ArticlePDF Available

Abstract

mRNA is the central molecule of all forms of life. It is generally accepted that current life on Earth descended from an RNA world. mRNA, after its first therapeutic description in 1992, has recently come into increased focus as a method to deliver genetic information. The recent solution to the two main difficulties in using mRNA as a therapeutic, immune stimulation and potency, has provided the basis for a wide range of applications. While mRNA-based cancer immunotherapies have been in clinical trials for a few years, novel approaches; including, in vivo delivery of mRNA to replace or supplement proteins, mRNA-based generation of pluripotent stem cells, or genome engineering using mRNA-encoded meganucleases are beginning to be realized. This review presents the current state of mRNA drug technologies and potential applications, as well as discussing the challenges and prospects in mRNA development and drug discovery.
mRNA transcript therapy
Expert Rev. Vaccines Early online, 1–17 (2014)
Drew Weissman
Department of Medicine, University of
Pennsylvania, Philadelphia, PA 19104,
USA
Tel.: +1 215 573 8491
Dreww@upenn.edu
mRNA is the central molecule of all forms of life. It is generally accepted that current life on
Earth descended from an RNA world. mRNA, after its first therapeutic description in 1992, has
recently come into increased focus as a method to deliver genetic information. The recent
solution to the two main difficulties in using mRNA as a therapeutic, immune stimulation and
potency, has provided the basis for a wide range of applications. While mRNA-based cancer
immunotherapies have been in clinical trials for a few years, novel approaches; including, in vivo
delivery of mRNA to replace or supplement proteins, mRNA-based generation of pluripotent
stem cells, or genome engineering using mRNA-encoded meganucleases are beginning to be
realized. This review presents the current state of mRNA drug technologies and potential
applications, as well as discussing the challenges and prospects in mRNA development and
drug discovery.
KEYWORDS:modified nucleoside • mRNA • PKR • pseudouridine • purification • therapeutic protein • vaccine
The first use of mRNA encoding a potentially
therapeutic protein delivered in vivo occurred in
1990 when Wolff et al. demonstrated the
expression of reporter proteins after direct injec-
tion of mRNA and occurred with the first
description of the in vivo delivery of plasmid
DNA [1]. Interestingly, mRNAs were used ther-
apeutically in only a single follow-up with injec-
tion into the brain and amelioration of disease
symptoms [2]. In the mean time, the field
focused on plasmid and DNA based technolo-
gies and, then, viral vector delivery systems.
In considering the advantages and disadvan-
tages of mRNA-based therapy, there are several
conceptual advantages compared to other
nucleic acid-based approaches. Unlike DNA-
based therapy, mRNA does not have the risk
of integration into the chromosomes [37],
which can lead to insertional mutagenesis with
potentially disastrous results [8,9]. mRNA deliv-
ered therapeutically only results in transient
translation that can be controlled by both
changes in the UTRs or coding sequence and
is completely degraded through physiologic
pathways [1013]. This is considered both an
advantage and a disadvantage depending on
therapeutic needs. In principle, mRNA-based
therapies appear to be much safer than DNA
or viral and are applicable to a broad spectrum
of disorders both acute and chronic.
The use of mRNA in the field of cancer vac-
cination has experienced the greatest amount of
preclinical investigation and has achieved
multiple stages of clinical testing. With the
advent of new discoveries that have both
increased the amount of protein produced per
delivered mRNA, through improvements in
mRNA structure and delivery, and the reduc-
tion of intrinsic immunogenicity of mRNA
[1424], new approaches to replace proteins in
cardiology, oncology, endocrinology and the
treatment of genetic disorders, such as cystic
fibrosis or hemophilia [25,26], are underway.
These therapeutic approaches are still very early
in development and require both basic science
and pharmaceutical development of nearly every
step from synthesis to delivery of mRNA.
In this review, we attempt to offer a summary
on the current state of therapeutic mRNA tech-
nology, including vaccines against epidemic and
pandemic infections, protein replacement for
specific disease therapy and genetic deficiency,
and gene editing. The strengths and the upcom-
ing challenges that will likely impact on the
progress of this new class of drug will also
be highlighted.
mRNA pharmacology
At the present time, therapeutic mRNA is made
through the process of in vitro transcription,
where a linearized plasmid, PCR product or
synthesized DNA contains a promoter that is
recognized by a phage polymerase, such as T7,
T3 or SP6. In an in vitro reaction, multiple cop-
ies of mRNA are made from the DNA. The
DNA to be transcribed, in addition to
informahealthcare.com 10.1586/14760584.2015.973859 2014 Informa UK Ltd ISSN 1476-0584 1
SPECIAL FOCUS yRNA Vaccines Review
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
containing the coding sequence of the protein of interest, can
contain optimized and specialized 5and 3UTRs and extended
lengths of A:T to produce a defined mRNA. In addition to the
cognate nucleotides, modified nucleoside triphosphates and vari-
ous versions of 5caps can be included in the transcription reac-
tion. The DNA template is digested with DNases to terminate
transcription, and the mRNA is purified by conventional meth-
ods of nucleic acid isolation. The mRNA is engineered to resem-
ble fully processed mature mRNA molecules, as they would be
found in the cytoplasm of eukaryotic cells. Additional posttran-
scriptional modifications can be added, such as the addition of a
5cap and 2-O-methylation of the cap or extension of the poly
(A) tail. The resulting product should optimally contain a 5cap,
a coding sequence or ORF that encodes the protein of interest
and starts and stops with the appropriate codons, which can be
flanked by UTRs that can contain additional instructions for
translating the coding region.
While DNA and RNA are both protein-encoding molecules
that utilize cellular systems to produce the protein of interest,
there are many differences in the use of mRNA versus DNA.
The first involves the site of initial activity, which is the
nucleus for plasmid DNA and the cytoplasm for mRNA. This
represents a major difference between DNA and RNA, as
DNA requires nuclear envelope breakdown during cell division
to reach the nucleus and produce encoded protein, while
mRNA only needs to reach the cytoplasm for translation.
For mRNA to be translated into protein, it must survive in
the extracellular space that contains high levels of ubiquitous
RNases; it must reach or be targeted to the cells of interest for
translation, and finally, it must cross the cell membrane. The
cell membrane hinders passive diffusion of large negatively
charged mRNA molecules. Although, it has been demonstrated
that eukaryotic cells can actively engulf naked mRNA through
a receptor mediated mechanism, in most cell types, the rate of
uptake with transfer to the cytoplasm is extremely low [27,28],
and naked mRNA has a very short half-life in tissues and fluids
containing high levels of RNase activity. The delivery of
mRNA to cells both in vitro and in vivo can be greatly
increased by the use of complexing agents that protect the
mRNA from degradation by RNases, increase cellular uptake
and can perform targeting to specific tissues and cell types.
The route for complexed and naked nucleic acid uptake by
cells is via endosomes [29]. The mRNA must then escape from
the endosome and reach a translationally competent region of the
cytoplasm. Once in the cytoplasm, the pharmacology of mRNA
is subject to the same complex cellular mechanisms that regulate
the stability and translation of native mRNAs [30,31], with the pos-
sible exception that the bound proteins will at least initially differ
between mRNAs that have been processed and exited the nucleus
versus being in vitro transcribed. The roles that these proteins
play in cytoplasmic regulation of translation and mRNA degra-
dation are unknown. The encoded protein that is translated from
the in vitro transcribed mRNA undergoes posttranslational modi-
fication as directed by both the host that the mRNA is injected
into and the cell type that is translating the mRNA. The half-
lives of both the mRNA and the protein are critical factors for
the pharmacokinetics of delivered mRNA therapeutics and both
can be easily modulated to increase or decrease the duration of
protein activity.
Therapeutically, mRNA may be directly conveyed into a
patients cells, ex vivo, which are then administered back to the
patient. This allows a more precise control of elements of mRNA
pharmacology that have not been solved for in vivo delivery,
including specific targeting of a cell type for protein expression
and control over the amount of expression. Alternatively, the
future of mRNA therapeutics will rely upon mRNA used in a
suitable formulation with direct in vivo administration. Advances
in cell or tissue targeting are needed to make this possible.
Factors modulating the translation & stability of mRNA
Significant advances have been made in modifying the 5cap, 5-
and 3-UTRs, the coding sequence and the poly(A) tail to
increase the translational efficiency and cytoplasmic stability of
mRNA. The optimization of translation can take on many forms
depending on the therapeutic needs and can include very rapid
initial production of the protein and sustained duration of trans-
lation or high peak translation followed by rapid reduction of
protein production. At present, protein production can be modu-
lated from minutes to nearly a week [23,31,32]. The field of modu-
lating mRNA pharmacology will advance as more therapeutic
mRNAs are described, but, at present, there is much space for
advances and deeper understanding. Further research into
mRNA binding factors and their respective mRNA binding sites
will unlock continued chances to engineer mRNA for diverse
pharmacokinetic activities.
In vivo mRNA delivery
In vitro transcribed mRNA used for immunization has been in
numerous clinical trials. Clinical trials using uncomplexed
(naked) or protamine-complexed mRNA in vaccines with deliv-
ery either by the intradermal route [3337] or with direct injec-
tion into lymph nodes [3840] are currently ongoing. The
delivery of naked mRNA seems to be satisfactory to produce
potent immune responses; it is likely to not be sufficient for
other clinical applications where higher amounts of protein will
be required and the delivery to a variety of cell types is needed.
The in vivo delivery of therapeutic mRNA for nonimmuniza-
tion purposes is currently the greatest hurdle and challenge.
Depending on the therapeutic goal of the delivered mRNA,
delivery can take on many different objectives; including, target-
ing specific cell types; delivery of high level protein production to
as many cells as possible without regard to cell type; delivery to a
majority of a particular cell type, such as bone marrow stem cells,
with or without regard to delivery to other cell types; and delivery
to specific organs, such as liver, brain, heart, and so on.
Therapeutic mRNAs can be divided between extracellular
and intracellular site of action of delivered encoded proteins.
The pharmacokinetics of each are very different. Systemic cir-
culation of an effector protein, such as observed with many
hormones and cytokines, is much easier to achieve, especially if
Review Weissman
doi: 10.1586/14760584.2015.973859 Expert Rev. Vaccines
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
any producer cell is qualified to make the appropriate func-
tional protein. The targeting of a specific cell type in vivo is
much more difficult and approaches that attempt this are still
early in development. This is further compounded if the major-
ity or that entire cell type needs to receive the mRNA for a
complete therapeutic effect.
Basic science studies in both mice and primates have demon-
strated that delivery of mRNA by the intraperotineal, subcuta-
neous and intramuscular routes achieves elevated plasma levels
for protein hormones, including erythropoietin (EPO) ([41] and
[DREW WEISSMAN,UNPUBLISHED DATA]). Lipid nanoparticles (LNPs), origi-
nally called stable nucleic acid lipid particles [42,43], have been
used in clinical trials for the delivery of siRNA to the liver
(reviewed in [44]). Liver cells are reached by many types of lipo-
somal and polymer delivery systems, and the liver serves as the
production site for many secreted systemic proteins; thus, liver
targeting could be an efficient method for producing high
amounts of recombinant proteins with mRNA delivery [45].
There are many other in vivo delivery systems using a variety
of approaches including lipids, polymers, sugars and den-
drimers being developed for nucleic acids including mRNA.
The results of these studies will direct future approaches to use
mRNA therapeutically.
The delivery of intracellular actingproteins to correct defec-
tive or deficient proteins with mRNA is a greater challenge,
especially when the fraction of cells that need to be restored is
high, in order to have a clinical result. The first determinations
for such approaches needs to consider both the percent of tar-
get cells that have to produce the encoded protein and the
amount of protein that needs to be present per cell for ade-
quate function/activity. Unlike in vitro transfection of cells that
can typically achieve over 80% and often close to 95% effi-
ciency, except for certain cell types that are highly resistant to
most chemical transfection methods, such as lymphocytes,
in vivo delivery is highly variable and poorly studied. At pres-
ent, the most work has been performed with LNPs and the
delivery of siRNA, including human clinical trials. It has been
reported that with such a delivery, over 80% of all cell types in
the liver can have siRNA within them [46]. For most delivery
systems with mRNA, especially when used as a vaccine, the
percent and type of cell that takes up the mRNA and produces
the protein is rarely identified and the measure for efficiency is
the resultant immune response.
5‘ cap
Eukaryotic mRNAs have a 7-methylguanosine (m7G) cap at
the 5end of the mRNA that is appended during the transcrip-
tion process through a 5´-5´-triphosphate (ppp) bridge
(m7GpppNpNp…). The 5cap is critical for efficient transla-
tion through its binding to eIF4E (reviewed in [47]). Three
decapping enzymes, Dcp1, Dcp2, and DcpS, also bind the 5
cap and are involved in the regulation of mRNA decay [48].
There are 2 main approaches to adding a 5cap to in vitro syn-
thesized mRNA. The first uses recombinant Vaccinia virus-
derived enzymes, one of which adds the m
7
GpppN cap, and
the second adds a 2-O-methyl group to the penultimate nucle-
otide (m
7
Gppp 2-O-methyl-NpNp…) [49,50]. The Vaccinia
virus enzymes make a 5cap-1 that is identical to the cap struc-
ture that is most frequently found in eukaryotic mRNAs.
A second more commonly used approach is to include a syn-
thetic cap analog in the in vitro transcription reaction. There
are advantages and disadvantages to both approaches, adding
the cap after transcription requires a second reaction, while
including a cap analog during transcription results in a decrease
in the total amount of mRNA made and a fraction of the
mRNA produced will not contain a cap analog. This is because
the cap analog competes with GTP in the transcription reac-
tion and a balance must be struck between the total amount of
mRNA produced and the percent that is capped by varying the
ratio of GTP to cap analog.
Current cap analogs comprise 3 basic families: m7GpppG cap
analog; antireverse cap analogs (ARCA), 3´-O-Me-m7GpppG
and modified antireverse cap analogs. Initial mRNA research was
done using mRNAs produced with m7 cap analog (GpppG) [1,51]
and continues to be used in most mRNA clinical trials. Unfortu-
nately, a significant amount of m7GpppG cap analog used dur-
ing in vitro transcription is incorporated in an opposite
orientation [52] resulting in inefficient recognition by translational
machinery. To combat this, ARCAs were developed [53,54].
mRNAs constructed with ARCA-caps exhibit superior transla-
tional efficiency in a variety of cell types compared to standard
cap analog [5557]. A new class of ARCA caps has been established
that contains a phosphorothioate bond [58]. This results in resis-
tance to decapping by the Dcp2 enzyme leading to an extension
of the half-life of the mRNA [59]. Phosphorothioate containing
cap analog mRNA encoding an antigen when given to mice
resulted in a potent antigen-specific response that was greater
than that observed for mRNA with standard ARCA-mRNA [24].
Untranslated regions
Eukaryotic cells use the UTRs in mRNAs to regulate transla-
tion; multiple different approaches are used, including
upstream ORFs, sequences that promote degradation, modula-
tory protein binding sequences and micro RNA sites [6063].
The description of these mechanisms is beyond the scope of
this review. For the purpose of therapeutic mRNA, most appli-
cations desire UTRs that lead to the highest level of translation,
although some applications may seek very fast translation or
short durations of protein translation.
Many mRNAs in clinical trials and in use for basic science
research utilize the 3UTRs derived from a- and b-globin
mRNAs. Multiple sequence elements have been observed to
increase the translation and stability of mRNAs that harbor
them [6466]. Studies have also identified multiple cellular and
viral 5and 3UTRs that enhance the translational efficiency
and stability of mRNAs they are appended to. 5UTRs found
in numerous orthopoxvirus mRNAs have been demonstrated to
inhibit both cap removal and 3to 5exonuclease degradation
(reviewed in [6]). Depending on the therapeutic need, mRNA
with limited lengths of protein production may be needed.
mRNA transcript therapy Review
informahealthcare.com doi: 10.1586/14760584.2015.973859
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
This can be achieved by incorporating sequence elements, such
as AU-rich sequences, into 3UTRs, which results in rapid
mRNA degradation and associated short interval of protein
expression [67]. New approaches to develop and optimize UTR
function using techniques such as Systematic Evolution of
Ligands by Exponential enrichment [68,69] are currently being
used. The results of these studies will be important in defining
methods to improve UTR function and in the creation of
improved mRNA therapeutics.
Coding sequence
The first element of the mRNA coding sequence is the signal
sequence/peptide, which is a short (5–30 amino acids) peptide
found at the N-terminus of most newly synthesized proteins
that will enter the secretory pathway. In the construction of
mRNA encoding therapeutic proteins, the signal sequence can
be the natural version found in the cellular mRNA or it can be
engineered, which is typically used when the fate of the desired
proteins needs to be redirected, such as membrane bound to
secreted. In the case of mRNA encoding vaccine immunogens,
the desired processing of the immunogen by antigen presenting
cells helps to determine the type of signal sequence, in order to
obtain the correct immune response. As part of proteasome
processing, peptides that can be loaded onto MHC class I mol-
ecules are generated that can then be presented to CD8
+
T cells. The loading of MHC class II molecules to generate a
CD4
+
T cell response to a vaccine immunogen can be pro-
moted by modifying the mRNA with an endosome targeting
sequence [7072].
Codon optimization serves several purposes. It replaces rarely
used codons with more abundant and frequently used codons
based on tRNA preferences of the host, but it also removes
RNA secondary structures, potential splice motifs, and other
sequences that may interfere with efficient transcription, nuclear
export and translation of an mRNA. The role of codon optimi-
zation to avoid rare and seldom used tRNAs in mRNA-based
therapies has been minimally studied. Mammalian codons usu-
ally have G or C in their third degenerative position, and such
sequences are expressed more efficiently than those in which
the codons end with A or T [73]. Typically, codon optimiza-
tions are performed by comparing protein production from
DNA plasmids carrying the wild-type sequence to the opti-
mized sequence [73]. In this setting, it is difficult to sort out if
the increased protein expression from the codon-optimized
construct is caused by enhanced transcription, increased RNA
stability, or augmented translation. However, studies that have
directly compared codon optimization that optimizes tRNA use
and removes RNA secondary structure, potential splice motifs
and other wild-type sequences in mRNA have found small
(~1.6 to threefold) increases in the protein produced [74,75].
CureVac GMBH, the first company to perform Phase II clini-
cal trials of mRNA-based vaccination for cancer, has identified
and patented the replacement of all third position Us with G
or C. Preliminary studies in our lab demonstrated an approxi-
mately fivefold increase in protein produced. This approach to
codon optimization is mechanistically different from other
studies [74,75] and only seeks to reduce third position Us and
replace them with Gs or Cs, which likely accounts for the dif-
ference in enhancement to translation.
While the increase in protein production by certain types of
codon-optimization is significant, there are potential difficulties
that could be encountered. Certain proteins require slow trans-
lation for their proper folding and modification and the pres-
ence of rare codons allows this to happen [76]. Responses to
hidden and unexpected epitopes have been identified in a num-
ber of cancers and in HIV infection [SHAW,GEORGE M, PERS.
COMM.] [7782]. Some of these occur due to translation in different
frames caused by ribosomal frame shifting or alternate start site
utilization. Codon optimization of vaccine immunogens would
not contain these epitopes.
Poly(A) tail
The poly(A)-binding protein bound to the poly(A) tail interacts
with the eIF4F complex bound to the cap as part of the transla-
tion initiation complex. The poly(A) tail regulates the transla-
tional efficiency and stability of mRNA in synergy with the 5
cap and with various other determinants [83].In vitro synthesized
mRNA can either have an encoded poly(A) tail in the template
vector or as a separate reaction after transcription where the poly
(A) tail is added by recombinant poly(A) polymerase. A downside
that precludes clinical use of enzymatic polyadenylation is that
each preparation of RNA contains a mixture of RNA species
with different lengths of poly(A) tail, which is unacceptable to
the US FDA. Typical lengths of encoded poly(A) tail range
around 120–150 nucleotides [23,56,84].
mRNA immunogenicity
There are more innate immune receptors that recognize RNA
than any other foreign or self-molecule, indicating the impor-
tance of RNA in the immune system. RNA, including in vitro
transcribed, induces innate immune responses through the activa-
tion of pattern recognition receptors, whose function is to iden-
tify and respond to pathogenic RNAs (FIGURE 1). Innate immune
recognition of RNA has been extensively reviewed [8593].
RNA sensor activation can have both direct and indirect effects
on translation. In vitro transcribed RNA activates protein kinase
R and 2-5-oligoadenylate synthetase that lead to a direct inhibi-
tion of protein synthesis [14,20,21]. Activation of other cytoplasmic
and endosomal RNA sensors, which include TLR3, TLR7,
TLR8, RIG-I, MDA5, NOD2, IFIT-2, DDX60, DHX9,
DDX3, the DDX1-DDX21-DHX36 complex, HMGB proteins,
and LRRFIP1 (reviewed in [94]) can result in the release of Type I
interferon that activate interferon-inducible genes including pro-
tein kinase R, RNase L, 2-5-oligoadenylate synthetase and
others that when activated, directly inhibit translation (FIGURE 1).
The negative effects of RNA sensor activation has been
highlighted for mRNA vaccination where lipid-complexed
mRNA encoding HIV gag induced Type I interferon that led to
an inhibition in translation of the mRNA and a reduced immune
response [95].
Review Weissman
doi: 10.1586/14760584.2015.973859 Expert Rev. Vaccines
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
The intrinsic immunogenicity of in vitro transcribed mRNA
is viewed as advantageous in vaccine therapeutics due to its
adjuvant activity [9698] that results in potent antigen-specific
humoral and cellular immune responses [99,100]. RNA has a dis-
tinctive pattern of immune stimulation, but this can be par-
tially modulated by altering the characteristics (type and size)
of the particle used to deliver it [22,101,102].
mRNA was first used therapeutically with in vivo administra-
tion to treat a disease in 1992 [2], but was not used again outside
of vaccination until the source of its immunogenicity and meth-
ods to overcome it were discovered. The observation that natu-
rally occurring RNAs differed in their immune activating
potential and the level of immune activation correlated with the
amount of modified nucleosides contained in the RNA led to the
finding that naturally occurring modified nucleosides such as
pseudouridine, 5-methyluridine, 2-thiouridine, 5-methylcytidine
(m5C), and N6-methyladenosine suppressed RNAs immune-
stimulatory effect by avoiding activation of TLR3, TLR7 and
TLR8 [103]. This was followed by the identification of other RNA
sensors, including RIG-I [18,19], protein kinase R [20,21],2-5-oli-
goadenylate synthetase and RNase L [14], similarly not being
activated by RNA containing nucleoside modifications. Double-
stranded RNA contaminants are formed during phage polymer-
ase transcription [104107], and the presence of pseudouridine
and/or m5C did not obviate their immunogenicity. The dsRNA
contaminants were eliminated using an HPLC purification
method that resulted in mRNA that did not have any immune-
stimulatory activity [22]. The absence of the activation of RNA
sensors led to mRNA that was translated at much greater levels
(> 1000-fold) in vivo without inducing proinflammatory
cytokines, Type I interferons or adverse events [22,41] and led to
the current resurgence of nonvaccine mRNA therapeutics.
Preclinical & clinical applications
The main major therapeutic areas that mRNA or nucleoside
modified mRNA are currently being explored are the fields of
regenerative medicine, immunotherapy, vaccines and protein
replacement (FIGURE 2).
Vaccines
The first preclinical application of mRNA was used to induce a
protective immune response and employed direct injection of
carcinoembryonic antigen encoding mRNA [108]. Direct addi-
tion of mRNA derived from tumor cells or encoding tumor
specific antigens to dendritic cells (DCs) ex vivo followed by
administration of DCs back to the host, followed this first
demonstration of efficacy and led to extensive development
[109112] leading to clinical trials (reviewed in [5,113]). Multiple
enhancements to this approach to therapy were attempted,
including the addition of mRNAs encoding costimulators
and cytokines [55,114117]. Some spectacular results have been
reported, including the use of dendritic cells electroporated
with melanoma-associated antigen fused to a HLA-class II
targeting signal (DC-LAMP), adjuvanted with mRNAs
encoding CD40 ligand, a constitutively active Toll-like
receptor 4, and CD70 (TriMix [118]), in treated melanoma
patients where antitumor activity with durable disease con-
trol was observed [119]. Tumor-derived mRNA loaded DCs
have entered a Phase III clinical trial for patients with
advanced renal cell cancer [120].
Signals to IVT RNA
????????XXXX XXX
RNase LOASPKR
Murine
IFIT-2
LRRFIP1HMGB1, 3DDX3
DHX9
+
IPS1
DDX60NOD2MDA5RIG-ITLR8TLR7TLR3
TRIF MYD88
MAVS
Type I IFNs
Inhibits viruses
unable to 2-O-
methylate caps
Inhibits protein
translation
2-5 linked
oligoadenylates
- Rnase L
elF2α
β-catenin
-IRF3
Promote signaling
through TLRs
and RIG-I
Proinflammatory
cytokines and
type I IFNs
DDX1-
DDX21-
DHX36
complex
X
Figure 1. Innate immune receptors demonstrated to respond to RNA. In vitro transcribed mRNA and other forms of
single-stranded and double-stranded RNA are recognized by numerous innate immune receptors. Whether signaling is initiated by in vitro
transcribed RNA is indicated. Signaling occurs through different pathways, as indicated, and results in inflammation associated with Type
I interferon, activation of cascades of transcriptional programs, proinflammatory cytokines and chemokines or inhibition of
translational machinery.
Indicates DREW WEISSMAN [UNPUBLISHED DATA].
mRNA transcript therapy Review
informahealthcare.com doi: 10.1586/14760584.2015.973859
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
Direct injection of mRNA encoding an antigen has many
advantages over ex vivo manipulations, which led to multiple
investigations using direct injection of mRNA with different
strategies, including a variety of complexing agents and gene
gun delivery (reviewed in [7]). CureVac (Tu
¨bingen, Germany)
took the lead early on to develop direct intradermal injection
of mRNA with and without protamine complexing [33,34,101].
They have completed multiple clinical trials including ones
directed at prostate and non-small cell lung cancer [121] and
continue to develop this approach for other types of cancer.
BioNTech (Mainz, Germany) took an alternative approach
that has entered a clinical trial in melanoma patients [40]. They
used direct injection of uncomplexed mRNA into lymph
nodes. The mRNAs were optimized for efficient and high level
translation. Their studies demonstrated selective uptake by local
DCs with RNA induced DC maturation and effective presenta-
tion to T cells [38,122124]. They have also begun to personalize
mRNA antigen vaccination therapy by massive parallel
sequencing of tumor cells with identification of tumor antigens
and construction of multi-epitope polypeptides containing each
patients tumor antigens and identified mutations [125127].
Testing this approach has been initiated. Recently, the Univer-
sity Hospital Brussels in collaboration with the Vrije University
Brussels have started a clinical trial where hepatocellular
Protein replacement
Direct injection
Ex vivo treatment
Transposases
CRISPR/Cas9
TALE nuclease
Zinc-finger nuclease
Preclinical
Direct injection
Ex vivo treatment
Clinical
Direct injection
Ex vivo treatment
T cells: CAR-mesothelin [132]
Preclinical
Ex vivo treatment
HIV [79,133,134]
Direct injection
RSV [135,136], influenza [135,137,138], TB-associated Hsp65 [139], SV40 large T [140]
Clinical
Ex vivo treatment
iPS cell generating [146-151]
Injection of mRNA encoding allergens [144,145]
Allergy
Reprogramming of cells
DCs: HIV antigens [141-143]
Infectious disease vaccines
DCs: pancreatic adenocarcinoma [122,123], prostate cancer [124,125], melanoma [126-130], colon cancer [131]
renal cell carcinoma [12], melanoma [14]
T cells: CAR-Her-2/neo [117], CAR-mesothelin [118,119], CAR-CD19 [120,121]
DCs: MUC1[114], iLRP [115], survivin[116]
CEA [110], NY-ESO [111], Trp2, tyrosinase [112], gp100 [113]
Cancer Immunotherapy
knock out in rats [108], gene editing in zebrafish [109]
transgenic rat [104], mutated mice [105,106], zebrafish [43,107] editing
gene editing in mice [102], zebrafish [103]
gene editing in Drosophila [97], mouse embryo [98,99], HT1080 cells, mouse liver in vivo [100], HeLa cells [101]
Genome editing
MSCs with Slex, P-selectin glycoprotein ligand-1, IL-10 [95], DCs with IL-4 [96]
VEGF-A [90], BAX [91], Vasopressin [2], EPO [15,92], HSV1-TK, [93]SP-B [92], Foxp3 [94]
Figure 2. Fields of therapeutic applications of mRNA. Clinical and preclinical development of mRNA therapeutics, whether delivered
by ex vivo treatment of cells or direct injection on mRNA is described.
OAS: 2-5-oligoadenylate synthetase.
Review Weissman
doi: 10.1586/14760584.2015.973859 Expert Rev. Vaccines
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
carcinoma patients are treated with intranodal injections of mRNA
encoding tumor antigens and TriMix adjuvant (EUDRACT
2012-005572-34).
Infectious diseases vaccines
Curevac has expanded their 2-component mRNA vaccine, an
RNA adjuvant and a naked mRNA encoding antigen, to viral
influenza. It was demonstrated to induce strain-specific protective
immune responses in aged and newborn mice, along with long-
lasting protection in pigs and ferrets [128]. Alpha Viruses replicate
their genome by self-amplifying the RNA. An encoded RNA-
dependent RNA polymerase makes reverse (antisense) copies of
the entire genome and then full-length genomic RNA plus sub-
genomic structural proteins. These structural proteins can be
replaced with antigen to form a self-replicating RNA vaccine.
This approach has been used for flaviviruses [129], parainfluenza,
respiratory syncytial virus and influenza viruses [130132]. Novartis
recently demonstrated that intramuscular injection of an RNA
self-replicating vaccine complexed in LNPs at very low doses
induced protective immune responses against influenza virus in
mice and non-human primates [133,134].
Two Phase I/II clinical trials using mRNA coding for HIV
antigens have been performed. These studies delivered DCs
ex vivo transfected with antigen-encoding mRNA to patients
on antiretroviral therapy. Immunization was deemed safe and
CD4
+
and CD8
+
T cell responses against the antigens delivered
were induced, but no antiviral effects were observed [135,136].
Vaccines for IgE hypersensitivity
Environmental allergic diseases are a hypersensitivity disorder of
the immune system mediated by IgE antibodies. Current treat-
ments involve immunization with graded dosing of the allergen to
modulate the type of T cell response and induce IgG antibodies
that compete and inhibit IgE binding to allergens. New and effec-
tive, but very costly, alternatives involve the injection of anti-IgE
monoclonal antibody [137140]. The molecular identification of the
most common antigens leading to hypersensitivity has allowed
the development of recombinant vaccine methods. mRNA encod-
ing allergen vaccination prompted long-lasting allergen specific
Th1 immune responses that protected mice from allergen
exposure-mediated inflammation of the lung [141145].
Cancer immunotherapy
Chimeric antigen receptor therapy has shown very good results
in several diseases (reviewed in [146]). Cells transfected with and
expressing such receptors are able to kill tumor cells that
express the targeted antigen. Multiple methods can be used to
load expanded T or NK cells, including mRNA. The advantage
of mRNA delivery is its transient nature, which can reduce the
risk of toxicity by uncontrolled expansion of the adoptively
transferred cells. A variety of antigen-specific receptors have
been delivered using mRNA and tested for antitumor immu-
nity in animal models [147150]. Expanded T cells loaded with
mRNA encoding chimeric antigen receptors is being used in a
Phase I clinical trial for pancreatic cancer [151].
mRNA encoding engineered nucleases for gene editing
Early forms of gene therapy for the repair of inborn errors of a
particular gene sought to replace the defective gene by delivering
a functional copy containing its own promoter and regulatory
regions and inserting it into the chromosomes using a viral vec-
tor. Genome editing has emerged as a potential alternative for
gene therapy. Zinc finger nucleases and transcription activator-
like effector nucleases use meganucleases linked to protein
sequences that bind to specific DNA sequences that allows site-
specific cutting of DNA in chromatin [152,153]. The CRISPR/
cas9 system is derived from the acquired immune system of cer-
tain bacteria that uses RNA tags linked to a protein with nucle-
ase activity called cas9. The RNA tags identify the site for
cutting [154,155]. The major adverse event encountered in all
forms of gene editing is the risk of nonspecific editing. The
amount of off site effects associates with increasing duration of
functional enzyme as mediated by plasmid or viral delivery sys-
tems [156]. All three of the gene editing technologies only require
the nucleases to be present for a short duration, thus, their tran-
sient expression from encoding mRNA would meet this criteria
and likely minimize the potential for nonspecific effects.
mRNAs encoding Cas9, transcription activator-like effector
nucleases and zinc finger nucleases and ZFNa have been success-
fully used to edit genomes ex vivo in embryonic cells from dif-
ferent species and in vivo in rodents and zebrafish [157164].
In summary, for all three gene editing approaches, the use of
mRNA either by direct injection in vivo or ex vivo treatment
would allow the fine tuning of dosing that cannot be achieved
with plasmid and viral delivery. The additional advantages of
high transfection efficiency without cell toxicity would be bene-
ficial for the generation of transgenic animals and the treatment
of human genetic diseases with potential application to other
types of diseases, including cancer.
Cell fate reprogramming with mRNA
In 2010, mRNAs coding for the Yamanaka stem cell factors
(Oct3/4, Sox2, Klf4, c-Myc) [165] containing pseudouridine and
5-methylcytydine were used to efficiently reprogram cells to
pluripotency (iPS cells) without any integration events [166].
A number of variants using the nucleoside modified mRNA
approach have been described that claim a more effective
induction of pluripotent stem cells or cell fate conversion
(reviewed in [3]). Prior to the use of nucleoside modified
mRNA with its lack of innate immune signaling, mRNA was
already being used to induce iPS cells [167], because of its high
in vitro transfection efficiency and transient expression with
lack of genomic integration. The transient expression of iPS
factors makes the use of nucleoside-modified mRNA for creat-
ing iPS cells attractive for different fields, including disease
modeling [168] and therapy for a variety of diseases [169] with
potential application to treatment.
Protein replacement therapies
mRNA based therapeutics deliver a natural product, that is, all
of its components are found in an organism and depending on
mRNA transcript therapy Review
informahealthcare.com doi: 10.1586/14760584.2015.973859
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
how it is made, is indistinguishable from other RNA in the
cell. It allows robust and tunable delivery of the encoded pro-
tein. It avoids the expensive manufacturing of therapeutic pro-
teins in fermentation tanks and protein-specific purification
procedures. All of these attributes allow mRNA the potential to
reduce cost and time for entry into clinical testing, which is
particularly helpful for epidemic vaccines and therapies.
mRNA therapeutics can be viewed as a form of transient
gene therapy without the potential complications of long-term
gene therapy, including insertional mutagenesis, vector immu-
nity and effects of viral replication on cell function. The deliv-
ery of therapeutic proteins by mRNA is an obvious therapeutic
objective. The delivery of such encoded proteins can be initially
divided between extracellular acting and systemic proteins ver-
sus intracellular acting proteins and then further divided based
on; short-term delivery to treat deficient or non-functional pro-
teins; long-term replacement of deficient or non-functional pro-
teins; delivery of exogenous therapeutic proteins (monoclonal
antibodies), and acute site-specific or systemic delivery of a pro-
tein during a medical emergency or therapeutic procedure.
As described earlier, while the first delivery of an mRNA
in vivo used reporter proteins [1], it was quickly followed by
the delivery of vasopressin to the hypothalamus to reverse dia-
betes insipidus [2]. Therapeutic mRNA research outside of vac-
cines became dormant after this description until the
identification that the inclusion of naturally occurring modified
nucleosides in mRNA along with purification procedures
resulted in an mRNA that did not activate any RNA sensors
and was translated in primary cells and in vivo at much greater
levels [15,22,41,170]. These findings led to a reemergence into the
study of mRNA based protein therapy for the treatment of a
wide range of diseases.
The most advanced studies delivered EPO to mice and mac-
aques and used pseudouridine-modified, HPLC-purified
mRNA. As little as 10 ng of mRNA resulted in a physiologic
response, increase in reticulocytes, in mice [41].In vitro tran-
scribed mRNA coding for surfactant protein B and containing
25% of the modified nucleosides 2-thiouridine and m5C were
analyzed in a mouse model of congenital surfactant protein B
deficiency. Surfactant protein B mRNA 20 mg was delivered
two-times a week and was found to protect mice from respira-
tory failure and prolong their lifespan [16]. The intratracheal
delivery of a nucleoside-modified mRNA containing 10% of
the modified nucleosides 2-thiouridine and m5C and coding
for the regulatory T cell transcription factor Foxp3 (20 mg
before of during challenge) protected the mice from allergen-
induced tissue inflammation and airway hyperresponsive-
ness [171]. In a new therapeutic approach that will require much
additional development before it is ready for clinical use, direct
intramyocardial injection of mRNA containing pseudouridine-
and m5C, encoding VEGF-A improved heart function and
enhanced long-term survival in a model of mouse myocardial
infarction [172]. In another report, mesenchymal stem cells were
engineered ex vivo to express functional rolling machinery
(P-selectin glycoprotein ligand-1 and Sialyl-Lewisx) and the
immunosuppressive cytokine IL-10 using pseudouridine-
modified mRNA. When delivered in vivo, the engineered cells
rapidly localized to an inflammatory site in the ear and deliv-
ered IL-10 that resulted in a reduction in inflammation [173].
Extensive studies have so far identified similar levels of transla-
tion after HPLC purification [22,174] of mRNA containing pseu-
douridine and pseudouridine plus m5C, whereas mRNA
containing 1-methyl-pseudouridine consistently gave increased
levels of translation [DREW WEISSMAN,UNPUBLISHED DATA], suggesting the
superiority of this nucleoside modification.
There are multiple factors that must be considered beyond
simply injecting mRNA and hoping for a physiological effect.
Protein signal sequences are optimized for both the protein it
is attached to and the cell type that secretes the protein [175,176].
Relative secretory signal strength differs and optimal signal pep-
tides can be substituted to increase the amount of protein
released per mRNA molecule [177]. Cell type specific differences
in posttranslational modification need to be considered, espe-
cially when the delivered mRNA will not be translated by the
cells that naturally produce the protein. Certain cell types do
not have the capability to properly glycosylate every protein,
especially when highly complex glycosylation is needed [178].
Other types of posttranslational modification include proteo-
lytic processing and cofactor-dependent folding and clearance
of misfolded versions. Certain proteins require chaperones to
be properly folded or cleared, if misfolded, and complete post-
translational processing (reviewed in [179]). Endoprotease proc-
essing is an important part of the maturation of many
polypeptides, including cytokines, growth factors, neuropepti-
des, receptors, hormones, enzymes and plasma proteins [180].
A second form of proteolytic processing uses convertases that
function in the Golgi apparatus and secretory granules [181].
Cells that take up mRNA encoding a protein of interest need
to have any required endoprotease or convertase in order to
process and successfully produce functional proteins.
Clinical development of therapeutic mRNA
The only clinical applications beyond Phase I trials where
mRNA has been injected into patients or patient cells were
transfected ex vivo followed by readministration are vaccine tri-
als. Other uses of therapeutic mRNA, including gene editing,
protein replacement, cell fate reprogramming and cancer
immunotherapy will require the determination of the pharma-
cokinetic characteristics of the drug and the performance of
dose finding, which are only indirectly determined for mRNA
vaccine studies. The principal difficulty in determining mRNA-
encoded protein pharmacokinetics is that what is delivered
(mRNA) is not the final active agent (encoded protein). This
compounds the difficulty in determining potential variability in
the amount of protein produced, which is further compounded
by the bioavailability, defined as the percent of delivered
mRNA that is translated in a cell. Such potential variability
would appear to make the delivery of a protein therapeutic that
needs consistent and exact levels of active protein, such as insu-
lin, more difficult for mRNA. The variation of the protein
Review Weissman
doi: 10.1586/14760584.2015.973859 Expert Rev. Vaccines
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
product effect after administration of the same dose to the
same patient and then for different individuals will need to be
investigated. The analysis of interaction with other drugs will
also need to consider the use of drugs that alter mRNA uptake,
such as chloroquine-like drugs, or alter translation or RNA
metabolism, such as cancer and autoimmunity therapeutics.
The FDA and the European equivalent (EMA) acted on
multiple Investigational New Drug applications for the use of
mRNA as a vaccine, but they have not reviewed any protocols
where therapeutic mRNA to replace a protein or edit the
genome will be delivered in vivo. This leaves open many possi-
bilities for what they will require in preclinical testing. One can
assume that since mRNA, unlike DNA and viral vectors, does
not contain promoters or antibiotic resistance genes, and can-
not integrate into the genome and disruption of genes does not
occur, at least for protein replacement, that testing for genome
integration, germline transmission, carcinogenicity and genotox-
icity should not be required.
Potential immune recognition of encoded proteins could
lead to adverse events, including anaphylaxis and other immu-
nologic reactions, as well as local and systemic infusion-related
reactions. Immune responses can not only neutralize the bio-
logical activity of the delivered therapeutic protein, it could
also bind and impair or neutralize the activity of the endoge-
nous protein counterpart [182,183]. The classic examples of this
occurred with EPO. Adeno-associated virus (AAV) delivery of
EPO to cynomolgus macaques led to severe anemia due to
neutralization of encoded and endogenous protein [184].
A change in the formulation of therapeutic EPO protein led to
pure red cell aplasia in treated humans by neutralization of
endogenous EPO [185187]. The observation that both gene ther-
apy and protein delivery of EPO led to autoimmunity raises
the potential risk for any mRNA-delivered protein. Recombi-
nant proteins require production in cell lines, yeast, plants or
bacteria that can result in altered posttranslational modification
leading to immune recognition; thus, the advantage of delivery
of mRNA encoding the gene is that the protein is made and
processed by the host. The mechanism behind AAV EPO-
induced antibodies remains unknown but serves as a warning
to any genetic delivery of a protein. Current data does not
demonstrate any potential for the induction of immunogenicity
against the mRNA-encoded protein.
A theoretical potential for the induction of antibodies against
the in vitro-transcribed mRNA or the worsening of diseases
where anti-RNA antibodies already exist, such as systemic lupus
erythematosus, needs to be considered. Evidence suggests that
anti-RNA autoantibodies may play a role in both the induction
and progression of autoimmunity in systemic lupus erythema-
tosus patients [188]. This raises the potential that mRNA ther-
apy using in vivo-administered mRNA could initiate or worsen
autoimmune diseases, including systemic lupus erythematosus.
Epitopes bound by autoantibodies have been identified and
potentially such epitopes could be engineered out of therapeutic
mRNA [189], but additional study into the ability of in vitro-
transcribed mRNA to induce anti-RNA autoantibodies and
worsen diseases where anti-RNA antibodies are present and
potentially pathogenic are needed.
The potential for immunogenicity of the mRNA-encoded
protein when delivered as a replacement for a defective or defi-
cient endogenous version needs to be considered. This has been
similarly addressed for other forms of gene therapy where the
delivered protein could potentially be seen as foreign, due to
the lack of the protein in the host.
The manufacturing of mRNA under current day techni-
ques is relatively simple. Therapeutic mRNA can be made in
a single reaction using template DNA and phage polymerase
with NTPs and buffer. There are no animal or cellular com-
ponents; thus, the risks are considerably lower compared to
recombinant proteins. An additional advantage of mRNA
over recombinant proteins for protein replacement therapies
is that the same procedure for making, purifying, and formu-
lating mRNA can be used for just about any mRNA encod-
ing a protein, while different purifications and formulations
are required for recombinant proteins. The manufacturing
process for producing mRNA has low variation and is easy to
control.
Expert commentary
Recently, with the identification of methods to reduce the immu-
nogenicity and increase the translational potential of mRNA,
there has been an expansion in research and development of
protein-encoding mRNA as a new therapeutic drug class. This
has fallen on the heels of a similar expansion of RNA therapeutics
involving siRNA, which has brought many new researchers and
hypotheses into the field. mRNA therapeutics is at the beginning
of its development, which informs us of the possibility that great
success and unexpected findings are likely to appear. The fields
of therapy where mRNA shows promise include vaccines, protein
replacement, gene editing and cell therapy, but new fields of use
are likely to be identified.
There has been a significant increase in interest in mRNA
therapeutics in the biotechnology arena and a number of uni-
versity derived and repurposed companies have developed,
including CureVac [190], BioNTech [191], Argos Therapeu-
tics [192], eTheRNA [193], Factor Bioscience [194], Trilink [195],
Ethris [196], Cellscript [197], and Moderna [198]. They have gar-
nered significant amounts of venture funding. A number of
pharmaceutical companies have also entered into the develop-
ment of mRNA-based therapeutics, including Shire, Novartis,
Astra Zeneca, Takeda, and Sanofi-Pasteur and a number of
others have expressed a peripheral interest and appear to be
waiting for potential products with greater development.
There are currently 33 clinical trials registered that are using
mRNA as a therapeutic (FIGURE 3). Most involve vaccines for a
variety of cancers. The most advanced are at Phase IIIB/IV,
suggesting the possibility of acceptance for clinical use in the
near future. It is expected with the expansion in biotech and
pharmaceutical companiesinterest in mRNA therapeutics that
there will be a significant expansion in clinical trails in all
aspects of mRNA therapy.
mRNA transcript therapy Review
informahealthcare.com doi: 10.1586/14760584.2015.973859
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
Enrolling by invitation
NCT00204516 Vaccination with tumor mRNA in metastatic melanoma ... Completed
NCT00204607 Intradermal vaccination with stabilized tumor mRNA-a
clinical phase l/lI trial in melanoma patients Completed
NCT00228189 Carcinoembryonic antigen-loaded dendritic cells in advanced
colorectal cancer patients
Completed
NCT00243529 Peptide-pulsed vs. RNA-transfected dendritic cell vaccines in
melanoma patients
Completed
NCT00510133 A study of active immunotherapy with GRNVAC1 in patients
with acute myelogenous leukemia (AML) Active, not recruiting
NCT00514189 Feasibility study of acute myelogenous leukemia mRNA plus
lysate loaded dendritic cell vaccines
Terminated
NCT00626483 Basiliximab in treating patients with newly diagnosed
glioblastoma multiforme ...
Recruiting
NCT00639639 Vaccine therapy in treating patients with newly diagnosed
glioblastoma multiforme
Active, not recruiting
NCT00833781 A pilot study of a dendritic cell vaccine in HIV-1 Infected ... Completed
NCT00834002 Dendritic cell vaccination for patients with acute myeloid
leukemia in remission
Completed
NCT00846456 Safe study of dendritic ceil (DC) based therapy targeting
tumor stem cells in glioblastoma Completed
NCT00890032 Vaccine therapy in treating patients undergoing surgery for
recurrent glioblastoma multiforme
Active, not recruiting
NCT00923312 Trial of an RNActive®-derived cancer vaccine in stage IIIB/IV
non small cell lung cancer (NSCLC)
Active, not recruiting
NCT00929019 Messenger ribonucleic acid (mRNA) transfected dendritic
ceil vaccination in high risk uveal melanoma patients
Recruiting
NCT00940004 Toll-like receptor (TLR) ligand matured dendritic cell
vaccination in melanoma patients
Active, not recruiting
NCT00961844 Trial for vaccine therapy with dendritic cells in patients with
metastatic malignant melanoma
Terminated
NCT00965224 Efficacy of dendritic cell therapy for myeloid leukemia and
myeloma
NCT00978913 Transfected dendritic cell based therapy for patients with
breast cancer or malignant melanoma
Recruiting
NCT01066390 A study on the safety and immunogenicity of... administration
of an autologous mRNA electroporated dendritic cell vaccine
in ... unresectable stage III or IV melanoma
Completed
NCT01197625 Vaccine therapy in curative resected prostate cancer
patients
Recruiting
NCT01278914 Trial of vaccine therapy with mRNA-transfected dendritic
cells in patients with androgen resistant metastatic prostate
cancer
Completed
NCT01278940 Trial of vaccine therapy with mRNA-transfected dendritic
cells in patients with advanced malignant melanoma
Completed
NCT01291420 Dendritic cell vaccination for patients with solid tumors Enrolling by invitation
NCT01334047 Trial of vaccine therapy in recurrent platinum sensitive
ovarian cancer patients
Recruiting
NCT01446731 Dendritic cell vaccination and docetaxei for patients with
prostate cancer
Recruiting
NCT01456065 Safety of active immunotherapy in subjects with ovarian
cancer
Active, not recruiting
NCT01456104 Immune responses to autologous langerhans-type dendritic
cells electroporated with mRNA encoding a tumor-associated
antigen in patients with malignancy:... melanoma
Recruiting
NCT01530698 Single-step antigen loading and TLR activation of dendritic
cells in melanoma patients
Completed
NCT01686334 Efficacy study of dendritic ceil vaccination in patients with
acute myeloid leukemia in remission
Recruiting
NCT01973322 Vaccination with autologous dendritic cells loaded with
autologous tumor lysate ... preleukapheresis IFN-alfa in ...
metastatic melano ma...
Recruiting
NCT01995708 CT7, MAGEA3, and WT1 mRNA electroporated autologous
langerhans-type dendritic cells ... multiple myeloma patients
undergoing autologous stem ceil transplantation
Recruiting
NCT02140138 An open label randomised trial of RNActive Cancer
vaccine in high risk and intermediate risk patients with
prostate cancer
Recruiting
Trial no. Title Status
EUDRACT:
2012-005572-34
A phase I study on the feasibility and safety of mRNA
immunotherapy in combination with RFA in patients with
hepatocellular carcinoma
Unknown
Figure 3. Clinical trials registered at clinicaltrials.gov using therapeutic mRNA. The clinicaltrials.gov and European Clinical Trials
Database (EudraCT) websites were queried for mRNA and results were analyzed for the use of mRNA as a therapeutic as of 8 January
2014.
Review Weissman
doi: 10.1586/14760584.2015.973859 Expert Rev. Vaccines
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
The choice of a target for mRNA drugs requires a thorough
understanding of the basic science, the clinical applicability and
the market. When an mRNA drug candidate is identified, the
process optimization and scale up to clinical grade GMP pro-
duction can be performed rapidly, because of the simplicity of
mRNA production. The expected costs for producing GMP
grade mRNA for clinical studies are likely 5–10-fold lower
compared to recombinant protein therapeutics that requires
complicated and specialized purification procedures that are dif-
ferent for every protein. The major difficulty for all mRNA
therapeutics that involve in vivo conveyance of the mRNA is
complexing and delivery. Most direct-injected mRNA vaccines
use uncomplexed mRNA for the delivery of mRNA that is to
be translated, which greatly simplifies these protocols. Naked
mRNA for protein replacement therapies or gene editing will
likely fail due to the inefficiency of uptake and the lack of pro-
tection from ubiquitous RNases. This leaves the field of
mRNA complexing open to new innovations. Beyond complex-
ing, the delivery of the mRNA to a particular cell type or tissue
is even more complex given all of the different cell types and
tissues and the complexity of targeting particular cells. The cur-
rent leader for in vivo delivery was borrowed from siRNA ther-
apeutics, LNPs [4244], that very efficiently target the liver. New
complexing agents that can be directed to other cell types and
organs are needed. At present, directing therapeutic develop-
ment to tissues and cells that can be more easily targeted (liver,
skin and respiratory tact) makes the most sense.
mRNA therapeutics is at an early stage of development, but
with the amount of interest from biotech and pharmaceutical
companies, it is expected to rapidly develop to clinical trials
and new drugs. Many obstacles are present in the path to
development, but the extremely broad distribution of potential
therapeutic uses, vaccines, protein replacement, gene editing,
and cell reprogramming and the wide distribution of potential
drugs within each of these categories, for example, protein
replacement can include both intracellular and extracellular
proteins, replacement versus new therapeutic, transient or
chronic treatment and systemic administration versus organ or
tissue-specific therapy has brought many new resources and a
large amount of capital into the field. I believe that mRNA
therapeutics will have an extremely broad base that will aid in
the development of new therapeutic avenues in the future.
Five-year view
mRNA vaccine therapy has been in clinical trials and has
shown initial promise for a number of different types of
cancer. Phase IIIb/VI clinical trials will start soon and will
be instrumental in propelling the field, if positive. mRNA
therapeutics is at the beginning of its development with no
clinical trials currently being performed. An enormous
amount of research and resources by biotech and pharma-
ceutical companies is involved in the development of
nucleoside-modified mRNA as a therapeutic. It is expected
that many different types of therapeutics will be developed,
including ones for acute treatment, inflammation, tissue
regeneration, gene editing, and chronic therapies, including
hormone, cytokine and monoclonal antibody treatment.
Studies in both fields of use of mRNA will benefit develop-
ment of each other, likely leading to rapidly advancing
expansion of effective therapeutics. It is likely that
nucleoside-modified mRNA encoding a therapeutic protein
will enter clinical trials within 2 years and begin a new
approach to treatment with therapeutic proteins with poten-
tial rapid expansion to other diseases.
Financial & competing interests disclosure
D Weissman is named in patents describing nucleoside modified mRNA as
a therapeutic. The author has no other relevant affiliations or financial
involvement with any organization or entity with a financial interest in
or financial conflict with the subject matter or materials discussed in the
manuscript apart from those disclosed.
No writing assistance was utilized in the production of this manuscript.
Key issues
• mRNA delivers therapeutic proteins to cells in vitro and in vivo, but its intrinsic immunogenicity has limited clinical development to
mainly vaccine studies.
Nucleoside modification reduces the ability of in vitro transcribed RNA to activate RNA innate immune sensors.
The further removal of phage polymerase produced contaminants, namely double-stranded RNA, can completely ablate the
immunogenicity of RNA containing certain nucleoside modifications.
• mRNA encoding a variety of antigens is an effective vaccine approach that can be delivered by direct injection of mRNA or ex vivo
treatment of cells.
• A major hurdle to mRNA therapeutics is the development of safe and effective in vivo delivery technologies, which is further
complicated when the targeting of particular cells or organs is required.
• mRNA therapy has therapeutic advantages, no risk of chromosomal integration, high transfection efficiency in both dividing and non-
dividing cells, superior modulation of the amount and duration of protein production, in many different fields, including vaccines, pro-
tein replacement, genome editing, cell fate reprogramming and cancer immunotherapy.
Nucleoside-modified mRNA has the potential to revolutionize protein therapeutics, but is at an early stage of development and requires
optimization and understanding of all elements of production and use.
mRNA transcript therapy Review
informahealthcare.com doi: 10.1586/14760584.2015.973859
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
References
Papers of special note have been highlighted as:
• of interest
•• of considerable interest
1. Wolff JA, Malone RW, Williams P, et al.
Direct gene transfer into mouse muscle in
vivo. Science 1990;247(4949 Pt 1):1465-8
The first use of mRNA with in vivo
injection to deliver a protein.
2. Jirikowski GF, Sanna PP,
Maciejewski-Lenoir D, Bloom FE. Reversal
of diabetes insipidus in Brattleboro rats:
intrahypothalamic injection of vasopressin
mRNA. Science 1992;255(5047):996-8
The first therapeutic use of mRNA to
treat a disease.
3. Bernal JA. RNA-based tools for nuclear
reprogramming and lineage-conversion:
towards clinical applications. J Cardiovasc
Transl Res 2013;6(6):956-68
4. Cannon G, Weissman D. RNA based
vaccines. DNA Cell Biol 2002;21(12):
953-61
5. Kreiter S, Diken M, Selmi A, et al. Tumor
vaccination using messenger RNA: prospects
of a future therapy. Curr Opin Immunol
2011;23(3):399-406
6. Kuhn AN, Beibetaert T, Simon P, et al.
mRNA as a versatile tool for exogenous
protein expression. Curr Gene Ther 2012;
12(5):347-61
7. Pascolo S. Vaccination with messenger
RNA (mRNA). Handbook of experimental
pharmacology 2008(183):221-35
8. Hacein-Bey-Abina S, von Kalle C,
Schmidt M, et al. A serious adverse event
after successful gene therapy for X-linked
severe combined immunodeficiency. N Engl
J Med 2003;348(3):255-6
9. Hacein-Bey-Abina S, Von Kalle C,
Schmidt M, et al. LMO2-associated clonal
T cell proliferation in two patients after
gene therapy for SCID-X1. Science 2003;
302(5644):415-19
10. Kapeli K, Yeo GW. Genome-wide
approaches to dissect the roles of
RNA binding proteins in translational
control: implications for neurological
diseases. Frontiers in neuroscience
2012;6:144
11. Lodish HF. Translational control of protein
synthesis: the early years. J Biol Chem
2012;287(43):36528-35
12. Morgens DW. The protein invasion:
a broad review on the origin of the
translational system. J Mol Evol 2013;77(4):
185-96
13. Roy B, Jacobson A. The intimate
relationships of mRNA decay and
translation. Trends Genet 2013;29(12):
691-9
14. Anderson BR, Muramatsu H, Jha BK, et al.
Nucleoside modifications in RNA limit
activation of 2-5-oligoadenylate synthetase
and increase resistance to cleavage by RNase
L. Nucleic Acids Res 2011;39(21):9329-38
15. Kariko K, Muramatsu H, Welsh FA, et al.
Incorporation of pseudouridine into
mRNA yields superior nonimmunogenic
vector with increased translational capacity
and biological stability. Mol Ther 2008;
16(11):1833-40
The first demonstration that
nucleoside-modified mRNA has reduced
immunogenicity and superior translation.
16. Kormann MS, Hasenpusch G, Aneja MK,
et al. Expression of therapeutic proteins
after delivery of chemically modified
mRNA in mice. Nat Biotechnol 2011;
29(2):154-7
17. Wang Y, Su HH, Yang Y, et al. Systemic
delivery of modified mRNA encoding
herpes simplex virus 1 thymidine kinase for
targeted cancer gene therapy. Mol Ther
2013;21(2):358-67
18. Hornung V, Ellegast J, Kim S, et al.
5-Triphosphate RNA is the ligand for
RIG-I. Science 2006;314(5801):994-7
19. Pichlmair A, Schulz O, Tan CP, et al.
RIG-I-mediated antiviral responses to
single-stranded RNA bearing 5-phosphates.
Science 2006;314(5801):997-1001
20. Anderson BR, Muramatsu H, Nallagatla SR
, et al. Incorporation of pseudouridine into
mRNA enhances translation by diminishing
PKR activation. Nucleic Acids Res 2010;
38(17):5884-92
21. Nallagatla SR, Bevilacqua PC. Nucleoside
modifications modulate activation of the
protein kinase PKR in an
RNA structure-specific manner. RNA 2008;
14(6):1201-13
22. Kariko K, Muramatsu H, Ludwig J,
Weissman D. Generating the optimal
mRNA for therapy: HPLC purification
eliminates immune activation and improves
translation of nucleoside-modified,
protein-encoding mRNA. Nucleic Acids Res
2011;39(21):e142
23. Holtkamp S, Kreiter S, Selmi A, et al.
Modification of antigen-encoding
RNA increases stability, translational
efficacy, and T-cell stimulatory capacity of
dendritic cells. Blood 2006;108(13):4009-17
24. Kuhn AN, Diken M, Kreiter S, et al.
Phosphorothioate cap analogs increase
stability and translational efficiency of
RNA vaccines in immature dendritic cells
and induce superior immune responses in
vivo. Gene Ther 2010;17(8):961-71
25. Dimitrov DS. Therapeutic proteins.
Methods in molecular biology 2012;899:
1-26
26. Leader B, Baca QJ, Golan DE. Protein
therapeutics: a summary and
pharmacological classification. Nat Rev
Drug discovery 2008;7(1):21-39
27. Lorenz C, Fotin-Mleczek M, Roth G, et al.
Protein expression from exogenous mRNA:
uptake by receptor-mediated endocytosis
and trafficking via the lysosomal pathway.
RNA Biol 2011;8(4):627-36
28. Van Tendeloo VF, Ponsaerts P, Lardon F,
et al. Highly efficient gene delivery by
mRNA electroporation in human
hematopoietic cells: superiority to
lipofection and passive pulsing of
mRNA and to electroporation of plasmid
cDNA for tumor antigen loading of
dendritic cells. Blood 2001;98(1):49-56
29. Chen AK, Behlke MA, Tsourkas A.
Efficient cytosolic delivery of molecular
beacon conjugates and flow cytometric
analysis of target RNA. Nucleic Acids Res
2008;36(12):e69
30. Besse F, Ephrussi A. Translational control
of localized mRNAs: restricting protein
synthesis in space and time. Nat Rev
Molecular cell biology 2008;9(12):971-80
31. Zubiaga AM, Belasco JG, Greenberg ME.
The nonamer UUAUUUAUU is the key
AU-rich sequence motif that mediates
mRNA degradation. Mol Cell Biol 1995;
15(4):2219-30
32. Meyer S, Temme C, Wahle E. Messenger
RNA turnover in eukaryotes: pathways and
enzymes. Crit Rev Biochem Mol Biol 2004;
39(4):197-216
33. Rittig SM, Haentschel M, Weimer KJ, et al.
Intradermal vaccinations with RNA coding
for TAA generate CD8+ and CD4+
immune responses and induce clinical
benefit in vaccinated patients. Mol Ther
2011;19(5):990-9
34. Weide B, Carralot JP, Reese A, et al.
Results of the first phase I/II clinical
vaccination trial with direct injection of
mRNA. J Immunother 2008;31(2):180-8
35. Weide B, Pascolo S, Scheel B, et al. Direct
injection of protamine-protected mRNA:
results of a phase 1/2 vaccination trial in
metastatic melanoma patients. J
Immunother 2009;32(5):498-507
36. Trial of RNActive-Derived Cancer
Vaccine and Local Radiation in in Stage IV
Review Weissman
doi: 10.1586/14760584.2015.973859 Expert Rev. Vaccines
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
Non Small Cell Lung Cancer (NSCLC).
Available from: https://clinicaltrials.gov/ct2/
results?term=NCT01915524&Search=Search
37. Trial of RNActive-Derived Prostate
Cancer Vaccine in Metastatic
Castrate-refractory Prostate Cancer.
Available from: https://clinicaltrials.gov/ct2/
results?term=NCT01817738&Search=Search
38. Kreiter S, Selmi A, Diken M, et al.
Intranodal vaccination with naked
antigen-encoding RNA elicits potent
prophylactic and therapeutic antitumoral
immunity. Cancer Res 2010;70(22):
9031-40
39. Van Lint S, Goyvaerts C, Maenhout S,
et al. Preclinical evaluation of TriMix and
antigen mRNA-based antitumor therapy.
Cancer Res 2012;72(7):1661-71
40. RBL001/RBL002 Phase I Clinical Trial.
Available from: https://clinicaltrials.gov/ct2/
results?term=NCT01684241&Search=Search
41. Kariko K, Muramatsu H, Keller JM,
Weissman D. Increased erythropoiesis in
mice injected with submicrogram quantities
of pseudouridine-containing
mRNA encoding erythropoietin. Mol Ther
2012;20(5):948-53
42. Morrissey DV, Lockridge JA, Shaw L, et al.
Potent and persistent in vivo anti-HBV
activity of chemically modified siRNAs. Nat
Biotechnol 2005;23(8):1002-7
43. Zimmermann TS, Lee AC, Akinc A, et al.
RNAi-mediated gene silencing in
non-human primates. Nature 2006;
441(7089):111-14
44. Brower V. RNA interference advances to
early-stage clinical trials. J Natl Cancer Inst
2010;102(19):1459-61
45. Davis ME, Zuckerman JE, Choi CH, et al.
Evidence of RNAi in humans from
systemically administered siRNA via
targeted nanoparticles. Nature 2010;
464(7291):1067-70
46. Shi B, Keough E, Matter A, et al.
Biodistribution of small interfering RNA at
the organ and cellular levels after lipid
nanoparticle-mediated delivery. The journal
of histochemistry and cytochemistry :
official journal of the Histochemistry
Society 2011;59(8):727-40
47. Ziemniak M, Strenkowska M, Kowalska J,
Jemielity J. Potential therapeutic
applications of RNA cap analogs. Future
medicinal chemistry 2013;5(10):1141-72
48. Li Y, Kiledjian M. Regulation of
mRNA decapping. Wiley interdisciplinary
reviews. RNA 2010;1(2):253-65
49. Ensinger MJ, Martin SA, Paoletti E,
Moss B. Modification of the 5-terminus of
mRNA by soluble guanylyl and methyl
transferases from vaccinia virus. Proc Natl
Acad Sci USA 1975;72(7):2525-9
50. Martin SA, Paoletti E, Moss B. Purification
of mRNA guanylyltransferase and
mRNA (guanine-7-) methyltransferase from
vaccinia virions. J Biol Chem 1975;250(24):
9322-9
51. Malone RW, Felgner PL, Verma IM.
Cationic liposome-mediated
RNA transfection. Proc Natl Acad Sci USA
1989;86(16):6077-81
52. Pasquinelli AE, Dahlberg JE, Lund E.
Reverse 5caps in RNAs made in vitro by
phage RNA polymerases. RNA 1995;1(9):
957-67
53. Jemielity J, Fowler T, Zuberek J, et al.
Novel anti-reversecap analogs with
superior translational properties. RNA 2003;
9(9):1108-22
54. Stepinski J, Waddell C, Stolarski R, et al.
Synthesis and properties of mRNAs
containing the novel anti-reversecap
analogs 7-methyl(3-O-methyl)GpppG and
7-methyl (3-deoxy)GpppG. RNA 2001;
7(10):1486-95
55. Bontkes HJ, Kramer D, Ruizendaal JJ, et al.
Dendritic cells transfected with
interleukin-12 and tumor-associated antigen
messenger RNA induce high avidity
cytotoxic T cells. Gene Ther 2007;14(4):
366-75
56. Mockey M, Goncalves C, Dupuy FP, et al.
mRNA transfection of dendritic cells:
synergistic effect of ARCA mRNA capping
with Poly(A) chains in cis and in trans for a
high protein expression level. Biochem
Biophys Res Commun 2006;340(4):1062-8
57. Rabinovich PM, Komarovskaya ME, Ye ZJ,
et al. Synthetic messenger RNA as a tool for
gene therapy. Hum Gene Ther 2006;
17(10):1027-35
58. Grudzien-Nogalska E, Jemielity J,
Kowalska J, et al. Phosphorothioate cap
analogs stabilize mRNA and increase
translational efficiency in mammalian cells.
RNA 2007;13(10):1745-55
59. Kowalska J, Lewdorowicz M, Zuberek J,
et al. Synthesis and characterization of
mRNA cap analogs containing
phosphorothioate substitutions that bind
tightly to eIF4E and are resistant to the
decapping pyrophosphatase DcpS. RNA
2008;14(6):1119-31
60. Jia J, Yao P, Arif A, Fox PL. Regulation
and dysregulation of 3UTR-mediated
translational control. Curr Opin Genet Dev
2013;23(1):29-34
61. Perez-Ortin JE, Alepuz P, Chavez S,
Choder M. Eukaryotic mRNA decay:
methodologies, pathways, and links to other
stages of gene expression. J Mol Biol 2013;
425(20):3750-75
62. Somers J, Poyry T, Willis AE. A perspective
on mammalian upstream open reading
frame function. Int J Biochem Cell Biol
2013;45(8):1690-700
63. Szostak E, Gebauer F. Translational control
by 3-UTR-binding proteins. Briefings in
functional genomics 2013;12(1):58-65
64. Ross J, Sullivan TD. Half-lives of beta and
gamma globin messenger RNAs and of
protein synthetic capacity in cultured
human reticulocytes. Blood 1985;66(5):
1149-54
65. Yu J, Russell JE. Structural and functional
analysis of an mRNP complex that mediates
the high stability of human beta-globin
mRNA. Mol Cell Biol 2001;21(17):
5879-88
66. Benteyn D, Anguille S, Van Lint S, et al.
Design of an Optimized WilmsTumor 1
(WT1) mRNA Construct for Enhanced
WT1 Expression and Improved
Immunogenicity In Vitro and In Vivo. Mol
Ther Nucleic acids 2013;2:e134
67. Chen CY, Shyu AB. AU-rich elements:
characterization and importance in
mRNA degradation. Trends Biochem Sci
1995;20(11):465-70
68. Belmont BJ, Niles JC. Engineering a direct
and inducible protein-RNA interaction to
regulate RNA biology. ACS Chem Biol
2010;5(9):851-61
69. Clery A, Bourguignon-Igel V, Allmang C,
et al. An improved definition of the
RNA-binding specificity of SECIS-binding
protein 2, an essential component of the
selenocysteine incorporation machinery.
Nucleic Acids Res 2007;35(6):1868-84
70. Bakke O, Nordeng TW. Intracellular traffic
to compartments for MHC class II peptide
loading: signals for endosomal and polarized
sorting. Immunol Rev 1999;172:171-87
71. Marks MS. Protein sorting within the
MHC class II antigen-processing pathway.
Immunol Res 1998;17(1-2):141-54
72. Bonehill A, Heirman C, Tuyaerts S, et al.
Messenger RNA-electroporated dendritic
cells presenting MAGE-A3 simultaneously
in HLA class I and class II molecules.
Journal of immunology 2004;172(11):
6649-57
mRNA transcript therapy Review
informahealthcare.com doi: 10.1586/14760584.2015.973859
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
73. Zhong F, Cao W, Chan E, et al. Deviation
from major codons in the Toll-like receptor
genes is associated with low Toll-like
receptor expression. Immunology 2005;
114(1):83-93
74. Ngumbela KC, Ryan KP, Sivamurthy R,
et al. Quantitative effect of suboptimal
codon usage on translational efficiency of
mRNA encoding HIV-1 gag in intact T
cells. PLoS ONE 2008;3(6):e2356
75. Van Gulck ER, Ponsaerts P, Heyndrickx L,
et al. Efficient stimulation of HIV-1-specific
T cells using dendritic cells electroporated
with mRNA encoding autologous
HIV-1 Gag and Env proteins. Blood 2006;
107(5):1818-27
76. Kimchi-Sarfaty C, Oh JM, Kim IW, et al.
Asilentpolymorphism in the
MDR1 gene changes substrate specificity.
Science 2007;315(5811):525-8
77. Ho O, Green WR. Alternative translational
products and cryptic T cell epitopes:
expecting the unexpected. J Immunol 2006;
177(12):8283-9
78. Ho O, Green WR. Cytolytic CD8+ T cells
directed against a cryptic epitope derived
from a retroviral alternative reading frame
confer disease protection. Journal of
immunology 2006;176(4):2470-5
79. Malarkannan S, Horng T, Shih PP, et al.
Presentation of out-of-frame peptide/MHC
class I complexes by a novel translation
initiation mechanism. Immunity 1999;
10(6):681-90
80. Saulquin X, Scotet E, Trautmann L, et al.
+1 Frameshifting as a novel mechanism to
generate a cryptic cytotoxic T lymphocyte
epitope derived from human interleukin 10.
J Exp Med 2002;195(3):353-8
81. Schirmbeck R, Riedl P, Fissolo N, et al.
Translation from cryptic reading frames of
DNA vaccines generates an extended
repertoire of immunogenic, MHC class
I-restricted epitopes. J Immunol 2005;
174(8):4647-56
82. Schwab SR, Li KC, Kang C, Shastri N.
Constitutive display of cryptic translation
products by MHC class I molecules. Science
2003;301(5638):1367-71
83. Gallie DR. The cap and poly(A) tail
function synergistically to regulate
mRNA translational efficiency. Genes Dev
1991;5(11):2108-16
84. Tcherepanova IY, Adams MD, Feng X,
et al. Ectopic expression of a truncated
CD40L protein from synthetic
post-transcriptionally capped RNA in
dendritic cells induces high levels of
IL-12 secretion. BMC Mol Biol 2008;9:90
85. Brencicova E, Diebold SS. Nucleic acids
and endosomal pattern recognition: how to
tell friend from foe? Frontiers in cellular
and infection microbiology 2013;3:37
86. Cerboni S, Gentili M, Manel N. Diversity
of pathogen sensors in dendritic cells. Adv
Immunol 2013;120:211-37
87. Crampton SP, Bolland S. Spontaneous
activation of RNA-sensing pathways in
autoimmune disease. Curr Opin Immunol
2013;25(6):712-19
88. Gantier MP. Processing of double-stranded
RNA in Mammalian cells: a direct antiviral
role? J Interferon Cytokine Res 2014;34(6):
469-77
89. Kindler E, Thiel V. To sense or not to
sense viral RNA-essentials of coronavirus
innate immune evasion. Curr Opin
Microbiol 2014;20C:69-75
90. Nellimarla S, Mossman KL. Extracellular
dsRNA: its function and mechanism of
cellular uptake. J Interferon Cytokine Res
2014;34(6):419-26
91. Rawling DC, Pyle AM. Parts, assembly and
operation of the RIG-I family of motors.
Curr Opin Struct Biol 2014;25:25-33
92. Vabret N, Blander JM. Sensing microbial
RNA in the cytosol. Frontiers in
immunology 2013;4:468
93. Wu J, Chen ZJ. Innate immune sensing
and signaling of cytosolic nucleic acids.
Annu Rev Immunol 2014;32:461-88
94. Goubau D, Deddouche S, Reis e Sousa C.
Cytosolic sensing of viruses. Immunity
2013;38(5):855-69
95. Pollard C, Rejman J, De Haes W, et al.
Type I IFN counteracts the induction of
antigen-specific immune responses by
lipid-based delivery of mRNA vaccines. Mol
Ther 2013;21(1):251-9
96. Bourquin C, Schmidt L, Hornung V, et al.
Immunostimulatory RNA oligonucleotides
trigger an antigen-specific cytotoxic T-cell
and IgG2a response. Blood 2007;109(7):
2953-60
97. Bourquin C, Wurzenberger C, Heidegger S,
et al. Delivery of immunostimulatory
RNA oligonucleotides by gelatin
nanoparticles triggers an efficient
antitumoral response. J Immunother 2010;
33(9):935-44
98. Sander LE, Davis MJ, Boekschoten MV,
et al. Detection of prokaryotic
mRNA signifies microbial viability and
promotes immunity. Nature 2011;
474(7351):385-9
99. Hoerr I, Obst R, Rammensee HG, Jung G.
In vivo application of RNA leads to
induction of specific cytotoxic T
lymphocytes and antibodies. Eur J Immunol
2000;30(1):1-7
100. Weissman D, Ni H, Scales D, et al. HIV
gag mRNA transfection of dendritic cells
(DC) delivers encoded antigen to MHC
class I and II molecules, causes DC
maturation, and induces a potent human
in vitro primary immune response. Journal
of immunology 2000;165(8):4710-17
101. Fotin-Mleczek M, Duchardt KM, Lorenz C,
et al. Messenger RNA-based vaccines with
dual activity induce balanced
TLR-7 dependent adaptive immune
responses and provide antitumor activity. J
Immunother 2011;34(1):1-15
102. Rettig L, Haen SP, Bittermann AG, et al.
Particle size and activation threshold: a new
dimension of danger signaling. Blood 2010;
115(22):4533-41
103. 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(2):165-75
•• The first demonstration that nucleoside
modification reduces RNA immunogenicity.
104. Arnaud-Barbe N, Cheynet-Sauvion V,
Oriol G, et al. Transcription of
RNA templates by T7 RNA polymerase.
Nucleic Acids Res 1998;26(15):3550-4
105. Milligan JF, Groebe DR, Witherell GW,
Uhlenbeck OC. Oligoribonucleotide
synthesis using T7 RNA polymerase and
synthetic DNA templates. Nucleic Acids Res
1987;15(21):8783-98
106. Nacheva GA, Berzal-Herranz A. Preventing
nondesired RNA-primed RNA extension
catalyzed by T7 RNA polymerase. European
journal of biochemistry/ FEBS 2003;270(7):
1458-65
107. Triana-Alonso FJ, Dabrowski M, Wadzack J,
Nierhaus KH. Self-coded 3-extension of
run-off transcripts produces aberrant
products during in vitro transcription with
T7 RNA polymerase. J Biol Chem 1995;
270(11):6298-307
108. Conry RM, LoBuglio AF, Wright M, et al.
Characterization of a messenger
RNA polynucleotide vaccine vector. Cancer
Res 1995;55(7):1397-400
109. Heiser A, Coleman D, Dannull J, et al.
Autologous dendritic cells transfected with
prostate-specific antigen RNA stimulate
CTL responses against metastatic prostate
tumors. J Clin Invest 2002;109(3):409-17
110. Nair SK, Boczkowski D, Morse M, et al.
Induction of primary carcinoembryonic
Review Weissman
doi: 10.1586/14760584.2015.973859 Expert Rev. Vaccines
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
antigen (CEA)-specific cytotoxic T
lymphocytes in vitro using human dendritic
cells transfected with RNA. Nat Biotechnol
1998;16(4):364-9
111. Nair SK, Heiser A, Boczkowski D, et al.
Induction of cytotoxic T cell responses and
tumor immunity against unrelated tumors
using telomerase reverse transcriptase
RNA transfected dendritic cells. Nat Med
2000;6(9):1011-17
112. Rains N, Cannan RJ, Chen W, Stubbs RS.
Development of a dendritic cell (DC)-based
vaccine for patients with advanced colorectal
cancer. Hepatogastroenterology 2001;
48(38):347-51
113. Van Lint S, Heirman C, Thielemans K.
Breckpot K. mRNA: from a chemical
blueprint for protein production to an
off-the-shelf therapeutic. Hum Vaccin
Immunother 2013;9(2):265-74
114. Bonehill A, Tuyaerts S, Van Nuffel AM,
et al. Enhancing the T-cell stimulatory
capacity of human dendritic cells by
co-electroporation with CD40L, CD70 and
constitutively active TLR4 encoding mRNA.
Mol Ther 2008;16(6):1170-80
115. Calderhead DM, DeBenedette MA,
Ketteringham H, et al. Cytokine maturation
followed by CD40L mRNA electroporation
results in a clinically relevant dendritic cell
product capable of inducing a potent
proinflammatory CTL response. J
Immunother 2008;31(8):731-41
116. Cisco RM, Abdel-Wahab Z, Dannull J,
et al. Induction of human dendritic cell
maturation using transfection with
RNA encoding a dominant positive toll-like
receptor 4. Journal of immunology 2004;
172(11):7162-8
117. Routy JP, Boulassel MR, Yassine-Diab B,
et al. Immunologic activity and safety of
autologous HIV RNA-electroporated
dendritic cells in HIV-1 infected patients
receiving antiretroviral therapy. Clinical
immunology 2010;134(2):140-7
118. Van Lint S, Wilgenhof S, Heirman C, et al.
Optimized dendritic cell-based
immunotherapy for melanoma: the
TriMix-formula. Cancer Immunol
Immunother 2014;63(9):959-67
119. Wilgenhof S, Van Nuffel AM, Benteyn D,
et al. A phase IB study on intravenous
synthetic mRNA electroporated dendritic
cell immunotherapy in pretreated advanced
melanoma patients. Ann Oncol 2013;
24(10):2686-93
Very impressive clinical results in a
Phase I trial with disease stabilization.
120. Phase 3 Trial of Autologous Dendritic Cell
Immunotherapy (AGS-003) Plus Standard
Treatment of Advanced Renal Cell
Carcinoma (RCC). Available from: https://
clinicaltrials.gov/ct2/results?
term=NCT01582672&Search=Search
121. Fotin-Mleczek M, Zanzinger K,
Heidenreich R, et al. Highly potent
mRNA based cancer vaccines represent an
attractive platform for combination
therapies supporting an improved
therapeutic effect. J Gene Med 2012;14(6):
428-39
122. Diken M, Kreiter S, Vascotto F, et al.
mTOR inhibition improves antitumor
effects of vaccination with antigen-encoding
RNA. Cancer Immunol Res 2013;1(6):
386-92
123. Kreiter S, Diken M, Selmi A, et al.
FLT3 ligand enhances the cancer
therapeutic potency of naked RNA vaccines.
Cancer Res 2011;71(19):6132-42
124. Van Driessche A, Van de Velde AL, Nijs G,
et al. Clinical-grade manufacturing of
autologous mature mRNA-electroporated
dendritic cells and safety testing in acute
myeloid leukemia patients in a phase I
dose-escalation clinical trial. Cytotherapy
2009;11(5):653-68
125. Castle JC, Kreiter S, Diekmann J, et al.
Exploiting the mutanome for tumor
vaccination. Cancer Res 2012;72(5):
1081-91
126. Kreiter S, Castle JC, Tureci O, Sahin U.
Targeting the tumor mutanome for
personalized vaccination therapy.
Oncoimmunology 2012;1(5):768-9
127. IVAC MUTANOME Phase I Clinical
Trial. Available from: https://clinicaltrials.
gov/ct2/results?term=NCT02035956&
Search=Search
128. Petsch B, Schnee M, Vogel AB, et al.
Protective efficacy of in vitro synthesized,
specific mRNA vaccines against influenza
A virus infection. Nat Biotechnol 2012;
30(12):1210-16
129. Mandl CW, Aberle JH, Aberle SW, et al.
In vitro-synthesized infectious RNA as an
attenuated live vaccine in a flavivirus model.
Nat Med 1998;4(12):1438-40
130. Anraku I, Harvey TJ, Linedale R, et al.
Kunjin virus replicon vaccine vectors induce
protective CD8+ T-cell immunity. J Virol
2002;76(8):3791-9
131. Fleeton MN, Chen M, Berglund P, et al.
Self-replicative RNA vaccines elicit
protection against influenza A virus,
respiratory syncytial virus, and a tick borne
encephalitis virus. J Infect Dis 2001;183(9):
1395-8
132. Greer CE, Zhou F, Legg HS, et al.
A chimeric alphavirus RNA replicon
gene-based vaccine for human parainfluenza
virus type 3 induces protective immunity
against intranasal virus challenge. Vaccine
2007;25(3):481-9
133. Geall AJ, Verma A, Otten GR, et al.
Nonviral delivery of self-amplifying
RNA vaccines. Proc Natl Acad Sci USA
2012;109(36):14604-9
134. Brito LA, Chan M, Shaw CA, et al.
A cationic nanoemulsion for the delivery of
next generation RNA vaccines. Mol Ther
2014;doi: 10.1038/mt.2014.133
135. Allard SD, De Keersmaecker B,
de Goede AL, et al. A phase I/IIa
immunotherapy trial of HIV-1-infected
patients with Tat, Rev and Nef expressing
dendritic cells followed by treatment
interruption. Clinical Immunol 2012;
142(3):252-68
136. Van Gulck E, Vlieghe E, Vekemans M,
et al. mRNA-based dendritic cell
vaccination induces potent antiviral T-cell
responses in HIV-1-infected patients. Aids
2012;26(4):F1-12
137. Baena-Cagnani CE, Gomez RM. Current
status of therapy with omalizumab in
children. Curr Opin Allergy Clin Immunol
2014;14(2):149-54
138. Keet CA, Wood RA. Emerging therapies for
food allergy. J Clin Invest 2014;124(5):
1880-6
139. Normansell R, Walker S, Milan SJ, et al.
Omalizumab for asthma in adults and
children. Cochrane Database Syst Rev
2014;1:Cd003559
140. Vashisht P, Casale T. Omalizumab for
treatment of allergic rhinitis. Expert Opin
Biol Ther 2013;13(6):933-45
141. Roesler E, Weiss R, Weinberger EE, et al.
Immunize and disappear-safety-optimized
mRNA vaccination with a panel of
29 allergens. J Allergy Clin Immunol 2009;
124(5):1070-7.e1071-1011
142. Weinberger EE, Isakovic A, Scheiblhofer S,
et al. The influence of antigen targeting to
sub-cellular compartments on the
anti-allergic potential of a DNA vaccine.
Vaccine 2013;31(51):6113-21
143. Weiss R, Scheiblhofer S, Roesler E, et al.
Prophylactic mRNA vaccination against
allergy. Curr Opin Allergy Clin Immunol
2010;10(6):567-74
144. Weiss R, Scheiblhofer S, Roesler E, et al.
mRNA vaccination as a safe approach for
mRNA transcript therapy Review
informahealthcare.com doi: 10.1586/14760584.2015.973859
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
specific protection from type I allergy.
Expert Rev Vaccines 2012;11(1):55-67
145. Weiss R, Scheiblhofer S, Thalhamer J.
Allergens are not pathogens: Why
immunization against allergy differs from
vaccination against infectious diseases. Hum
Vaccin Immunother 2013;10:3
146. Maus MV, Grupp SA, Porter DL, June CH.
Antibody-modified T cells: CARs take the
front seat for hematologic malignancies.
Blood 2014;123(17):2625-35
147. Almasbak H, Rian E, Hoel HJ, et al.
Transiently redirected T cells for adoptive
transfer. Cytotherapy 2011;13(5):629-40
148. Barrett DM, Liu X, Jiang S, et al.
Regimen-specific effects of RNA-modified
chimeric antigen receptor T cells in mice
with advanced leukemia. Hum Gene Ther
2013;24(8):717-27
149. Barrett DM, Zhao Y, Liu X, et al.
Treatment of advanced leukemia in mice
with mRNA engineered T cells. Hum Gene
Ther 2011;22(12):1575-86
150. Zhao Y, Moon E, Carpenito C, et al.
Multiple injections of electroporated
autologous T cells expressing a chimeric
antigen receptor mediate regression of
human disseminated tumor. Cancer Res
2010;70(22):9053-61
151. Autologous Redirected RNA Meso CAR T
Cells for Pancreatic Cancer. Available from:
https://clinicaltrials.gov/ct2/results?
term=NCT01897415&Search=Search
152. Hockemeyer D, Wang H, Kiani S, et al.
Genetic engineering of human pluripotent
cells using TALE nucleases. Nat Biotechnol
2011;29(8):731-4
153. Miller JC, Holmes MC, Wang J, et al. An
improved zinc-finger nuclease architecture
for highly specific genome editing. Nat
Biotechnol 2007;25(7):778-85
154. Mali P, Aach J, Stranges PB, et al.
CAS9 transcriptional activators for target
specificity screening and paired nickases for
cooperative genome engineering. Nat
Biotechnol 2013;31(9):833-8
155. Mali P, Yang L, Esvelt KM, et al.
RNA-guided human genome engineering
via Cas9. Science 2013;339(6121):823-6
156. Fu Y, Foden JA, Khayter C, et al.
High-frequency off-target mutagenesis
induced by CRISPR-Cas nucleases in
human cells. Nature biotechnology 2013;
31(9):822-6
157. Auer TO, Duroure K, De Cian A, et al.
Highly efficient CRISPR/Cas9-mediated
knock-in in zebrafish by
homology-independent DNA repair.
Genome Res 2014;24(1):142-53
158. Bedell VM, Wang Y, Campbell JM, et al.
In vivo genome editing using a
high-efficiency TALEN system. Nature
2012;491(7422):114-18
159. Geurts AM, Cost GJ, Freyvert Y, et al.
Knockout rats via embryo microinjection of
zinc-finger nucleases. Science 2009;
325(5939):433
160. Hwang WY, Fu Y, Reyon D, et al. Efficient
genome editing in zebrafish using a
CRISPR-Cas system. Nat Biotechnol 2013;
31(3):227-9
161. Tesson L, Usal C, Menoret S, et al.
Knockout rats generated by embryo
microinjection of TALENs. Nat Biotechnol
2011;29(8):695-6
162. Wang H, Yang H, Shivalila CS, et al.
One-step generation of mice carrying
mutations in multiple genes by CRISPR/
Cas-mediated genome engineering. Cell
2013;153(4):910-18
163. Wefers B, Panda SK, Ortiz O, et al.
Generation of targeted mouse mutants by
embryo microinjection of TALEN mRNA.
Nat Protoc 2013;8(12):2355-79
164. Wood AJ, Lo TW, Zeitler B, et al.
Targeted genome editing across species
using ZFNs and TALENs. Science 2011;
333(6040):307
165. Yamanaka S, Takahashi K. [Induction of
pluripotent stem cells from mouse fibroblast
cultures]. Tanpakushitsu kakusan koso
2006;51(15):2346-51
166. Warren L, Ni Y, Wang J, Guo X.
Feeder-free derivation of human induced
pluripotent stem cells with messenger RNA.
Sci Rep 2012;2:657
167. Yakubov E, Rechavi G, Rozenblatt S,
Givol D. Reprogramming of human
fibroblasts to pluripotent stem cells using
mRNA of four transcription factors.
Biochem Biophys Res Commun 2010;
394(1):189-93
168. Miller JD, Ganat YM, Kishinevsky S, et al.
Human iPSC-based modeling of late-onset
disease via progerin-induced aging. Cell
Stem Cell 2013;13(6):691-705
169. Okano H, Nakamura M, Yoshida K, et al.
Steps toward safe cell therapy using induced
pluripotent stem cells. Circ Res 2013;
112(3):523-33
170. Kariko K, Weissman D. Naturally occurring
nucleoside modifications suppress the
immunostimulatory activity of RNA:
implication for therapeutic
RNA development. Curr Opin Drug Discov
Dev 2007;10(5):523-32
171. Mays LE, Ammon-Treiber S, Mothes B,
et al. Modified Foxp3 mRNA protects
against asthma through an IL-10-dependent
mechanism. J Clin Invest 2013;123(3):
1216-28
172. Zangi L, Lui KO, von Gise A, et al.
Modified mRNA directs the fate of heart
progenitor cells and induces vascular
regeneration after myocardial infarction. Nat
Biotechnol 2013;31(10):898-907
Preclinical use of nucleoside-modified
mRNA in a myocardial infarction model.
173. Levy O, Zhao W, Mortensen LJ, et al.
mRNA-engineered mesenchymal stem cells
for targeted delivery of interleukin-10 to
sites of inflammation. Blood 2013;122(14):
e23-32
174. Weissman D, Pardi N, Muramatsu H,
Kariko K. HPLC purification of in vitro
transcribed long RNA. Methods Mol Biol
2013;969:43-54
175. Fattori E, Cappelletti M, Zampaglione I,
et al. Gene electro-transfer of an improved
erythropoietin plasmid in mice and
non-human primates. J Gene Med 2005;
7(2):228-36
176. Lee S, Lee Y, Hong K, et al. Effect of
recombinant alpha1-antitrypsin Fc-fused
(AAT-Fc)protein on the inhibition of
inflammatory cytokine production and
streptozotocin-induced diabetes. Mol Med
2013;19:65-71
177. Barash S, Wang W, Shi Y. Human
secretory signal peptide description by
hidden Markov model and generation of a
strong artificial signal peptide for secreted
protein expression. Biochem Biophys Res
Commun 2002;294(4):835-42
178. Kolarich D, Weber A, Turecek PL, et al.
Comprehensive glyco-proteomic analysis of
human alpha1-antitrypsin and its charge
isoforms. Proteomics 2006;6(11):3369-80
179. Sontag EM, Vonk WI, Frydman J. Sorting
out the trash: the spatial nature of
eukaryotic protein quality control. Curr
Opin Cell Biol 2014;26:139-46
180. Seidah NG, Chretien M. Proprotein and
prohormone convertases: a family of
subtilases generating diverse bioactive
polypeptides. Brain Res 1999;848(1-2):
45-62
181. Nakayama K. Furin: a mammalian
subtilisin/Kex2p-like endoprotease involved
in processing of a wide variety of precursor
proteins. Biochem J 1997;327(Pt 3):625-35
Review Weissman
doi: 10.1586/14760584.2015.973859 Expert Rev. Vaccines
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
182. Koren E, Smith HW, Shores E, et al.
Recommendations on risk-based strategies
for detection and characterization of
antibodies against biotechnology products. J
Immunol Methods 2008;333(1-2):1-9
183. Murphy K. The Humoral Immune
Response. In: Janeway C, editor.
Immunobiology. Garland Science
Publishing; New York: 2011. p. 367-408
184. Gao G, Lebherz C, Weiner DJ, et al.
Erythropoietin gene therapy leads to
autoimmune anemia in macaques. Blood
2004;103(9):3300-2
185. Casadevall N, Nataf J, Viron B, et al. Pure
red-cell aplasia and antierythropoietin
antibodies in patients treated with
recombinant erythropoietin. N Engl J Med
2002;346(7):469-75
186. Cournoyer D, Toffelmire EB, Wells GA,
et al. Anti-erythropoietin antibody-mediated
pure red cell aplasia after treatment with
recombinant erythropoietin products:
recommendations for minimization of risk.
J Am Soc Nephrol 2004;15(10):2728-34
187. Kromminga A, Schellekens H. Antibodies
against erythropoietin and other
protein-based therapeutics: an overview.
Ann N Y Acad Sci 2005;1050:257-65
188. Hwang SH, Lee H, Yamamoto M, et al. B
cell TLR7 expression drives
anti-RNA autoantibody production and
exacerbates disease in systemic lupus
erythematosus-prone mice. Journal of
immunology 2012;189(12):5786-96
189. Lipes BD, Keene JD. Autoimmune epitopes
in messenger RNA. RNA 2002;8(6):762-71
190. Available from: www.curevac.com/
191. Available from: www.biontech.de/
192. Available from: www.argostherapeutics.com/
193. Available from: www.etherna.be/
194. Available from: http://factorbio.com/
195. Available from: www.trilinkbiotech.com/
196. Available from: http://ethris.com/
197. Available from: www.cellscript.com/
198. Available from: http://modernatx.com/
mRNA transcript therapy Review
informahealthcare.com doi: 10.1586/14760584.2015.973859
Expert Review of Vaccines Downloaded from informahealthcare.com by University of Pennsylvania on 11/07/14
For personal use only.
... In recent years, mRNA therapy has shown considerable potential in the treatment of various diseases [1][2][3][4][5][6]. The rapid development and effective use of mRNA-based vaccines in combating the COVID-19 pandemic [7][8][9][10] has further propelled the adoption of this technology in various clinical directions, including the prevention of infectious and treatment of hereditary diseases, monoclonal antibody production, and cancer immunotherapy [2,[11][12][13][14]. ...
... The diverse clinical applications of mRNA therapy [1][2][3][4][5][11][12][13][14] underscore the vital importance of developing a broad array of effective delivery systems with varied properties. To date, dozens of cationic and ionizable lipid-based systems providing effective mRNA delivery have been described (reviewed in [14][15][16]33,35]). ...
Article
Full-text available
Over the past decade, mRNA-based therapy has displayed significant promise in a wide range of clinical applications. The most striking example of the leap in the development of mRNA technologies was the mass vaccination against COVID-19 during the pandemic. The emergence of large-scale technology and positive experience of mRNA immunization sparked the development of antiviral and anti-cancer mRNA vaccines as well as therapeutic mRNA agents for genetic and other diseases. To facilitate mRNA delivery, lipid nanoparticles (LNPs) have been successfully employed. However, the diverse use of mRNA therapeutic approaches requires the development of adaptable LNP delivery systems that can control the kinetics of mRNA uptake and expression in target cells. Here, we report effective mRNA delivery into cultured mammalian cells (HEK293T, HeLa, DC2.4) and living mouse muscle tissues by liposomes containing either 1,26-bis(cholest-5-en-3β-yloxycarbonylamino)-7,11,16,20-tetraazahexacosane tetrahydrochloride (2X3) or the newly applied 1,30-bis(cholest-5-en-3β-yloxycarbonylamino)-9,13,18,22-tetraaza-3,6,25,28-tetraoxatriacontane tetrahydrochloride (2X7) cationic lipids. Using end-point and real-time monitoring of Fluc mRNA expression, we showed that these LNPs exhibited an unusually delayed (of over 10 h in the case of the 2X7-based system) but had highly efficient and prolonged reporter activity in cells. Accordingly, both LNP formulations decorated with 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[amino(polyethylene glycol)-2000] (DSPE-PEG2000) provided efficient luciferase production in mice, peaking on day 3 after intramuscular injection. Notably, the bioluminescence was observed only at the site of injection in caudal thigh muscles, thereby demonstrating local expression of the model gene of interest. The developed mRNA delivery systems hold promise for prophylactic applications, where sustained synthesis of defensive proteins is required, and open doors to new possibilities in mRNA-based therapies.
... In the end, mRNA translation efficiency may be predicted using a bioinformatics method based on the 5′UTR sequence. 74 To control the translation elongation rate, the ORF's G and cytosine (C) contents is able to be optimized. Another codon optimization technique that has a strong connection to a higher level of GC is uridine removal. ...
... Although 250 units is a length of Poly (A) that is frequently utilized, various cells might possess distinct preferences. For instance, human Mo-DCs and human primary T cells require poly (A) strands to have an ideal length of 120-150 nucleotides and 300 nucleotides,respectively.74 Furthermore, by interacting with 5′cap via translational initiation factors like eIF4G and eIF4E, Poly (A) binding protein can generate a close-loop that affects mRNA architecture.77 Adding different nucleotides to mRNA transcripts is an additional method to increase mRNA stability, translation effectiveness, and mRNA vaccination strength.78,79 ...
Article
The messenger RNA (mRNA) vaccines have progressed from a theoretical concept to a clinical reality over the last few decades. Compared to conventional vaccination methods, these vaccines have a number of benefits, such as substantial potency, rapid growth, inexpensive production, and safe administration. Nevertheless, their usefulness was restricted up to now due to worries about the erratic and ineffective circulation of mRNA in vivo. Thankfully, these worries have largely been allayed by recent technological developments, which have led to the creation of multiple mRNA vaccination platforms for cancer and viral infections. The mRNA vaccines have been demonstrated as a powerful alternative to traditional conventional vaccines because of their high potency, safety and efficacy, capacity for rapid clinical development, and potential for rapid, low-cost manufacturing. The paper will examine the present status of mRNA vaccine technology and suggest future paths for the advancement and application of this exciting vaccine platform as a common therapeutic choice.
... Other design modifications are also applied. Encoding a 3′ poly(A) tail in the template adds 120-150 adenines, which facilitates initiation of translation and enhances the stability of the mRNA strand 43,44 . Equally important, modifications to include a 5′ N 7 -methylguanosine triphosphate structure on the 2′-O-methylated ribose of the first and sometimes second nucleotide constitute a cap that marks IVT mRNA as 'self' and abrogates recognition by interferon-induced protein with tetratricopeptide repeats 1 (IFIT1) 45,46 . ...
... Moreover, mRNA provides a promising platform to treat diseases caused by malfunctioning or missing proteins or enzymes 18 . In situ protein expression enables their subcellular transfer, which also improves therapeutic efficacy 43 . Natural mRNA decay and protein turnover rates limit the duration of their therapeutic effect 247,248 but their efficacy and longevity still outperform those of DNA or traditional protein-based therapies 249 . ...
Article
mRNA formulated with lipid nanoparticles is a transformative technology that has enabled the rapid development and administration of billions of coronavirus disease 2019 (COVID-19) vaccine doses worldwide. However, avoiding unacceptable toxicity with mRNA drugs and vaccines presents challenges. Lipid nanoparticle structural components, production methods, route of administration and proteins produced from complexed mRNAs all present toxicity concerns. Here, we discuss these concerns, specifically how cell tropism and tissue distribution of mRNA and lipid nanoparticles can lead to toxicity, and their possible reactogenicity. We focus on adverse events from mRNA applications for protein replacement and gene editing therapies as well as vaccines, tracing common biochemical and cellular pathways. The potential and limitations of existing models and tools used to screen for on-target efficacy and de-risk off-target toxicity, including in vivo and next-generation in vitro models, are also discussed.
... We have previously validated this approach as efficient, safe, non-integrative, and liver-targeted when the LNP were injected intravenously, enabling transient and robust protein expression in the murine liver 23,24 . mRNA-LNP safety in humans has been validated with current cancer immunotherapy 25 and mRNA-LNP-based COVID-19 vaccines 26 , as well as in many clinical trials aiming to treat infections, cancer, and genetic disorders 27 . The regenerative stimuli we deliver are two human hepatocyte mitogens-hepatocyte growth factor (HGF) and epidermal growth factor (EGF). ...
Article
Full-text available
Primary human hepatocyte (PHH) transplantation is a promising alternative to liver transplantation, whereby liver function could be restored by partial repopulation of the diseased organ with healthy cells. However, currently PHH engraftment efficiency is low and benefits are not maintained long-term. Here we refine two male mouse models of human chronic and acute liver diseases to recapitulate compromised hepatocyte proliferation observed in nearly all human liver diseases by overexpression of p21 in hepatocytes. In these clinically relevant contexts, we demonstrate that transient, yet robust expression of human hepatocyte growth factor and epidermal growth factor in the liver via nucleoside-modified mRNA in lipid nanoparticles, whose safety was validated with mRNA-based COVID-19 vaccines, drastically improves PHH engraftment, reduces disease burden, and improves overall liver function. This strategy may overcome the critical barriers to clinical translation of cell therapies with primary or stem cell-derived hepatocytes for the treatment of liver diseases.
... However, the use of the modified nucleic acid m1ψ can significantly reduce dsRNA production. Due to this creative innovation, Karikó and her colleague won the 2023 Nobel Prize in Physiology or Medicine [46][47][48][49]. mRNA vaccines accepted by the FDA replace uridine with m1ψTP as the raw material to limit the potential for immunogenicity [50][51][52]. ...
Article
Full-text available
: mRNA vaccines are entering a period of rapid development. However, their synthesis is still plagued by challenges related to mRNA impurities and fragments (incomplete mRNA). Most impurities of mRNA products transcribed in vitro are mRNA fragments. Only full-length mRNA transcripts containing both a 5′-cap and a 3′-poly(A) structure are viable for in vivo expression. Therefore, RNA fragments are the primary product-related impurities that significantly hinder mRNA efficacy and must be effectively controlled; these species are believed to originate from either mRNA hydrolysis or premature transcriptional termination. In the manufacturing of commercial mRNA vaccines, T7 RNA polymerase-catalyzed in vitro transcription (IVT) synthesis is a well-established method for synthesizing long RNA transcripts. This study identified a pivotal domain on the T7 RNA polymerase that is associated with erroneous mRNA release. By leveraging the advantageous properties of a T7 RNA polymerase mutant and precisely optimized IVT process parameters, we successfully achieved an mRNA integrity exceeding 91%, thereby further unlocking the immense potential of mRNA therapeutics.
... These attempts have relied on the use of a single protein or combinations of up to three proteins, such as VEGF, PDGF, TGFβ, ANGPT1, and other growth factors, to improve overall wound repair [37][38][39] . Interest has turned to gene therapy, since it may be possible to facilitate sustained production of angiogenic products by altering gene expression [40][41][42] . Shams et al. demonstrated that fibroblasts transfected with recombinant VEGF plasmid enhanced angiogenesis in vitro and in vivo. ...
Article
Full-text available
Hypoxia-Inducible Factor-1α (HIF-1α) has presented a new direction for ischemic preconditioning of surgical flaps to promote their survival. In a previous study, we demonstrated the effectiveness of HIF-1a DNA plasmids in this application. In this study, to avoid complications associated with plasmid use, we sought to express HIF-1α through mRNA transfection and determine its biological activity by measuring the upregulation of downstream angiogenic genes. We transfected six different HIF-1a mRNAs–one predominant, three variant, and two novel mutant isoforms–into primary human dermal fibroblasts using Lipofectamine, and assessed mRNA levels using RT-qPCR. At all time points examined after transfection (3, 6, and 10 h), the levels of HIF-1α transcript were significantly higher in all HIF-1α transfected cells relative to the control (all p < 0.05, unpaired Student’s T-test). Importantly, the expression of HIF-1α transcription response genes (VEGF, ANG-1, PGF, FLT1, and EDN1) was significantly higher in the cells transfected with all isoforms than with the control at six and/or ten hours post-transfection. All isoforms were transfected successfully into human fibroblast cells, resulting in the rapid upregulation of all five downstream angiogenic targets tested. These findings support the potential use of HIF-1α mRNA for protecting ischemic dermal flaps.
Article
Full-text available
Infection with human immunodeficiency virus type 1 (HIV-1) is characterized by dysfunction of HIV-1-specific T cells. To control the virus, antigen-loaded den-dritic cells (DCs) might be useful to boost and broaden HIV-specific T-cell responses. In the present study, monocyte-derived DCs from nontreated HIV-1-seropositive patients were electroporated with codon-optimized ("humanized") mRNA encoding consensus HxB-2 (hHXB-2) Gag protein. These DCs elicited a strong HIV-1 Gag-specific interferon-(IFN-) response by an HLA-A2-restricted CD8 T-cell line. Moreover, hHXB-2 gag mRNA-electropo-rated DCs also triggered IFN-secretion by autologous peripheral blood mononuclear cells (PBMCs), CD4 T cells, and CD8 T cells from all patients tested. Next, a novel strategy was developed using autologous virus sequences. Significant specific IFN-T-cell responses were induced in all patients tested by DCs electroporated with patients' autologous polymerase chain reaction (PCR)-amplified and in vitro-transcribed proviral and plasma viral mRNA encoding either Gag or Env. The stimulatory effect was seen on PBMCs, CD8 T cells, and CD4 T cells, demonstrating both major histocompatibility complex (MHC) class I and MHC class II antigen presentation. Moreover, a significant interleukin-2 (IL-2) T-cell response was induced by DCs elec-troporated with hHxB-2 or proviral gag mRNA. These findings open a major perspective for the development of patient-specific immunotherapy for HIV-1 dis
Article
Full-text available
Targeted genetic engineering of human pluripotent cells is a prerequisite for exploiting their full potential. Such genetic manipulations can be achieved using site-specific nucleases. Here we engineered transcription activator–like effector nucleases (TALENs) for five distinct genomic loci. At all loci tested we obtained human embryonic stem cell (ESC) and induced pluripotent stem cell (iPSC) clones carrying transgenic cassettes solely at the TALEN-specified location. Our data suggest that TALENs employing the specific architectures described here mediate site-specific genome modification in human pluripotent cells with similar efficiency and precision as do zinc-finger nucleases (ZFNs).
Article
Full-text available
More than 70% of the RNA synthesized by T7 RNA polymerase during run-off transcription in vitro can be incorrect products, up to twice as long as the expected transcripts. Transcriptions with model templates indicate that false transcription is mainly observed when the correct product cannot form stable secondary structures at the 3′-end. Therefore, the following hypothesis is tested: after leaving the DNA template, the polymerase can bind a transcript to the template site and the 3′-end of the transcript to the product site and extend it, if the 3′-end is not part of a stable secondary structure. Indeed, incubation of purified transcripts with the polymerase in transcription conditions triggers a 3′-end prolongation of the RNA. When two RNAs of different lengths are added to the transcription mix, both generate distinct and specific patterns of prolonged RNA products without any interference, demonstrating the self-coding nature of the prolongation process. Furthermore, sequencing of the high molecular weight transcripts demonstrates that their 5′-ends are precisely defined in sequence, whereas the 3′-ends contain size-variable extensions which show complementarity to the correct transcript. Surprisingly, a reduction of the UTP concentration to 0.2-1.0 mM in the presence of 3.5-4.0 mM of the other NTPs leads to faithful transcription and good yields, irrespective of the nucleotide composition of the template.
Article
Full-text available
Nucleic acid-based vaccines such as viral vectors, plasmid DNA, and mRNA are being developed as a means to address a number of unmet medical needs that current vaccine technologies have been unable to address. Here we describe a cationic nanoemulsion delivery system developed to deliver a self-amplifying mRNA vaccine. This non-viral delivery system is based on Novartis's proprietary adjuvant MF59, which has an established clinical safety profile and is well tolerated in children, adults and the elderly. We show that non-viral delivery of a 9 kb self-amplifying mRNA elicits potent immune responses in mice, rats, rabbits, and non-human primates comparable to a viral delivery technology, and demonstrate that, relatively low doses (75 µg) induce antibody and T-cell responses in primates. We also show the cationic nanoemulsion-delivered self-amplifying mRNA enhances the local immune environment through recruitment of immune cells similar to an MF59 adjuvanted sub-unit vaccine. Lastly, we show that the site of protein expression within the muscle and magnitude of protein expression is similar to a viral vector. Given the demonstration that self-amplifying mRNA delivered using a cationic nanoemulsion is well tolerated and immunogenic in a variety of animal models, we are optimistic about the prospects for this technology.Molecular Therapy (2014); doi:10.1038/mt.2014.133.
Article
Full-text available
Processing of viral double-stranded RNA (dsRNA) into small interfering RNAs (siRNAs) contributes directly to an antiviral effect in plants and invertebrates, which is amplified through the recruitment of RNA interference (RNAi). In mammals, viral dsRNAs are the substrate of the innate immune response and limit viral spread by impacting on cellular translation and cytokine production, as well as promoting cell death. Recent studies suggest that viral siRNAs also exert a direct antiviral activity in mammalian cells. Here, I review the current knowledge of dsRNA processing in mammalian cells and discuss the recent findings in light of the complex interplay between RNAi and dsRNA-driven innate immune responses toward the common goal of virus restriction.
Article
Full-text available
Double-stranded RNA (dsRNA) is arguably the most potent viral trigger of innate immune signaling. Its activity has been recognized for over 5 decades, first as a toxin, then as a central component of the interferon system, as an efficient activator of antiviral responses and an immunomodulator for therapeutic applications. Nucleic acid sensing is the main basis for antiviral defense systems throughout the diverse forms of life from bacteria to plants and animals. Pattern recognition receptors of the host defense system not only sense viral dsRNA as a pathogen-associated molecular pattern in infected cells, but also recognize circulating endogenous dsRNA, a nonmicrobial signal, as a danger-associated molecular pattern, often leading to autoimmunity. Despite the effects of extracellular viral and host dsRNA associated with infection and autoimmunity, respectively, the understanding of cellular mechanisms for its recognition and uptake has only been appreciated in recent years. This review presents an overview of this unique form of nucleic acid, addressing its roles in infection, autoimmunity, and host sensing mechanisms. The goal of this review is to highlight the novel findings with a focus on extracellular recognition and uptake by the cell.
Article
The turnover rates of beta and gamma globin messenger RNAs and of beta and gamma globin protein synthesis in human reticulocytes have been measured. Our goal was to determine whether beta globin mRNA is significantly more stable than gamma globin mRNA during the final stages of erythroid cell maturation. Such a result could explain the reported increase in the beta-gamma protein synthetic ratio during erythroid maturation. As determined by molecular hybridization and cell- free translation assays, the half-lives of both mRNAs are 20 to 29 hours in adult and neonatal reticulocytes. Protein synthetic capacity in intact cells decays with a half-life of six to eight hours, but beta protein synthesis declines at the same rate as gamma. Therefore, the changing ratio of fetal to adult hemoglobin synthesis during late erythroid maturation does not result from differences in mRNA turnover rates or changes in translation efficiencies. These data, coupled with those obtained with immature erythroid cells (Farquhar et al, Dev Biol 85: 403, 1981), suggest that, during erythroid maturation, the gamma- beta globin protein synthesis ratio declines because gamma gene transcription ceases earlier than beta gene transcription. Our results also indicate that the protein synthetic machinery, not the quantity of mRNA, becomes rate-limiting for globin production in cultured reticulocytes.
Article
We have previously shown correction of X-linked severe combined immunodeficiency [SCID-X1, also known as γ chain (γc) deficiency] in 9 out of 10 patients by retrovirus-mediated γc gene transfer into autologous CD34 bone marrow cells. However, almost 3 years after gene therapy, uncontrolled exponential clonal proliferation of mature T cells (with γδ+ or αβ+ T cell receptors) has occurred in the two youngest patients. Both patients' clones showed retrovirus vector integration in proximity to the LMO2 proto-oncogene promoter, leading to aberrant transcription and expression of LMO2. Thus, retrovirus vector insertion can trigger deregulated premalignant cell proliferation with unexpected frequency, most likely driven by retrovirus enhancer activity on the LMO2 gene promoter.