. Engl. J. Med. 2020, 382, 1969−

1973.

(5) Fauci, A. S.; Morens, D. M. Zika Virus in the Americas Yet

Another Arbovirus Threat. N. Engl. J. Med. 2016, 374, 601−604.

(6) Regules, J. A.; Beigel, J. H.; Paolino, K. M.; Voell, J.; Castellano,

A. R.; Hu, Z.; Muñoz, P.; Moon, J. E.; Ruck, R. C.; Bennett, J. W.;

Twomey, P. S.; Gutie

́


rrez, R. L.; Remich, S. A.; Hack, H. R.;

Wisniewski, M. L.; Josleyn, M. D.; Kwilas, S. A.; Van Deusen, N.;

Mbaya, O. T.; Zhou, Y.; et al. A Recombinant Vesicular Stomatitis

Virus Ebola Vaccine. N. Engl. J. Med. 2017, 376, 330−341.

(7) Mulangu, S.; Dodd, L. E.; Davey, R. T.; Tshiani Mbaya, O.;

Proschan, M.; Mukadi, D.; Lusakibanza Manzo, M.; Nzolo, D.;

Tshomba Oloma, A.; Ibanda, A.; Ali, R.; Coulibaly, S.; Levine, A. C.;

Grais, R.; Diaz, J.; Lane, H. C.; Muyembe-Tamfum, J.-J.; The, P. W.

G. A Randomized, Controlled Trial of Ebola Virus Disease

Therapeutics. N. Engl. J. Med. 2019, 381, 2293−2303.

(8) Fineberg, H. V. Pandemic Preparedness and Response 

Lessons from the H1N1 Influenza of 2009. N. Engl. J. Med. 2014, 370,

1335−1342.

(9) Le, T. T.; Andreadakis, Z.; Kumar, A.; Roman, R. G.; Tollefsen,

S.; Saville, M.; Mayhew, S. The COVID-19 Vaccine Development

Landscape. Nat. Rev. Drug Discovery 2020, 19, 305−306.

(10) Draft Landscape of COVID-19 Candidate Vaccines. https://

www.who.int/publications/m/item/draft-landscape-of-covid-19-

candidate-vaccines (accessed 2020-08-28).

(11) Cox, R. J.; Brokstad, K. A. Not Just Antibodies: B Cells and T

Cells Mediate Immunity to COVID-19. Nat. Rev. Immunol. 2020,

581−582.

(12) Irvine, D. J.; Swartz, M. A.; Szeto, G. L. Engineering Synthetic

Vaccines Using Cues from Natural Immunity. Nat. Mater. 2013, 12,

978−990.

(13) Callaway, E. The Race for Coronavirus Vaccines: A Graphical

Guide. Nature 2020, 580, 576.

(14) Nabel, G. J. Designing Tomorrow’s Vaccines. N. Engl. J. Med.

2013, 368, 551−560.

(15) Gao, Q.; Bao, L.; Mao, H.; Wang, L.; Xu, K.; Yang, M.; Li, Y.;

Zhu, L.; Wang, N.; Lv, Z.; Gao, H.; Ge, X.; Kan, B.; Hu, Y.; Liu, J.;

Cai, F.; Jiang, D.; Yin, Y.; Qin, C.; Li, J.; et al. Development of an

Inactivated Vaccine Candidate for SARS-CoV-2. Science 2020, 369,

77−81.

(16) Ura, T.; Okuda, K.; Shimada, M. Developments in Viral Vector-

Based Vaccines. Vaccines 2014, 2, 624−641.

(17) Hos, B. J.; Tondini, E.; van Kasteren, S. I.; Ossendorp, F.

Approaches to Improve Chemically Defined Synthetic Peptide

Vaccines. Front. Immunol. 2018, 9, 884.

(18) Donlan, A. N.; Petri, W. A. Mucosal Immunity and the

Eradication of Polio. Science 2020, 368, 362−363.

(19) Macklin, G. R.; O’Reilly, K. M.; Grassly, N. C.; Edmunds, W. J.;

Mach, O.; Santhana Gopala Krishnan, R.; Voorman, A.; Vertefeuille, J.

F.; Abdelwahab, J.; Gumede, N.; Goel, A.; Sosler, S.; Sever, J.;


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12384



Bandyopadhyay, A. S.; Pallansch, M. A.; Nandy, R.; Mkanda, P.; Diop,

O. M.; Sutter, R. W. Evolving Epidemiology of Poliovirus Serotype 2

following Withdrawal of the Serotype 2 Oral Poliovirus Vaccine.

Science 2020, 368, 401−405.

(20) 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.; Revets, H.; Macadam, A.; Andino, R. Engineering the Live-

Attenuated Polio Vaccine to Prevent Reversion to Virulence. Cell Host

Microbe 2020, 27, 736−751.

(21) Robert-Guroff, M. Replicating and Non-Replicating Viral

Vectors for Vaccine Development. Curr. Opin. Biotechnol. 2007, 18,

546−556.

(22) 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.; Jia, S.-Y.; Jiang, H.-

D.; Wang, L.; Jiang, T.; Hu, Y.; Gou, J.-B.; Xu, S.-B.; Xu, J.-J.; Wang,

X.-W.; Wang, W.; 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.

(23) van Doremalen, N.; Haddock, E.; Feldmann, F.; Meade-White,

K.; Bushmaker, T.; Fischer, R. J.; Okumura, A.; Hanley, P. W.;

Saturday, G.; Edwards, N. J.; Clark, M. H. A.; Lambe, T.; Gilbert, S.

C.; Munster, V. J. A Single Dose of ChAdOx1MERS Provides

Protective Immunity in Rhesus Macaques. Sci. Adv. 2020, 6,

No. eaba8399.

(24) Gary, E. N.; Weiner, D. B. DNA Vaccines: Prime Time Is Now.

Curr. Opin. Immunol. 2020, 65, 21−27.

(25) Modjarrad, K.; Roberts, C. C.; Mills, K. T.; Castellano, A. R.;

Paolino, K.; Muthumani, K.; Reuschel, E. L.; Robb, M. L.; Racine, T.;

Oh, M.-d.; Lamarre, C.; Zaidi, F. I.; Boyer, J.; Kudchodkar, S. B.;

Jeong, M.; Darden, J. M.; Park, Y. K.; Scott, P. T.; Remigio, C.;

Parikh, A. P.; et al. Safety and Immunogenicity of an Anti-Middle East

Respiratory Syndrome Coronavirus DNA Vaccine: A Phase 1, Open-

Label, Single-Arm, Dose-Escalation Trial. Lancet Infect. Dis. 2019, 19,

1013−1022.

(26) Muthumani, K.; Falzarano, D.; Reuschel, E. L.; Tingey, C.;

Flingai, S.; Villarreal, D. O.; Wise, M.; Patel, A.; Izmirly, A.; Aljuaid,

A.; Seliga, A. M.; Soule, G.; Morrow, M.; Kraynyak, K. A.; Khan, A. S.;

Scott, D. P.; Feldmann, F.; LaCasse, R.; Meade-White, K.; Okumura,

A.; et al. A Synthetic Consensus Anti-Spike Protein DNA Vaccine

Induces Protective Immunity against Middle East Respiratory

Syndrome Coronavirus in Nonhuman Primates. Sci. Transl. Med.

2015, 7, 301−132.

(27) Pardi, N.; Hogan, M. J.; Porter, F. W.; Weissman, D. mRNA

Vaccines A New Era in Vaccinology. Nat. Rev. Drug Discovery

2018, 17, 261−279.

(28) Suschak, J. J.; Williams, J. A.; Schmaljohn, C. S. Advancements

in DNA Vaccine Vectors, Non-Mechanical Delivery Methods, and

Molecular Adjuvants to Increase Immunogenicity. Hum. Vaccines

Immunother. 2017, 13, 2837−2848.

(29) Gallie, D. R. The Cap and Poly (A) Tail Function

Synergistically to Regulate mRNA Translational Efficiency. Genes

Dev. 1991, 5, 2108−2116.

(30) Gustafsson, C.; Govindarajan, S.; Minshull, J. Codon Bias and

Heterologous Protein Expression. Trends Biotechnol. 2004, 22, 346−

353.

(31) Kariko,́ K.; Muramatsu, H.; Welsh, F. A.; Ludwig, J.; Kato, H.;

Akira, S.; Weissman, D. Incorporation of Pseudouridine into mRNA

Yields Superior Nonimmunogenic Vector with Increased Transla-

tional Capacity and Biological Stability. Mol. Ther. 2008, 16, 1833−

1840.

(32) Kariko,́ K.; Buckstein, M.; Ni, H.; Weissman, D. Suppression of

RNA Recognition by Toll-Like Receptors: the Impact of Nucleoside

Modification and the Evolutionary Origin of RNA. Immunity 2005,

23, 165−175.

(33) Chen, N.; Xia, P.; Li, S.; Zhang, T.; Wang, T. T.; Zhu, J. RNA

Sensors of the Innate Immune System and Their Detection of

Pathogens. IUBMB Life 2017, 69, 297−304.


(34) Kariko, K.; Muramatsu, H.; Ludwig, J.; Weissman, D.

Generating the Optimal mRNA for Therapy: HPLC Purification

Eliminates Immune Activation and Improves Translation of Nucleo-

side-Modified, Protein-Encoding mRNA. Nucleic Acids Res. 2011, 39,

142.

(35) Cohen, J. Designer Proteins Produce Potent HIV Defense.

Science 2015, 348, 1297.

(36) Bijker, M. S.; van den Eeden, S. J.; Franken, K. L.; Melief, C. J.;

Offringa, R.; van der Burg, S. H. CD8+ CTL Priming by Exact

Peptide Epitopes in Incomplete Freund’s Adjuvant Induces a

Vanishing CTL Response, Whereas Long Peptides Induce Sustained

CTL Reactivity. J. Immunol. 2007, 179, 5033−5040.

(37) Bijker, M. S.; van den Eeden, S. J.; Franken, K. L.; Melief, C. J.;

van der Burg, S. H.; Offringa, R. Superior Induction of Anti-Tumor

CTL Immunity by Extended Peptide Vaccines Involves Prolonged,

DC-Focused Antigen Presentation. Eur. J. Immunol. 2008, 38, 1033−

1042.

(38) Jackson, D. C.; Lau, Y. F.; Le, T.; Suhrbier, A.; Deliyannis, G.;

Cheers, C.; Smith, C.; Zeng, W.; Brown, L. E. A Totally Synthetic

Vaccine of Generic Structure That Targets Toll-Like Receptor 2 on

Dendritic Cells and Promotes Antibody or Cytotoxic T Cell

Responses. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 15440.

(39) Enayatkhani, M.; Hasaniazad, M.; Faezi, S.; Guklani, H.;

Davoodian, P.; Ahmadi, N.; Einakian, M. A.; Karmostaji, A.; Ahmadi,

K. Reverse Vaccinology Approach to Design a Novel Multi-Epitope

Vaccine Candidate against COVID-19: An In Silico Study. J. Biomol.

Struct. Dyn. 2020, 1−16.

(40) Kapsenberg, M. L. Dendritic-Cell Control of Pathogen-Driven

T-Cell Polarization. Nat. Rev. Immunol. 2003, 3, 984−993.

(41) Malowany, J. I.; McCormick, S.; Santosuosso, M.; Zhang, X.;

Aoki, N.; Ngai, P.; Wang, J.; Leitch, J.; Bramson, J.; Wan, Y.; Xing, Z.

Development of Cell-Based Tuberculosis Vaccines: Genetically

Modified Dendritic Cell Vaccine Is a Much More Potent Activator

of CD4 and CD8 T Cells Than Peptide- or Protein-Loaded

Counterparts. Mol. Ther. 2006, 13, 766−775.

(42) Benteyn, D.; Heirman, C.; Bonehill, A.; Thielemans, K.;

Breckpot, K. mRNA-Based Dendritic Cell Vaccines. Expert Rev.

Vaccines 2015, 14, 161−176.

(43) Perez, C. R.; De Palma, M. Engineering Dendritic Cell Vaccines

to Improve Cancer Immunotherapy. Nat. Commun. 2019, 10, 5408.

(44) Vela Ramirez, J. E.; Sharpe, L. A.; Peppas, N. A. Current State

and Challenges in Developing Oral Vaccines. Adv. Drug Delivery Rev.

2017, 114, 116−131.

(45) Melief, C. J. M.; van der Burg, S. H. Immunotherapy of

Established (Pre)malignant Disease by Synthetic Long Peptide

Vaccines. Nat. Rev. Cancer 2008, 8, 351−360.

(46) Napolitani, G.; Rinaldi, A.; Bertoni, F.; Sallusto, F.;

Lanzavecchia, A. Selected Toll-Like Receptor Agonist Combinations

Synergistically Trigger a T Helper Type 1-Polarizing Program in

Dendritic Cells. Nat. Immunol. 2005, 6, 769−776.

(47) Coffman, R. L.; Sher, A.; Seder, R. A. Vaccine Adjuvants:

Putting Innate Immunity to Work. Immunity 2010, 33, 492−503.

(48) Stephan, M. T.; Moon, J. J.; Um, S. H.; Bershteyn, A.; Irvine, D.

J. Therapeutic Cell Engineering with Surface-Conjugated Synthetic

Nanoparticles. Nat. Med. 2010, 16, 1035−1041.

(49) Yong, C. Y.; Ong, H. K.; Yeap, S. K.; Ho, K. L.; Tan, W. S.

Recent Advances in the Vaccine Development against Middle East

Respiratory Syndrome-Coronavirus. Front. Microbiol. 2019, 10, 1781.

(50) Hajj, K. A.; Whitehead, K. A. Tools for Translation: Non-Viral

Materials for Therapeutic mRNA Delivery. Nat. Rev. Mater. 2017, 2,

17056.

(51) Moyer, T. J.; Zmolek, A. C.; Irvine, D. J. Beyond Antigens and

Adjuvants: Formulating Future Vaccines. J. Clin. Invest. 2016, 126,

799−808.

(52) Sun, W.; Hu, Q.; Ji, W.; Wright, G.; Gu, Z. Leveraging

Physiology for Precision Drug Delivery. Physiol. Rev. 2017, 97, 189−

225.

(53) Moon, J. J.; Huang, B.; Irvine, D. J. Engineering Nano-and

Microparticles to Tune Immunity. Adv. Mater. 2012, 24, 3724−3746.


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12385



(54) Oberli, M. A.; Reichmuth, A. M.; Dorkin, J. R.; Mitchell, M. J.;

Fenton, O. S.; Jaklenec, A.; Anderson, D. G.; Langer, R.; Blankschtein,

D. Lipid Nanoparticle Assisted mRNA Delivery for Potent Cancer

Immunotherapy. Nano Lett. 2017, 17, 1326−1335.

(55) Moon, J. J.; Suh, H.; Bershteyn, A.; Stephan, M. T.; Liu, H.;

Huang, B.; Sohail, M.; Luo, S.; Ho Um, S.; Khant, H.; Goodwin, J. T.;

Ramos, J.; Chiu, W.; Irvine, D. J. Interbilayer-Crosslinked Multi-

lamellar Vesicles as Synthetic Vaccines for Potent Humoral and

Cellular Immune Responses. Nat. Mater. 2011, 10, 243−251.

(56) Kranz, L. M.; Diken, M.; Haas, H.; Kreiter, S.; Loquai, C.;

Reuter, K. C.; Meng, M.; Fritz, D.; Vascotto, F.; Hefesha, H.;

Grunwitz, C.; Vormehr, M.; Hüsemann, Y.; Selmi, A.; Kuhn, A. N.;

Buck, J.; Derhovanessian, E.; Rae, R.; Attig, S.; Diekmann, J.; et al.

Systemic RNA Delivery to Dendritic Cells Exploits Antiviral Defence

for Cancer Immunotherapy. Nature 2016, 534, 396−401.

(57) Irvine, D. J.; Read, B. J. Shaping Humoral Immunity to

Vaccines through Antigen-Displaying Nanoparticles. Curr. Opin.

Immunol. 2020, 65, 1−6.

(58) Kuai, R.; Sun, X.; Yuan, W.; Ochyl, L. J.; Xu, Y.; Hassani

Najafabadi, A.; Scheetz, L.; Yu, M.-Z.; Balwani, I.; Schwendeman, A.;

Moon, J. J. Dual TLR Agonist Nanodiscs as a Strong Adjuvant System

for Vaccines and Immunotherapy. J. Controlled Release 2018, 282,

131−139.

(59) Miao, L.; Li, L.; Huang, Y.; Delcassian, D.; Chahal, J.; Han, J.;

Shi, Y.; Sadtler, K.; Gao, W.; Lin, J.; Doloff, J. C.; Langer, R.;

Anderson, D. G. Delivery of mRNA Vaccines with Heterocyclic Lipids

Increases Anti-Tumor Efficacy by STING-Mediated Immune Cell

Activation. Nat. Biotechnol. 2019, 37, 1174−1185.

(60) Chahal, J. S.; Khan, O. F.; Cooper, C. L.; McPartlan, J. S.;

Tsosie, J. K.; Tilley, L. D.; Sidik, S. M.; Lourido, S.; Langer, R.; Bavari,

S.; Ploegh, H. L.; Anderson, D. G. Dendrimer-RNA Nanoparticles

Generate Protective Immunity against Lethal Ebola, H1N1 Influenza,

and Toxoplasma Gondii Challenges with a Single Dose. Proc. Natl.

Acad. Sci. U. S. A. 2016, 113, E4133−42.

(61) Zeng, Q.; Jiang, H.; Wang, T.; Zhang, Z.; Gong, T.; Sun, X.

Cationic Micelle Delivery of Trp2 Peptide for Efficient Lymphatic

Draining and Enhanced Cytotoxic T-Lymphocyte Responses. J.

Controlled Release 2015, 200, 1−12.

(62) Zhao, M.; Li, M.; Zhang, Z.; Gong, T.; Sun, X. Induction of

HIV-1 Gag Specific Immune Responses by Cationic Micelles

Mediated Delivery of Gag mRNA. Drug Delivery 2016, 23, 2596−

2607.

(63) Li, M.; Zhao, M.; Fu, Y.; Li, Y.; Gong, T.; Zhang, Z.; Sun, X.

Enhanced Intranasal Delivery of mRNA Vaccine by Overcoming the

Nasal Epithelial Barrier via Intra- and Paracellular Pathways. J.

Controlled Release 2016, 228, 9−19.

(64) Demento, S. L.; Cui, W.; Criscione, J. M.; Stern, E.; Tulipan, J.;

Kaech, S. M.; Fahmy, T. M. Role of Sustained Antigen Release from

Nanoparticle Vaccines in Shaping the T Cell Memory Phenotype.

Biomaterials 2012, 33, 4957−4964.

(65) Gu, P.; Wusiman, A.; Zhang, Y.; Liu, Z.; Bo, R.; Hu, Y.; Liu, J.;

Wang, D. Rational Design of PLGA Nanoparticle Vaccine Delivery

Systems to Improve Immune Responses. Mol. Pharmaceutics 2019, 16,

5000−5012.

(66) Chen, X.; Liu, Y.; Wang, L.; Liu, Y.; Zhang, W.; Fan, B.; Ma, X.;

Yuan, Q.; Ma, G.; Su, Z. Enhanced Humoral and Cell-Mediated

Immune Responses Generated by Cationic Polymer-Coated PLA

Microspheres with Adsorbed HBsAg. Mol. Pharmaceutics 2014, 11,

1772−1784.

(67) Chen, H.; Li, P.; Yin, Y.; Cai, X.; Huang, Z.; Chen, J.; Dong, L.;

Zhang, J. The Promotion of Type 1 T Helper Cell Responses to

Cationic Polymers in Vivo via Toll-Like Receptor-4 Mediated IL-12

Secretion. Biomaterials 2010, 31, 8172−8180.

(68) Fields, R. J.; Cheng, C. J.; Quijano, E.; Weller, C.; Kristofik, N.;

Duong, N.; Hoimes, C.; Egan, M. E.; Saltzman, W. M. Surface

Modified Poly(β Amino Ester)-Containing Nanoparticles for Plasmid

DNA Delivery. J. Controlled Release 2012, 164, 41−48.

(69) Kowalski, P. S.; Capasso Palmiero, U.; Huang, Y.; Rudra, A.;

Langer, R.; Anderson, D. G. Ionizable Amino-Polyesters Synthesized


via Ring Opening Polymerization of Tertiary Amino-Alcohols for

Tissue Selective mRNA Delivery. Adv. Mater. 2018, 30, 1801151.

(70) Fornaguera, C.; Guerra-Rebollo, M.; Ángel Laźaro, M.;

Castells-Sala, C.; Meca-Corte

́


s, O.; Ramos-Pe

́


rez, V.; Cascante, A.;


Rubio, N.; Blanco, J.; Borro

́


s, S. mRNA Delivery System for Targeting

Antigen-Presenting Cells in Vivo. Adv. Healthcare Mater. 2018, 7,

1800335.

(71) Liu, H.; Moynihan, K. D.; Zheng, Y.; Szeto, G. L.; Li, A. V.;

Huang, B.; Van Egeren, D. S.; Park, C.; Irvine, D. J. Structure-Based

Programming of Lymph-Node Targeting in Molecular Vaccines.

Nature 2014, 507, 519−522.

(72) Son, S.; Nam, J.; Zenkov, I.; Ochyl, L. J.; Xu, Y.; Scheetz, L.;

Shi, J.; Farokhzad, O. C.; Moon, J. J. Sugar-Nanocapsules Imprinted

with Microbial Molecular Patterns for mRNA Vaccination. Nano Lett.

2020, 20, 1499−1509.

(73) Moyer, T. J.; Kato, Y.; Abraham, W.; Chang, J. Y. H.; Kulp, D.

W.; Watson, N.; Turner, H. L.; Menis, S.; Abbott, R. K.; Bhiman, J.

N.; Melo, M. B.; Simon, H. A.; Herrera-De la Mata, S.; Liang, S.;

Seumois, G.; Agarwal, Y.; Li, N.; Burton, D. R.; Ward, A. B.; Schief,

W. R.; et al. Engineered Immunogen Binding to Alum Adjuvant

Enhances Humoral Immunity. Nat. Med. 2020, 26, 430−440.

(74) Moyano, D. F.; Goldsmith, M.; Solfiell, D. J.; Landesman-Milo,

D.; Miranda, O. R.; Peer, D.; Rotello, V. M. Nanoparticle

Hydrophobicity Dictates Immune Response. J. Am. Chem. Soc.

2012, 134, 3965−3967.

(75) Xu, L.; Liu, Y.; Chen, Z.; Li, W.; Liu, Y.; Wang, L.; Liu, Y.; Wu,

X.; Ji, Y.; Zhao, Y.; Ma, L.; Shao, Y.; Chen, C. Surface-Engineered

Gold Nanorods: Promising DNA Vaccine Adjuvant for HIV-1

Treatment. Nano Lett. 2012, 12, 2003−2012.

(76) Qiu, Y.; Liu, Y.; Wang, L.; Xu, L.; Bai, R.; Ji, Y.; Wu, X.; Zhao,

Y.; Li, Y.; Chen, C. Surface Chemistry and Aspect Ratio Mediated

Cellular Uptake of Au Nanorods. Biomaterials 2010, 31, 7606−7619.

(77) Zhou, Q.; Zhang, Y.; Du, J.; Li, Y.; Zhou, Y.; Fu, Q.; Zhang, J.;

Wang, X.; Zhan, L. Different-Sized Gold Nanoparticle Activator/

Antigen Increases Dendritic Cells Accumulation in Liver-Draining

Lymph Nodes and CD8+ T Cell Responses. ACS Nano 2016, 10,

2678−2692.

(78) Mercuri, L. P.; Carvalho, L. V.; Lima, F. A.; Quayle, C.; Fantini,

M. C. A.; Tanaka, G. S.; Cabrera, W. H.; Furtado, M. F. D.;

Tambourgi, D. V.; Matos, J. d. R.; Jaroniec, M.; Sant’Anna, O. A.

Ordered Mesoporous Silica SBA-15: A New Effective Adjuvant to

Induce Antibody Response. Small 2006, 2, 254−256.

(79) Mahony, D.; Cavallaro, A. S.; Mody, K. T.; Xiong, L.; Mahony,

T. J.; Qiao, S. Z.; Mitter, N. In Vivo Delivery of Bovine Viral

Diahorrea Virus, E2 Protein Using Hollow Mesoporous Silica

Nanoparticles. Nanoscale 2014, 6, 6617−6626.

(80) Hess, K. L.; Oh, E.; Tostanoski, L. H.; Andorko, J. I.; Susumu,

K.; Deschamps, J. R.; Medintz, I. L.; Jewell, C. M. Engineering

Immunological Tolerance Using Quantum Dots to Tune the Density

of Self-Antigen Display. Adv. Funct. Mater. 2017, 27, 1700290.

(81) Hu, Z.; Song, B.; Xu, L.; Zhong, Y.; Peng, F.; Ji, X.; Zhu, F.;

Yang, C.; Zhou, J.; Su, Y.; Chen, S.; He, Y.; He, S. Aqueous

Synthesized Quantum Dots Interfere with the NF-κB Pathway and

Confer Anti-Tumor, Anti-Viral and Anti-Inflammatory Effects.

Biomaterials 2016, 108, 187−196.

(82) Faria, P. C. B. d.; Santos, L. I. d.; Coelho, J. P.; Ribeiro, H. B.;

Pimenta, M. A.; Ladeira, L. O.; Gomes, D. A.; Furtado, C. A.;

Gazzinelli, R. T. Oxidized Multiwalled Carbon Nanotubes as Antigen

Delivery System to Promote Superior CD8+ T Cell Response and

Protection against Cancer. Nano Lett. 2014, 14, 5458−5470.

(83) Zhang, L.; Chan, J. M.; Gu, F. X.; Rhee, J.-W.; Wang, A. Z.;

Radovic-Moreno, A. F.; Alexis, F.; Langer, R.; Farokhzad, O. C. Self-

Assembled Lipid-Polymer Hybrid Nanoparticles: A Robust Drug

Delivery Platform. ACS Nano 2008, 2, 1696−1702.

(84) Walsh, C. L.; Nguyen, J.; Tiffany, M. R.; Szoka, F. C. Synthesis,

Characterization, and Evaluation of Ionizable Lysine-Based Lipids for

siRNA Delivery. Bioconjugate Chem. 2013, 24, 36−43.


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12386



(85) Jokerst, J. V.; Lobovkina, T.; Zare, R. N.; Gambhir, S. S.

Nanoparticle PEGylation for Imaging and Therapy. Nanomedicine

2011, 6, 715−728.

(86) Beltra

́


n-Gracia, E.; Lo

́


pez-Camacho, A.; Higuera-Ciapara, I.;


Vela

́


zquez-Ferna

́


ndez, J. B.; Vallejo-Cardona, A. A. Nanomedicine

Review: Clinical Developments in Liposomal Applications. Cancer

Nanotechnol. 2019, 10, 11.

(87) 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.;

Sanders, R. W.; Moore, J. P.; Crotty, S.; Langer, R.; Anderson, D. G.;

Chakraborty, A. K.; Irvine, D. J. Sustained Antigen Availability during

Germinal Center Initiation Enhances Antibody Responses to

Vaccination. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, E6639−E6648.

(88) Diebold, S. S.; Kaisho, T.; Hemmi, H.; Akira, S.; Reis e Sousa,

C. Innate Antiviral Responses by Means of TLR7-Mediated

Recognition of Single-Stranded RNA. Science 2004, 303, 1529−1531.

(89) Kuai, R.; Ochyl, L. J.; Bahjat, K. S.; Schwendeman, A.; Moon, J.

J. Designer Vaccine Nanodiscs for Personalized Cancer Immunother-

apy. Nat. Mater. 2017, 16, 489−496.

(90) Dobrovolskaia, M. A.; McNeil, S. E. Immunological Properties

of Engineered Nanomaterials. Nat. Nanotechnol. 2007, 2, 469−478.

(91) Akinc, A.; Thomas, M.; Klibanov, A. M.; Langer, R. Exploring

Polyethylenimine-Mediated DNA Transfection and the Proton

Sponge Hypothesis. J. Gene Med. 2005, 7, 657−663.

(92) Boussif, O.; Lezoualch, F.; Zanta, M. A.; Mergny, M. D.;

Scherman, D.; Demeneix, B.; Behr, J. P. A Versatile Vector for Gene

and Oligonucleotide Transfer into Cells in Culture and in Vivo:

Polyethylenimine. Proc. Natl. Acad. Sci. U. S. A. 1995, 92, 7297−7301.

(93) Anderson, J. M.; Shive, M. S. Biodegradation and

Biocompatibility of PLA and PLGA Microspheres. Adv. Drug Delivery

Rev. 2012, 64, 72−82.

(94) Sunshine, J. C.; Perica, K.; Schneck, J. P.; Green, J. J. Particle

Shape Dependence of CD8+ T Cell Activation by Artificial Antigen

Presenting Cells. Biomaterials 2014, 35, 269−277.

(95) Wang, Q.; Tan, M. T.; Keegan, B. P.; Barry, M. A.; Heffernan,

M. J. Time Course Study of the Antigen-Specific Immune Response

to a PLGA Microparticle Vaccine Formulation. Biomaterials 2014, 35,

8385−8393.

(96) Chan, J. M.; Zhang, L.; Yuet, K. P.; Liao, G.; Rhee, J.-W.;

Langer, R.; Farokhzad, O. C. PLGA-Lecithin-PEG Core-Shell

Nanoparticles for Controlled Drug Delivery. Biomaterials 2009, 30,

1627−1634.

(97) Sun, W.; Wang, J.; Hu, Q.; Zhou, X.; Khademhosseini, A.; Gu,

Z. CRISPR-Cas12a Delivery by DNA-Mediated Bioresponsive Editing

for Cholesterol Regulation. Sci. Adv. 2020, 6, No. eaba2983.

(98) Dong, L.; Xia, S.; Chen, H.; Chen, J.; Zhang, J. Anti-Arthritis

Activity of Cationic Materials. J. Cell. Mol. Med. 2010, 14, 2015−2024.

(99) Little, S. R.; Lynn, D. M.; Ge, Q.; Anderson, D. G.; Puram, S.

V.; Chen, J.; Eisen, H. N.; Langer, R. Poly-β Amino Ester-Containing

Microparticles Enhance the Activity of Nonviral Genetic Vaccines.

Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 9534−9539.

(100) Liu, Z.; Jiao, Y.; Wang, Y.; Zhou, C.; Zhang, Z.

Polysaccharides-Based Nanoparticles as Drug Delivery Systems. Adv.

Drug Delivery Rev. 2008, 60, 1650−1662.

(101) Acharya, A. P.; Sinha, M.; Ratay, M. L.; Ding, X.; Balmert, S.

C.; Workman, C. J.; Wang, Y.; Vignali, D. A. A.; Little, S. R. Localized

Multi-Component Delivery Platform Generates Local and Systemic

Anti-Tumor Immunity. Adv. Funct. Mater. 2017, 27, 1604366.

(102) Dosta, P.; Segovia, N.; Cascante, A.; Ramos, V.; Borro

́


s, S.

Surface Charge Tunability as a Powerful Strategy to Control

Electrostatic Interaction for High Efficiency Silencing, Using Tailored

Oligopeptide-Modified Poly(Beta-Amino Ester)s (PBAEs). Acta

Biomater. 2015, 20, 82−93.

(103) Tsopelas, C.; Sutton, R. Why Certain Dyes Are Useful for

Localizing the Sentinel Lymph Node. J. Nucl. Med. 2002, 43, 1377−

1382.

(104) Seong, S.-Y.; Matzinger, P. Hydrophobicity: An Ancient

Damage-Associated Molecular Pattern That Initiates Innate Immune

Responses. Nat. Rev. Immunol. 2004, 4, 469−478.


(105) Shokouhi, B.; Coban, C.; Hasirci, V.; Aydin, E.; Dhanasingh,

A.; Shi, N.; Koyama, S.; Akira, S.; Zenke, M.; Sechi, A. S. The Role of

Multiple Toll-Like Receptor Signalling Cascades on Interactions

between Biomedical Polymers and Dendritic Cells. Biomaterials 2010,

31, 5759−5771.

(106) Glenny, A.; Pope, C.; Waddington, H.; Wallace, U.

Immunological Notes. xvii-xxiv. J. Pathol. Bacteriol. 1926, 29, 31−40.

(107) Levitz, S. M.; Golenbock, D. T. Beyond Empiricism:

Informing Vaccine Development through Innate Immunity Research.

Cell 2012, 148, 1284−1292.

(108) Mannhalter, J.; Neychev, H.; Zlabinger, G.; Ahmad, R.; Eibl,

M. Modulation of the Human Immune Response by the Non-Toxic

and Non-Pyrogenic Adjuvant Aluminium Hydroxide: Effect on

Antigen Uptake and Antigen Presentation. Clin. Exp. Immunol.

1985, 61, 143−151.

(109) McKee, A. S.; Munks, M. W.; Marrack, P. How Do Adjuvants

Work? Important Considerations for New Generation Adjuvants.

Immunity 2007, 27, 687−690.

(110) Hornung, V.; Bauernfeind, F.; Halle, A.; Samstad, E. O.; Kono,

H.; Rock, K. L.; Fitzgerald, K. A.; Latz, E. Silica Crystals and

Aluminum Salts Activate the NALP3 Inflammasome through

Phagosomal Destabilization. Nat. Immunol. 2008, 9, 847−856.

(111) Hess, K. L.; Medintz, I. L.; Jewell, C. M. Designing Inorganic

Nanomaterials for Vaccines and Immunotherapies. Nano Today 2019,

27, 73−98.

(112) Reddy, S. T.; van der Vlies, A. J.; Simeoni, E.; Angeli, V.;

Randolph, G. J.; O’Neil, C. P.; Lee, L. K.; Swartz, M. A.; Hubbell, J. A.

Exploiting Lymphatic Transport and Complement Activation in

Nanoparticle Vaccines. Nat. Biotechnol. 2007, 25, 1159−1164.

(113) Sokolova, V.; Epple, M. Inorganic Nanoparticles as Carriers of

Nucleic Acids into Cells. Angew. Chem., Int. Ed. 2008, 47, 1382−1395.

(114) Moon, J. J.; Suh, H.; Li, A. V.; Ockenhouse, C. F.; Yadava, A.;

Irvine, D. J. Enhancing Humoral Responses to a Malaria Antigen with

Nanoparticle Vaccines That Expand Tfh Cells and Promote Germinal

Center Induction. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 1080−

1085.

(115) Lin, W.-H. W.; Kouyos, R. D.; Adams, R. J.; Grenfell, B. T.;

Griffin, D. E. Prolonged Persistence of Measles Virus RNA Is

Characteristic of Primary Infection Dynamics. Proc. Natl. Acad. Sci. U.

S. A. 2012, 109, 14989−14994.

(116) Pape, K. A.; Catron, D. M.; Itano, A. A.; Jenkins, M. K. The

Humoral Immune Response Is Initiated in Lymph Nodes by B Cells

That Acquire Soluble Antigen Directly in the Follicles. Immunity

2007, 26, 491−502.

(117) McHugh, K. J.; Guarecuco, R.; Langer, R.; Jaklenec, A. Single-

Injection Vaccines: Progress, Challenges, and Opportunities. J.

Controlled Release 2015, 219, 596−609.

(118) Zhu, Q.; Zarnitsyn, V. G.; Ye, L.; Wen, Z.; Gao, Y.; Pan, L.;

Skountzou, I.; Gill, H. S.; Prausnitz, M. R.; Yang, C.; Compans, R. W.

Immunization by Vaccine-Coated Microneedle Arrays Protects

against Lethal Influenza Virus Challenge. Proc. Natl. Acad. Sci. U. S.

A. 2009, 106, 7968−7973.

(119) Sullivan, S. P.; Koutsonanos, D. G.; del Pilar Martin, M.; Lee,

J. W.; Zarnitsyn, V.; Choi, S.-O.; Murthy, N.; Compans, R. W.;

Skountzou, I.; Prausnitz, M. R. Dissolving Polymer Microneedle

Patches for Influenza Vaccination. Nat. Med. 2010, 16, 915−920.

(120) Kim, E.; Erdos, G.; Huang, S.; Kenniston, T. W.; Balmert, S.

C.; Carey, C. D.; Raj, V. S.; Epperly, M. W.; Klimstra, W. B.;

Haagmans, B. L.; Korkmaz, E.; Falo, L. D.; Gambotto, A. Microneedle

Array Delivered Recombinant Coronavirus Vaccines: Immunogenicity

and Rapid Translational Development. EBioMedicine 2020, 55,

102743.

(121) Rouphael, N. G.; Paine, M.; Mosley, R.; Henry, S.; McAllister,

D. V.; Kalluri, H.; Pewin, W.; Frew, P. M.; Yu, T.; Thornburg, N. J.;

Kabbani, S.; Lai, L.; Vassilieva, E. V.; Skountzou, I.; Compans, R. W.;

Mulligan, M. J.; Prausnitz, M. R.; Beck, A.; Edupuganti, S.; Heeke, S.;

et al. The Safety, Immunogenicity, and Acceptability of Inactivated

Influenza Vaccine Delivered by Microneedle Patch (TIV-MNP


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12387



2015): A Randomised, Partly Blinded, Placebo-Controlled, Phase 1

Trial. Lancet 2017, 390, 649−658.

(122) Boopathy, A. V.; Mandal, A.; Kulp, D. W.; Menis, S.; Bennett,

N. R.; Watkins, H. C.; Wang, W.; Martin, J. T.; Thai, N. T.; He, Y.;

Schief, W. R.; Hammond, P. T.; Irvine, D. J. Enhancing Humoral

Immunity via Sustained-Release Implantable Microneedle Patch

Vaccination. Proc. Natl. Acad. Sci. U. S. A. 2019, 116, 16473−16478.

(123) DeMuth, P. C.; Moon, J. J.; Suh, H.; Hammond, P. T.; Irvine,

D. J. Releasable Layer-by-Layer Assembly of Stabilized Lipid

Nanocapsules on Microneedles for Enhanced Transcutaneous

Vaccine Delivery. ACS Nano 2012, 6, 8041−8051.

(124) Zaric, M.; Lyubomska, O.; Touzelet, O.; Poux, C.; Al-Zahrani,

S.; Fay, F.; Wallace, L.; Terhorst, D.; Malissen, B.; Henri, S.; Power,

U. F.; Scott, C. J.; Donnelly, R. F.; Kissenpfennig, A. Skin Dendritic

Cell Targeting via Microneedle Arrays Laden with Antigen-

Encapsulated Poly-d, l-lactide-co-Glycolide Nanoparticles Induces

Efficient Antitumor and Antiviral Immune Responses. ACS Nano

2013, 7, 2042−2055.

(125) Bae, W.-G.; Ko, H.; So, J.-Y.; Yi, H.; Lee, C.-H.; Lee, D.-H.;

Ahn, Y.; Lee, S.-H.; Lee, K.; Jun, J.; Kim, H.-H.; Jeon, N. L.; Jung, W.;

Song, C.-S.; Kim, T.; Kim, Y.-C.; Jeong, H. E. Snake Fang-Inspired

Stamping Patch for Transdermal Delivery of Liquid Formulations. Sci.

Transl. Med. 2019, 11, No. eaaw3329.

(126) Tzeng, S. Y.; McHugh, K. J.; Behrens, A. M.; Rose, S.;

Sugarman, J. L.; Ferber, S.; Langer, R.; Jaklenec, A. Stabilized Single-

Injection Inactivated Polio Vaccine Elicits a Strong Neutralizing

Immune Response. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, E5269−

E5278.

(127) McHugh, K. J.; Nguyen, T. D.; Linehan, A. R.; Yang, D.;

Behrens, A. M.; Rose, S.; Tochka, Z. L.; Tzeng, S. Y.; Norman, J. J.;

Anselmo, A. C.; Xu, X.; Tomasic, S.; Taylor, M. A.; Lu, J.; Guarecuco,

R.; Langer, R.; Jaklenec, A. Fabrication of Fillable Microparticles and

Other Complex 3D Microstructures. Science 2017, 357, 1138−1142.

(128) Lu, X.; Miao, L.; Gao, W.; Chen, Z.; McHugh, K. J.; Sun, Y.;

Tochka, Z.; Tomasic, S.; Sadtler, K.; Hyacinthe, A.; Huang, Y.; Graf,

T.; Hu, Q.; Sarmadi, M.; Langer, R.; Anderson, D. G.; Jaklenec, A.

Engineered PLGA Microparticles for Long-Term, Pulsatile Release of

STING Agonist for Cancer Immunotherapy. Sci. Transl. Med. 2020,

12, No. eaaz6606.

(129) Ali, O. A.; Huebsch, N.; Cao, L.; Dranoff, G.; Mooney, D. J.

Infection-Mimicking Materials to Program Dendritic Cells in Situ.

Nat. Mater. 2009, 8, 151−158.

(130) Kim, J.; Li, W. A.; Choi, Y.; Lewin, S. A.; Verbeke, C. S.;

Dranoff, G.; Mooney, D. J. Injectable, Spontaneously Assembling,

Inorganic Scaffolds Modulate Immune Cells in Vivo and Increase

Vaccine Efficacy. Nat. Biotechnol. 2015, 33, 64−72.

(131) Li, A. W.; Sobral, M. C.; Badrinath, S.; Choi, Y.; Graveline, A.;

Stafford, A. G.; Weaver, J. C.; Dellacherie, M. O.; Shih, T.-Y.; Ali, O.

A.; Kim, J.; Wucherpfennig, K. W.; Mooney, D. J. A Facile Approach

to Enhance Antigen Response for Personalized Cancer Vaccination.

Nat. Mater. 2018, 17, 528−534.

(132) Dellacherie, M. O.; Li, A.; Lu, B. Y.; Verbeke, C. S.; Gu, L.;

Stafford, A. G.; Doherty, E. J.; Mooney, D. J. Single-Shot Mesoporous

Silica Rods Scaffold for Induction of Humoral Responses against

Small Antigens. Adv. Funct. Mater. 2020, 30, 2002448.

(133) Shih, T.-Y.; Blacklow, S. O.; Li, A. W.; Freedman, B. R.;

Bencherif, S.; Koshy, S. T.; Darnell, M. C.; Mooney, D. J. Injectable,

Tough Alginate Cryogels as Cancer Vaccines. Adv. Healthcare Mater.

2018, 7, 1701469.

(134) Bencherif, S. A.; Sands, R. W.; Bhatta, D.; Arany, P.; Verbeke,

C. S.; Edwards, D. A.; Mooney, D. J. Injectable Preformed Scaffolds

with Shape-Memory Properties. Proc. Natl. Acad. Sci. U. S. A. 2012,

109, 19590−19595.

(135) Hori, Y.; Winans, A. M.; Huang, C. C.; Horrigan, E. M.;

Irvine, D. J. Injectable Dendritic Cell-Carrying Alginate Gels for

Immunization and Immunotherapy. Biomaterials 2008, 29, 3671−

3682.


(136) Guo, Z.; Zhang, M.; Tang, H.; Cao, X. Fas Signal Links Innate

and Adaptive Immunity by Promoting Dendritic-Cell Secretion of CC

and CXC Chemokines. Blood 2005, 106, 2033−2041.

(137) Hori, Y.; Winans, A. M.; Irvine, D. J. Modular Injectable

Matrices Based on Alginate Solution/Microsphere Mixtures That Gel

in Situ and Co-deliver Immunomodulatory factors. Acta Biomater.

2009, 5, 969−982.

(138) Roy, K.; Wang, D.; Hedley, M. L.; Barman, S. P. Gene

Delivery with In-Situ Crosslinking Polymer Networks Generates

Long-Term Systemic Protein Expression. Mol. Ther. 2003, 7, 401−

408.

(139) Wu, Y.; Wei, W.; Zhou, M.; Wang, Y.; Wu, J.; Ma, G.; Su, Z.

Thermal-Sensitive Hydrogel as Adjuvant-Free Vaccine Delivery

System for H5N1 Intranasal Immunization. Biomaterials 2012, 33,

2351−2360.

(140) Hiemstra, C.; van der Aa, L. J.; Zhong, Z.; Dijkstra, P. J.;

Feijen, J. Rapidly in Situ-Forming Degradable Hydrogels from

Dextran Thiols through Michael Addition. Biomacromolecules 2007,

8, 1548−1556.

(141) Singh, A.; Suri, S.; Roy, K. In-Situ Crosslinking Hydrogels for

Combinatorial Delivery of Chemokines and siRNA-DNA Carrying

Microparticles to Dendritic Cells. Biomaterials 2009, 30, 5187−5200.

(142) Singh, A.; Qin, H.; Fernandez, I.; Wei, J.; Lin, J.; Kwak, L. W.;

Roy, K. An Injectable Synthetic Immune-Priming Center Mediates

Efficient T-Cell Class Switching and T-Helper 1 Response against B

Cell Lymphoma. J. Controlled Release 2011, 155, 184−192.

(143) Umeki, Y.; Mohri, K.; Kawasaki, Y.; Watanabe, H.; Takahashi,

R.; Takahashi, Y.; Takakura, Y.; Nishikawa, M. Induction of Potent

Antitumor Immunity by Sustained Release of Cationic Antigen from a

DNA-Based Hydrogel with Adjuvant Activity. Adv. Funct. Mater.

2015, 25, 5758−5767.

(144) Prausnitz, M. R.; Mikszta, J. A.; Cormier, M.; Andrianov, A. K.

Microneedle-Based Vaccines. Curr. Top. Microbiol. Immunol. 2009,

333, 369−393.

(145) Chen, G.; Chen, Z.; Wen, D.; Wang, Z.; Li, H.; Zeng, Y.;

Dotti, G.; Wirz, R. E.; Gu, Z. Transdermal Cold Atmospheric Plasma-

Mediated Immune Checkpoint Blockade Therapy. Proc. Natl. Acad.

Sci. U. S. A. 2020, 117, 3687−3692.

(146) Ye, Y.; Wang, C.; Zhang, X.; Hu, Q.; Zhang, Y.; Liu, Q.; Wen,

D.; Milligan, J.; Bellotti, A.; Huang, L.; Dotti, G.; Gu, Z. A Melanin-

Mediated Cancer Immunotherapy Patch. Sci. Immunol. 2017, 2,

No. eaan5692.

(147) Erdos, G.; Balmert, S. C.; Carey, C. D.; Falo, G. D.; Patel, N.

A.; Zhang, J.; Gambotto, A.; Korkmaz, E.; Falo, L. D. Improved

Cutaneous Genetic Immunization by Microneedle Array Delivery of

an Adjuvanted Adenovirus Vaccine. J. Invest. Dermatol. 2020,

DOI: 10.1016/j.jid.2020.03.966.

(148) Lee, K.; Goudie, M. J.; Tebon, P.; Sun, W.; Luo, Z.; Lee, J.;

Zhang, S.; Fetah, K.; Kim, H.-J.; Xue, Y.; Darabi, M. A.; Ahadian, S.;

Sarikhani, E.; Ryu, W.; Gu, Z.; Weiss, P. S.; Dokmeci, M. R.;

Ashammakhi, N.; Khademhosseini, A., Non-Transdermal Micro-

needles for Advanced Drug Delivery. Adv. Drug Delivery Rev. 2019,

DOI: 10.1016/j.addr.2019.11.010.

(149) van der Maaden, K.; Jiskoot, W.; Bouwstra, J. Microneedle

Technologies for (Trans)dermal Drug and Vaccine Delivery. J.

Controlled Release 2012, 161, 645−655.

(150) Kim, Y.-C.; Quan, F.-S.; Compans, R. W.; Kang, S.-M.;

Prausnitz, M. R. Stability Kinetics of Influenza Vaccine Coated onto

Microneedles during Drying and Storage. Pharm. Res. 2011, 28, 135−

144.

(151) Sanders, R. W.; van Gils, M. J.; Derking, R.; Sok, D.; Ketas, T.

J.; Burger, J. A.; Ozorowski, G.; Cupo, A.; Simonich, C.; Goo, L.;

Arendt, H.; Kim, H. J.; Lee, J. H.; Pugach, P.; Williams, M.; Debnath,

G.; Moldt, B.; van Breemen, M. J.; Isik, G.; Medina-Ramírez, M.; et al.

HIV-1 Neutralizing Antibodies Induced by Native-Like Envelope

Trimers. Science 2015, 349, No. aac4223.

(152) Wang, W. Lyophilization and Development of Solid Protein

Pharmaceuticals. Int. J. Pharm. 2000, 203, 1−60.


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12388



(153) Dellacherie, M. O.; Seo, B. R.; Mooney, D. J. Macroscale

Biomaterials Strategies for Local Immunomodulation. Nat. Rev. Mater.

2019, 4, 379−397.

(154) Wu, Y.; Jane Grande-Allen, K.; West, J. L. Adhesive Peptide

Sequences Regulate Valve Interstitial Cell Adhesion, Phenotype and

Extracellular Matrix Deposition. Cell. Mol. Bioeng. 2016, 9, 479−495.

(155) Diehl, M. C.; Lee, J. C.; Daniels, S. E.; Tebas, P.; Khan, A. S.;

Giffear, M.; Sardesai, N. Y.; Bagarazzi, M. L. Tolerability of

Intramuscular and Intradermal Delivery by CELLECTRA® Adaptive

Constant Current Electroporation Device in Healthy Volunteers.

Hum. Vaccines Immunother. 2013, 9, 2246−2252.

(156) Smith, T. R. F.; Patel, A.; Ramos, S.; Elwood, D.; Zhu, X.; Yan,

J.; Gary, E. N.; Walker, S. N.; Schultheis, K.; Purwar, M.; Xu, Z.;

Walters, J.; Bhojnagarwala, P.; Yang, M.; Chokkalingam, N.; Pezzoli,

P.; Parzych, E.; Reuschel, E. L.; Doan, A.; Tursi, N.; et al.

Immunogenicity of a DNA Vaccine Candidate for COVID-19. Nat.

Commun. 2020, 11, 2601.

(157) Choi, S.-O.; Kim, Y.-C.; Lee, J. W.; Park, J.-H.; Prausnitz, M.

R.; Allen, M. G. Intracellular Protein Delivery and Gene Transfection

by Electroporation Using a Microneedle Electrode Array. Small 2012,

8, 1081−1091.

(158) Banga, A. K.; Bose, S.; Ghosh, T. K. Iontophoresis and

Electroporation: Comparisons and Contrasts. Int. J. Pharm. 1999,

179, 1−19.

(159) Tadros, A. R.; Romanyuk, A.; Miller, I. C.; Santiago, A.; Noel,

R. K.; O’Farrell, L.; Kwong, G. A.; Prausnitz, M. R. STAR Particles for

Enhanced Topical Drug and Vaccine Delivery. Nat. Med. 2020, 26,

341−347.

(160) Lamson, N. G.; Berger, A.; Fein, K. C.; Whitehead, K. A.

Anionic Nanoparticles Enable the Oral Delivery of Proteins by

Enhancing Intestinal Permeability. Nat. Biomed. Eng. 2020, 4, 84−96.

(161) Wei, X.; Beltra

́


n-Gaste

́


lum, M.; Karshalev, E.; Esteban-


Ferna

́


ndez de Ávila, B.; Zhou, J.; Ran, D.; Angsantikul, P.; Fang, R.

H.; Wang, J.; Zhang, L. Biomimetic Micromotor Enables Active

Delivery of Antigens for Oral Vaccination. Nano Lett. 2019, 19,

1914−1921.

(162) Karshalev, E.; Esteban-Ferna

́


ndez de Ávila, B.; Beltra

́


n-


Gaste

́


lum, M.; Angsantikul, P.; Tang, S.; Mundaca-Uribe, R.; Zhang,

F.; Zhao, J.; Zhang, L.; Wang, J. Micromotor Pills as a Dynamic Oral

Delivery Platform. ACS Nano 2018, 12, 8397−8405.

(163) Aran, K.; Chooljian, M.; Paredes, J.; Rafi, M.; Lee, K.; Kim, A.

Y.; An, J.; Yau, J. F.; Chum, H.; Conboy, I.; Murthy, N.; Liepmann, D.

An Oral Microjet Vaccination System Elicits Antibody Production in

Rabbits. Sci. Transl. Med. 2017, 9, No. eaaf6413.

(164) Abramson, A.; Caffarel-Salvador, E.; Khang, M.; Dellal, D.;

Silverstein, D.; Gao, Y.; Frederiksen, M. R.; Vegge, A.; Huba

́


lek, F.;

Water, J. J.; Friderichsen, A. V.; Fels, J.; Kirk, R. K.; Cleveland, C.;

Collins, J.; Tamang, S.; Hayward, A.; Landh, T.; Buckley, S. T.;

Roxhed, N.; et al. An Ingestible Self-Orienting System for Oral

Delivery of Macromolecules. Science 2019, 363, 611−615.

(165) Shin, M. D.; Shukla, S.; Chung, Y. H.; Beiss, V.; Chan, S. K.;

Ortega-Rivera, O. A.; Wirth, D. M.; Chen, A.; Sack, M.; Pokorski, J.

K.; Steinmetz, N. F. COVID-19 Vaccine Development and a Potential

Nanomaterial Path Forward. Nat. Nanotechnol. 2020, 15, 646−655.

(166) Weiss, C.; Carriere, M.; Fusco, L.; Capua, I.; Regla-Nava, J. A.;

Pasquali, M.; Scott, J. A.; Vitale, F.; Unal, M. A.; Mattevi, C.;

Bedognetti, D.; Merkoç


i, A.; Tasciotti, E.; Yilmazer, A.; Gogotsi, Y.;

Stellacci, F.; Delogu, L. G. Toward Nanotechnology-Enabled

Approaches against the COVID-19 Pandemic. ACS Nano 2020, 14,

6383−6406.

(167) Fang, J.; Hsueh, Y.-Y.; Soto, J.; Sun, W.; Wang, J.; Gu, Z.;

Khademhosseini, A.; Li, S. Engineering Biomaterials with Micro/

Nanotechnologies for Cell Reprogramming. ACS Nano 2020, 14,

1296−1318.

(168) Sun, J.; Zhuang, Z.; Zheng, J.; Li, K.; Wong, R. L.-Y.; Liu, D.;

Huang, J.; He, J.; Zhu, A.; Zhao, J.; Li, X.; Xi, Y.; Chen, R.; Alshukairi,

A. N.; Chen, Z.; Zhang, Z.; Chen, C.; Huang, X.; Li, F.; Lai, X.; et al.

Generation of a Broadly Useful Model for COVID-19 Pathogenesis,

Vaccination, and Treatment. Cell 2020, 182, 734−743.


(169) Si, L.; Bai, H.; Rodas, M.; Cao, W.; Oh, C. Y.; Jiang, A.;

Nurani, A.; Zhu, D. Y.; Goyal, G.; Gilpin, S. E.; Prantil-Baun, R.;

Ingber, D. E., Human Organs-on-Chips as Tools for Repurposing

Approved Drugs as Potential Influenza and COVID19 Therapeutics

in Viral Pandemics. bioRxiv, April 14, 2020, ver. 1. https://www.

biorxiv.org/content/10.1101/2020.04.13.039917v1 (accessed 2020-

08-25).

(170) Sun, W.; Luo, Z.; Lee, J.; Kim, H.-J.; Lee, K.; Tebon, P.; Feng,

Y.; Dokmeci, M. R.; Sengupta, S.; Khademhosseini, A. Organ-on-a-

Chip for Cancer and Immune Organs Modeling. Adv. Healthcare

Mater. 2019, 8, 1801363.

(171) Segaliny, A. I.; Li, G.; Kong, L.; Ren, C.; Chen, X.; Wang, J.

K.; Baltimore, D.; Wu, G.; Zhao, W. Functional TCR T Cell

Screening Using Single-Cell Droplet Microfluidics. Lab Chip 2018,

18, 3733−3749.

(172) Monteil, V.; Kwon, H.; Prado, P.; Hagelkrüys, A.; Wimmer, R.

A.; Stahl, M.; Leopoldi, A.; Garreta, E.; Hurtado del Pozo, C.;

Prosper, F.; Romero, J. P.; Wirnsberger, G.; Zhang, H.; Slutsky, A. S.;

Conder, R.; Montserrat, N.; Mirazimi, A.; Penninger, J. M. Inhibition

of SARS-CoV-2 Infections in Engineered Human Tissues Using

Clinical-Grade Soluble Human ACE2. Cell 2020, 181, 905−913.

(173) Shah, S. K.; Miller, F. G.; Darton, T. C.; Duenas, D.; Emerson,

C.; Lynch, H. F.; Jamrozik, E.; Jecker, N. S.; Kamuya, D.; Kapulu, M.;

Kimmelman, J.; MacKay, D.; Memoli, M. J.; Murphy, S. C.; Palacios,

R.; Richie, T. L.; Roestenberg, M.; Saxena, A.; Saylor, K.; Selgelid, M.

J.; et al. Ethics of Controlled Human Infection to Address COVID-19.

Science 2020, 368, 832−834.


ACS Nano www.acsnano.org Review


https://dx.doi.org/10.1021/acsnano.0c06109

ACS Nano 2020, 14, 12370−12389


12389

No comments:

Post a Comment

Note: only a member of this blog may post a comment.