ArticlePDF AvailableLiterature Review

Understanding and tackling immune responses to AAV vectors

Authors:
  • Massachusetts Eye and Ear - Harvard Medical School

Abstract and Figures

As the clinical experience in Adeno-Associated Viral (AAV) vector-based gene therapies is expanding, the necessity to better understand and control the host immune responses is also increasing. Immunogenicity of AAV vectors in humans has been linked to several limitations of the platform including lack of efficacy due to antibody-mediated neutralization, tissue inflammation, loss of transgene expression, and in some cases complement activation and acute toxicities. Nevertheless, significant knowledge gaps remain in our understanding of the mechanisms of immune responses to AAV gene therapies, further hampered by the failure of preclinical animal models to recapitulate clinical findings. In this review, we focus on the current knowledge regarding immune responses, spanning from innate immunity to humoral and adaptive responses, triggered by AAV vectors and how they can be mitigated for safer, durable, and more effective gene therapies.
Content may be subject to copyright.
Open camera or QR reader and
scan code to access this article
and other resources online.
Understanding and Tackling Immune Responses
to Adeno-Associated Viral Vectors
Helena Costa-Verdera,
1
Carmen Unzu,
2
Erika Valeri,
1
Sahil Adriouch,
3
Gloria Gonza
´lez Aseguinolaza,
2,4
Federico Mingozzi,
5
and Anna Kajaste-Rudnitski
1,
*
1
San Raffaele Telethon Institute for Gene Therapy (SR-TIGET), IRCSS Ospedale San Raffaele, Milan, Italy;
2
DNA and RNA Medicine Division, CIMA, Universidad de
Navarra, IdisNA, Pamplona, Spain;
3
University of Rouen, INSERM, U1234, Pathophysiology Autoimmunity and Immunotherapy (PANTHER), Normandie University,
Rouen, France;
4
Vivet Therapeutics S.L., Pamplona, Spain; and
5
Spark Therapeutics, Inc., Philadelphia, Pennsylvania, USA.
As the clinical experience in adeno-associated viral (AAV) vector-based gene therapies is expanding, the necessity to
better understand and control the host immune responses is also increasing. Immunogenicity of AAV vectors in humans
has been linked to several limitations of the platform, including lack of efficacy due to antibody-mediated neutralization,
tissue inflammation, loss of transgene expression, and in some cases, complement activation and acute toxicities.
Nevertheless, significant knowledge gaps remain in our understanding of the mechanisms of immune responses to AAV
gene therapies, further hampered by the failure of preclinical animal models to recapitulate clinical findings. In this
review, we focus on the current knowledge regarding immune responses, spanning from innate immunity to humoral and
adaptive responses, triggered by AAV vectors and how they can be mitigated for safer, durable, and more effective gene
therapies.
Keywords: gene therapy, AAV, immune responses, innate immunity, adaptive immunity, viral restriction
INTRODUCTION
FROM BACTERI A TO vertebrates, life has established sophis-
ticated mechanisms to detect and eliminate foreign mole-
cules or to restrict its function and replication. In
mammalian cells, pattern recognition receptors (PRR) have
a central role as they recognize evolutionarily conserved
structures on pathogens, termed pathogen-associated mo-
lecular patterns within the specific compartments that they
patrol. Engagement of toll-like receptors (TLRs) or of the
cytosolic nucleic acid-detecting immune receptors such as
the RIG-I family of helicases (RIG-I, MDA5, LGP2),
cGAS, IFI16, or AIM2 will ultimately elicit type-I inter-
feron (IFN)-mediated antiviral responses.
1–3
Because all
current and emerging gene transfer and editing technolo-
gies are bound to expose cells to exogenous nucleic acids
and, most often, also to viral vectors, host antiviral factors
and nucleic acid sensors may play a pivotal role in the
efficacy and safety of gene therapy.
4
These cell-intrinsic innate immune responses may also
promote both humoral and cellular adaptive immune re-
sponses against components of the viral vector and, po-
tentially, its cargo. The relevance of immune responses in
the context of other delivery platforms such as integrating
Lentiviral Vectors or nonviral delivery systems such as
Lipid Nanoparticles and their nucleic acid cargos have
been extensively reviewed elsewhere.
4–8
For adeno-
associated viral (AAV) vectors, studies indicate that the
establishment of adaptive immune responses starts by
the recognition of the AAV vector capsid and genome by
the innate immune system.
9
In particular, the stimulation of different TLRs present
on antigen-presenting cells (APCs) has been shown to be
*Correspondence: Dr. Anna Kajaste-Rudnitski, San Raffaele Telethon Institute for Gene Therapy (SR-TIGET), IRCSS Ospedale San Raffaele, Via Olgettina 58, Milan 20132,
Italy. E-mail: kajaste.anna@hsr.it
836 jHUMAN GENE THERAPY, VOLUME 34, NUMBERS 17 and 18 DOI: 10.1089/hum.2023.119
ª2023 by Mary Ann Liebert, Inc.
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
involved in initial innate sensing of the vector and sub-
sequent establishment of adaptive responses, but other
PRRs and nonimmune cell types may also contribute to
this initial priming. As wild-type AAV naturally infects
humans, cross-reactive pre-existing immunity to AAV
vectors, both humoral and cell-mediated, is highly prev-
alent and can also interfere with gene transfer.
Immunogenicity of AAV vectors in humans has been
associated with several observations, including lack of
efficacy due to antibody-mediated neutralization, tissue
inflammation, and loss of transgene expression. In some
cases, acute toxicities have also been documented, trig-
gered by complement activation. Importantly, the lack of
preclinical animal models that robustly recapitulate find-
ings in patients further adds to the complexity of the un-
derstanding of the interaction between the immune system
and AAV vectors. This review will focus on the current
knowledge regarding immune responses, spanning from
innate immunity to humoral and cellular adaptive re-
sponses, triggered by AAV vectors, and how they can be
mitigated for safer, durable, and more effective gene
therapies.
INNATE IMMUNITY TO AAV
Capsid sensing
In primary human nonparenchymal liver cell cultures,
including Kupffer cells (KCs) and liver sinusoidal endo-
thelial cells, the viral capsid has been seen to contribute to
innate immunity mainly through binding to TLR2 ex-
pressed on the cell surface
10
(Fig. 1). TLR2 stimulation by
capsid proteins has been shown to signal through MyD88/
nuclear factor kappa B (NFjB) and transiently upregulate
proinflammatory cytokines such as interleukin (IL)-8, IL-
1b, tumor necrosis factor-a, or IL-6.
10
However, there is
no clear evidence so far of the role of TLR2 sensing in the
induction of adaptive immune responses.
11,12
Work by
Kuranda et al., explored innate responses stimulated by the
AAV2 capsid in the context of human peripheral blood
mononuclear cells (PBMCs), reporting that peripheral
monocyte-derived dendritic cells (moDCs) play a key role
in the establishment of humoral and cellular responses
through IL-1band IL-6 secretion.
13
Exposure of human PBMCs to either full AAV2, AAV2
empty particles, or capsid-derived peptide pools resulted
in increased IL-1band IL-6 levels in supernatants inde-
pendently of the serological status of the donors, with the
main source being moDCs compared to plasmacytoid
(pDCs) or conventional dendritic cells (cDCs). In subjects
previously exposed to wild-type AAV, these cytokines
triggered the differentiation of capsid-specific memory B
cells into antibody-secreting cells resulting in anti-capsid
antibody production, and this process could be inhibited
in vitro and in vivo by IL-6 and IL-1bblockade. On the
contrary, the AAV-seronegative individuals responded to
the AAV capsid by transient NK activation. The innate
receptors involved in the response were not directly as-
sessed in this study, and whether moDC activation results
from direct interaction with AAV peptides or via inter-
action with other APCs remains to be elucidated.
13
Vector genome sensing
As opposed to the TLR2-dependent sensing of the viral
capsid, sensing of the AAV vector genome through en-
dosomal TLR9 has been clearly linked with the subse-
quent activation of innate and adaptive responses against
the AAV capsid and the transgene product
11,14–16
(Fig. 1).
TLR9 recognizes unmethylated CpG sequences present in
viral or bacterial but not mammalian DNA. TLR9 signals
through the adapter molecule MyD88, which induces the
expression of proinflammatory cytokines through NFkB
activation, as well as type I IFNs via IRF3 and IRF7.
In 2009, Zhu et al. reported that exposure of murine
bone marrow-derived pDCs to single-stranded AAV
(ssAAV) led to the induction of type I IFN responses,
whereas no evidence of proinflammatory signaling was
detected in cDCs, macrophages, or KCs cultured in vitro.
14
Type I IFN signaling in pDCs was shown to be dependent
on the TLR9-MyD88 pathway and independent of the
nature of the transgene. Later studies have shown that
TLR9 stimulation and type I IFN responses mediated by
pDCs are important to promote cross-presentation of AAV
capsid antigens by cDCs to CD8
+
T cells, being pivotal for
anti-capsid cellular responses.
12
In vivo,Tlr9
-/-
, and
Myd88
-/-
mice injected intramuscularly with a ssAAV2
encoding the influenza virus hemagglutinin showed di-
minished T cell responses against both capsid and trans-
gene compared to wild-type controls.
14
Moreover, the authors reported that the anti-HA and
anti-capsid humoral responses were also significantly di-
minished in both models. Cytotoxicity was also absent in
Ifnr
-/-
mice together with reduced antibody responses to
HA and AAV, leading to long-term transgene expression,
highlighting a role for type I IFN in induction of adaptive
immunity in mice. Moreover, stimulation of PBMC-
derived human pDCs ex vivo with AAV2 encoding for
lacZ led to the upregulation of human IFN-a(hIFN-a) and
hIFN-bmRNA. The response was blocked by the TLR9
antagonist H154 ODN, suggesting a similar mechanism as
observed in mice.
A different study addressing innate immune responses
in the liver showed that proinflammatory responses to
ssAAV were mild in this tissue, whereas they were en-
hanced when using self-complementary AAV (scAAV)
vectors, correlating with stronger cellular and humoral
responses to the AAV capsid, but not to the transgene
product.
17
Innate responses to scAAV in the liver were
also dependent on TLR9 signaling and could be prevented
by transient inhibition of TLR9. Some inconsistent results
have been obtained regarding the role of TLR9 signaling
IMMUNE RESPONSES TO AAV GENE THERAPY 837
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
on humoral responses to the transgene product upon in-
tramuscular AAV injection. While Wu et al., reported
stronger transgene-specific cellular and humoral responses
induced by scAAV compared to ssAAV,
18
Rogers and
colleagues observed enhanced cellular responses but un-
changed antibody responses.
19
Regarding the different impact of modulating TLR9
signaling on the regulation of cellular and humoral re-
sponses, different roles have been attributed to TLR9 and
MyD88 in the establishment of adaptive responses to the
capsid and transgene product.
11
In intramuscular setting,
cellular but not humoral responses to transgene product
were reported to depend both on TLR9 and MyD88. As for
anti-AAV capsid responses, only humoral immunity
seemed to partially depend on MyD88, but not on TLR9,
indicating the involvement of additional sensing mecha-
nisms in shaping adaptive responses.
11
Altogether, dif-
ferent works evidence that TLR9 is not required for
antibody formation, yet, it may have a modifying effect on
anti-AAV IgG titers and subclasses.
Another factor shown to affect the magnitude of TLR9-
mediated signaling in addition to the vector DNA structure
is the content of CpG motifs in regions such as the inverted
terminal repeats (ITRs), the enhancer and promoter re-
gions, intronic sequences, and polyA tail.
16,20–22
Im-
portantly, it has been suggested that the unexpected loss of
FIX expression observed in patients administered the in-
vestigational product BAX335 was due to an increase in
CpG content on the expression cassette resulting from
codon-optimization.
21,23
In the setting of ex vivo gene editing, while AAV effi-
ciently avoids activation of type I IFN responses in human
hematopoietic stem and progenitor cells (HSPC), it trig-
gers p53-mediated DNA damage responses (DDRs) in this
Figure 1. Immunological barriers to AAV-mediated gene therapy. (1) Pre-existing NAbs bind to AAV capsids hampering successful liver transduction. (2 and
3) TLRs recognize vital capsid (TLR2) and viral genomes (TLR9) and trigger innate immune pathways. (4) dsRNA can induce RLRs and MDA5 sensors that will
trigger an IFN type I immune response. (5) Antigenic capsid or recombinant protein peptides generated by proteasome degradation are presented by APCs via
MHC class I to CD8
+
T cells and via MHC class II to CD4
+
T cells. After presentation, antigen-specific cytotoxic CD8
+
T cells can eliminate transduced cells and
CD4
+
T cells stimulate the activation of antibody producing plasma B cells. AAV, adeno-associated viral; APCs, antigen-presenting cells; dsDNA, double-
stranded DNA; dsRNA, double-stranded RNA; IFN, interferon; MHC, major histocompatibility complex; NAbs, neutralizing antibodies; RLRs, Rig-I like receptors;
ssDNA, single-stranded DNA; TLRs, toll-like receptors.
838 COSTA-VERDERA ET AL.
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
cell type.
24–26
AAV vector DNA genome is sensed in the
nucleus, where it triggers p53-mediated DDR upon re-
cruiting the MRE11-RAD50-NBS1 (MRN) complex on
the AAV ITRs.
24,25
AAV-mediated DDR induced apo-
ptotic responses in vitro and reduced engraftment of short-
term (ST)-HSC in vivo.
24,27
Complement activation
Recent clinical studies have shown that complement re-
sponses constitute another toxicity risk associated with high-
dose AAV gene transfer, causing variable clinical mani-
festations, including thrombotic microangiopathy (TMA),
kidney injury, thrombocytopenia, and cardiopulmonary in-
sufficiency.
28–32
Complement-related acute toxicities have
been observed only at vector doses above 1 ·10
13
vector
genomes (vg)/kg, and not consistently across trials.
Although complement factors have been shown to di-
rectly bind the AAV capsid ex vivo in the absence of anti-
capsid antibodies,
33,34
leading to increased AAV uptake
and innate responses in APCs,
33,35
the correlation between
pre-existing antibodies and the risk for complement acti-
vation is unclear. Additional studies have shown increased
complement factor binding and activation in the presence
of anti-AAV IgG,
34,35
in particular IgG1,
34
constituting a
risk in particular when dosing patients with pre-existing
humoral immunity. Nevertheless, complement responses
have been reported in clinical studies of gene therapy for
diseases, including Fabry disease, spinal muscular atrophy
(SMA), Methylmalonic acidemia,
28,31,36
and Duchenne
muscular dystrophy (DMD),
32
in which patients had been
prescreened for pre-existing anti-AAV antibody titers,
whereas no complement responses were reported in an
hemophilia B clinical trial, in which seropositive patients
were treated with AAV vectors.
37,38
Therefore, the risk for complement activation could be
influenced by the total vector dose administered and the
subclass of pre-existing or de novo formed anti-AAV an-
tibodies, the AAV serotype, or the underlying genetic
background of the host. Furthermore, while similar level
of complement activation has been observed to full and
empty AAV particles in the presence of immunoglobulins
in vitro,
35
a recent study presented at the 26th annual
meeting of American Society of Gene and Cell Therapy
(ASGCT) by Buchlis and colleagues suggests that the
vector genome could also be involved in the mechanism of
complement activation upon liver transduction in vivo.
39
Other mechanisms potentially contributing
to AAV innate immune activation
Among the potential alternative mechanisms contrib-
uting to AAV-induced innate immunity, one study re-
ported the induction of type I IFN responses by AAV
through the MDA5/MAVS axis, due to the detection of
double-stranded RNA generated from the intrinsic pro-
moter activity of the ITR regions (Fig. 1). However, the
impact of MAVS signaling on adaptive immune responses
has not been demonstrated.
40
Importantly, AAV gene transfer was also shown to in-
duce dose-dependent toxicity in sensory neurons of the
dorsal root ganglia (DRG) upon systemic and local AAV
administration in large animal models.
28,41,42
The mecha-
nism responsible for this toxicity remains still unclear since
evaluation of T cell responses and studies in the presence of
immunosuppressive drugs do not support a clear role of T
cell-mediated immunity,
41,43
but studies suggest a corre-
lation between overabundance of the transgene product and
neuron loss,
42
as reported also by Henry et al., at the 2023
ASGT meeting.
39
Ongoing works recently presented at the
ASGCT annual meeting have detected the presence of
immune cell infiltrates in spinal cord as well as different
cytokines in cerebrospinal fluid of injected nonhuman pri-
mates (NHPs), suggesting the involvement of innate im-
mune mechanisms which require further investigations.
39
Finally, one, although rare but lethal innate immune
response, was recently described as cytokine-mediated
capillary leak syndrome in a N-of-1 clinical trial for
DMD.
44
The patient experienced cardiac and pulmonary
toxicities associated to AAV9-mediated innate signaling,
which were complicated by the advanced DMD disease in
the patient, resulting in cardiorespiratory failure. Post-
injection studies in serum and pericardial fluid had re-
vealed elevation of proinflammatory cytokines and
complement factors before the cardiorespiratory arrest.
Restriction factors
The high AAV doses still required for therapeutic ef-
ficacy in some applications of gene transfer are a potential
risk factor contributing to the adverse toxicity observed in
recent clinical trials. The efficiency of viral vector-based
genetic manipulation may be limited by recognition of
exogenous components by host cell restriction factors
(RF).
4
RF are intrinsically present in different cell types
and usually part of the interferon stimulated genes, thus
further inducible upon type I IFN responses.
45
The antiviral mechanisms that interfere with AAV
transduction are not completely understood. Thus far, the
factors identified to block AAV transduction act at key
steps of the viral life cycle such as vector entry or DNA
genome conversion into dsDNA or target the AAV capsid
for proteasomal degradation (Fig. 2). A whole-genome
siRNA screen led to the identification of members of the
small ubiquitin-like modifier pathway as critical RF for
AAV transduction that acts at the level of vector entry
within the cells and are shared across different AAV se-
rotypes.
46
More recently, the VP2 capsid protein was
proposed as a direct target of SUMOylation.
47
Restriction
of AAV gene transduction can be mediated by direct
SUMOylation of the AAV capsid or by AAV-mediated
SUMOylation of cellular proteins, which then indirectly
influence vector transduction.
IMMUNE RESPONSES TO AAV GENE THERAPY 839
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
More recently, the apical polarity determinant Crumbs 3
(Crb3) protein emerged as a key RF playing a role in viral
attachment and entry from a CRISPR screen in a human
hepatic cell line.
48
FKBP52, a cellular chaperone protein, is
among the factors reported to act at the level of dsDNA
conversion. Phosphorylated FKBP52 binds AAV2 ITR
49,50
inhibiting viral second-strand DNA synthesis and affecting
transduction efficiency. The dephosphorylated protein lo-
ses the ability to bind AAV genome thus allowing efficient
transgene expression. PHF5A, a U2 snRNP-associated
protein, is an additional critical host factor identified
through a siRNA screening as able to limit AAV transgene
expression in a serotype and cell type-independent manner,
influencing a step after second-strand synthesis.
51
Disruption of PHF5, a subunit of the splicing factor 3b
protein complex that forms the U2 small nuclear ribonu-
cleoproteins complex (U2 snRNP) together with other
proteins, is sufficient to increase AAV transcript levels,
suggesting a critical role for the U2 snRNP spliceosome
complex in host-mediated restriction of AAV.
51
AAV proteasome degradation can also potentially af-
fect transduction efficiency. Different AAV capsid surface
residues are targets of cellular protein kinases.
52,53
Phos-
phorylated capsids become a substrate for ubiquitin con-
jugation and proteasome-dependent degradation
53,54
with
overall decrease in transduction efficiency.
55
AAV genome silencing is another barrier to efficient
long-term transgene expression. It has been shown that the
dsDNA binding protein NP220 and the human silencing
hub (HUSH) complex mediate transcriptional silencing of
single-stranded and self-complementary rAAV genomes
by epigenetic modification of associated host histones in a
Figure 2. Innate immune restriction mechanisms to efficient AAV transduction. Host RF have been shown to limit AAV transduction efficiency acting at
different steps of the AAV life cycle. The main mechanisms are reported in the figure: (1) inhibition of viral attachment and entry (Crb3); (2) capsid ubiquitination
and proteasome degradation; (3) capsid SUMOylation and/or AAV-mediated SUMOylation of putative host AAV RF; (4) viral ITR binding and inhibition of dsDNA
conversion (FKBP52); (5) AAV genome silencing by epigenetic modification of associated histones (NP220; HUSH complex). HUSH, human silencing hub; ITR,
inverted terminal repeat; RF, restriction factors.
840 COSTA-VERDERA ET AL.
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
serotype-dependent manner.
56
Lack of NP220 or HUSH
complex favors higher AAV transcript levels with reduced
H3K9 histone methylation marks.
56
Many other proteins involved in different cellular
processes such as cell cycle regulation, DDR, chromatin
remodeling, and transcriptional regulation, have also been
identified through cellular screenings as potential RF
limiting AAV transduction efficiency.
57,58
Silencing or
knock-out of these proteins results in improved transgene
expression from different AAV serotypes. The potential
benefits of inhibiting proteins such as the Fanconi anemia
protein FANCA, the HUSH-associated methyltransferase
SETDB1, or the nuclear matrix protein MORC3 have been
demonstrated in human primary cells, highlighting their
potential relevance in therapeutic settings.
58
CRISPR
screenings in cell lines were also used to identify factors
critical in facilitating AAV transduction. Among these
factors, knock-out of the GPR108 and TM9SF2 proteins
decreased AAV transduction using different serotypes.
59
Since GPR108 localizes to the Golgi, it has been suggested
that it may interact with AAV, playing a critical role in
viral escape or trafficking.
59
Strategies to overcome innate immune
barriers to AAV transduction
Although most strategies currently tested to prevent in-
nate immune responses to AAV are still under preclinical
research, several investigations have shown the influence of
the AAV vector genome composition on the induction of
proinflammatory signals through TLR9 and the magnitude
of cellular and humoral responses to both capsid and
transgene product, in particular, when using single-stranded
or self-complementary vectors.
11,17,18
In agreement, re-
ducing the CpG content
16,20–22
or including TLR9 inhibi-
tory sequences in cis
60
can dampen this cascade of immune
responses. CpG-depleted AAV vectors have shown re-
duced CD8
+
T cell responses to both capsid and the trans-
gene in liver and muscle gene transfer settings.
16,20,22
Yet, despite the amount of evidence regarding the link
between TLR9 stimulation and cytotoxic responses to the
capsid and transgene product, it seems that the beneficial
effect of avoiding TLR9 signaling is vector-dose depen-
dent, and responses are still induced above a certain
threshold
61,62
as reported also by Kumar et al., at the 26th
annual meeting of the ASGCT.
39
Although removing CpG
dimers from the transgene cDNA is beneficial in reducing
AAV immunogenicity, it is inherently impossible to
completely eliminate TLR9-mediated immune sensing.
As a result, even CpG depleted vectors stimulate TLR9
above a certain dose. Moreover, additional pathways have
been seen to contribute to induction of such responses,
such as the IL-1 receptor pathway reported by Kumar and
colleagues,
39,63,64
suggesting that multiple pathways may
need to be blocked at high vector doses for effective im-
mune modulation.
To prevent cell toxicity derived from DNA-damage
responses, inhibition of Ataxia-telangiectasia mutated ki-
nase, upstream of p53 activation, can be exploited to
prevent vector-induced DDRs in HSPC, rescuing delayed
engraftment.
24,27
Similarly, transient overexpression of
GSE56, a p53 dominant negative mRNA, is exploited in
AAV/Cas9 gene-editing procedure to rescue engraftment
defects and clonality of manipulated HSPC when coelec-
troporated with the genome-editing components.
25–27
The C5 inhibitor Eculizumab has been used to treat
complement-associated symptoms such as TMA in clini-
cal trials of SMA
36
and DMD,
32,65,66
however, further
studies are needed to determine the efficacy of this drug in
limiting complement responses. Other strategies have at-
tempted detargeting transgene expression from unwanted
tissues through the introduction of microRNA (miR) target
sites on the transgene sequence, such as the introduction of
targets for miR122
67
or miR183
42
to prevent transgene
overexpression and associated toxicity in liver and DRG,
respectively, as well as the use of miR142 targets to de-
target transgene expression in APCs and thus decrease
immune sensing of the transgene product.
67–69
The use of
miR targets to improve the tissue-specificity of transgene
expression is already being used in an ex vivo gene transfer
clinical trial for glioblastoma for the delivery of IFNa
specifically into the tumor microenvironment.
70
Strategies aimed at preventing AAV vector restriction
are also important to improve transduction efficiency and
decrease the vector doses needed to achieve therapeutic
benefit. For instance, targeting SUMOylation has been
reported to increase AAV transduction, constituting a
potential strategy to enhance AAV-based transgene de-
livery.
47
Proteasome inhibitors have also shown potential
to increase AAV transduction efficiency in different cell
types, both in vitro and in vivo,
71
as well as epigenetic
modulation of target cells to prevent vector silencing.
56
As the molecular understanding of the early innate
immune activation associated with AAV transduction in-
creases, it will be possible to design and test improved
vector configurations that avoid cellular sensors such as
TLRs or DNA recognizing proteins. Transduction proto-
cols adjusted to transiently prevent activation of these
pathways through targeted inhibitor delivery behold po-
tential to dampen innate immune activation and thereby
potentially diminish also subsequent humoral and adaptive
responses in vivo. While it is still unclear whether infec-
tious AAV encodes for antagonists of some of the iden-
tified RF, the field of virology offers plenty of examples of
how viruses effectively counteract these innate immune
barriers.
72,73
Further studies will help elucidate if AAV-
encoded proteins could be harnessed for such purposes or
if other viral factors or synthetic agonists could be
exploited to improve target cell permissivity to transduc-
tion, allowing the lowering of vector doses with conse-
quent lowering of the risks of adverse immune toxicity.
IMMUNE RESPONSES TO AAV GENE THERAPY 841
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
ADAPTIVE IMMUNITY TO AAV
Humoral responses
Humoral immunity directed against the AAV capsid
represents one of the most important barriers to successful
gene transfer. Depending on the route of vector adminis-
tration, the impact of anti-capsids neutralizing antibodies
(NAbs) can be major (Fig. 3A), with systemic vector ad-
ministration being the setting, in which exposure to anti-
bodies and vector neutralization is the highest. The
antibody titer and the total capsid dose administered
74
may
also contribute to the outcome of gene transfer in the
presence of NAbs. Studies in animal models and in hu-
mans indicate that even low titer NAb can majorly affect
the outcome of gene transfer.
37,75,76
For example, NAb
titers of just 1:5 were shown to block AAV transduction of
the liver at vector doses of 5 ·10
12
vg/kg in NHPs.
77
Similarly, in humans, in the context of hemophilia B
trials, antibody titers of 1:17 were shown to prevent liver
transduction at a dose of 2 ·10
12
vg/kg,
37
whereas titers of
just 1:1 can reduce transgene expression by about 50%.
75
Nevertheless, an important aspect to be considered is that
NAb titers may vary from laboratory to laboratory de-
pending on the specific design of the NAb assay, due to the
lack of a standardized protocol.
Humoral immunity to AAV can be categorized in pre-
existing and postdosing. Pre-existing antibodies to AAV
originate from the exposure to the wild-type virus, which
usually occurs early in life.
78
Depending on the AAV se-
rotype, up to *70% of humans can be seropositive, with
an average of 30–40% positivity for anti-AAV NAbs
across serotypes in adults (Fig. 3B), and lower ser-
oprevalence in young children.
78,79
IgGs binding to AAV
are the antibody isotype mostly found in humans, with
positivity for IgM detected in a minority of individuals.
80
IgG-binding titers correlate with NAb titers,
81
and in the
context of natural immunity to AAV tend to be low to
moderate. Upon AAV vector administration, early pro-
duction of IgM is commonly seen, which is rapidly fol-
lowed by production of high-titer IgG binding to AAV
82
and, consequently, NAbs.
Post-treatment anti-AAV antibodies tend to persist at
high titers for several years and can cross-react with dif-
ferent AAV serotypes.
83
While pre-existing antibodies to
AAV affect eligibility to receive gene transfer, postdosing
humoral immunity prevents vector readministration, if
needed.
Anti-AAV antibodies are mainly measured with two
assays, a neutralization assay which measures inhibition of
transduction of a cell line by a reporter AAV vector, and a
binding assay measuring antibodies binding to the AAV
capsid.
84,85
Various protocols have been developed for
both the binding and neutralizing assays, and both have
been used for the prescreening of subjects ahead of AAV
vector administration. While binding and neutralizing
antibody titers correlate, some discrepancies can be found,
particularly at lower titers, with subjects scoring sero-
negative in one assay and positive in the other. The choice
of a method to screen for antibodies to AAV has important
Figure 3. Influence of pre-existing immunity on AAV delivery efficacy. (A) Potential impact of anti-AAV NAbs depending on the route of administration. Green
dots: the impact of anti-AAV antibodies is minimal when the vector is delivered intraparenchymally, subretinally, or into the cerebrospinal fluid. Yellow dot:
delivery of AAV intravitreally, appears to expose the vector to some degree of neutralization by NAbs. Red dot: systemic delivery of AAV vectors is the most
challenging when it comes to potential for vector neutralization by anti-capsid antibodies. (B) Prevalence of anti-capsid IgG in humans across several AAV
serotypes (adapted from Boutin et al.
151
).
842 COSTA-VERDERA ET AL.
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
implications, including ease of assay setup, validation, and
throughput; ultimately the selection of the assay platform
needs to be defined in the context of both preclinical and
clinical studies and efforts to develop a companion diag-
nostic for an investigational gene therapy need to start
early in the development.
Several solutions have been proposed to address the
issue of anti-AAV antibodies. Broadly, they can be cate-
gorized as (1) methods to prevent antibody formation; (2)
methods to protect the vector from neutralization; and (3)
methods to remove or inactivate circulating NAbs. Anti-
body prevention is a potential effective strategy to allow
for vector readministration, unlikely to be useful in the
context of pre-existing immunity. In general, most strat-
egies work best in the context of low to moderate NAb
titers, like those found in the context of natural immunity
to the virus, while are less effective in the context to high
titer antibodies. Thus, redosing of AAV vectors will pos-
sibly require a combination of NAb reduction strategies.
Methods to prevent antibody formation
Prevention of antibody formation via immunomodula-
tory regimens has been proposed to allow for vector
readministration. Various combinations of T and B cell
targeting drugs have been tested to block antibody for-
mation postvector dosing. Some of these combinations
have been tested in the clinic.
86
Overall, the task has been
challenging, with preclinical data not easily scalable to
humans, possibly reflecting the higher immunogenicity of
AAV vectors in humans and the long-term persistence of
the AAV capsid antigen postvector dosing. While im-
munomodulation is almost universally used in AAV gene
transfer to block detrimental inflammatory responses, the
risk benefit of the approach in the context of vector read-
ministration requires careful evaluation.
The idea of tolerizing the host immune system against
the AAV capsid has been explored using various strate-
gies. For example, the coadministration of tolerogenic
nanoparticles containing sirolimus has shown promising
results in preclinical studies.
87
In a recent healthy volun-
teer clinical trial, empty AAV capsids were administered
together with tolerogenic nanoparticles.
88,89
The approach
was effective in reducing NAb titers in the ST, while an-
tibodies appeared to rebound at later times to levels
comparable to the control group, suggesting that repeated
antigen-nanoparticle (i.e., AAV capsid-nanoparticle) or
sirolimus-nanoparticle exposures may be required to allow
for a more persistent suppression of antibody formation.
Vector engineering and serotype switching. Several
methodologies can be used to engineer AAV capsids with
better tissue tropism or the ability to cross physical bar-
riers.
61
Similarly, the development of non-natural variants
and their selection against pooled human sera can help
identify novel AAV capsids with low seroprevalence.
90
In
the context of vector readministration, serotype switch has
been shown to be effective in preclinical models.
91
While
the effectiveness of the approach needs clinical validation,
it should be kept in mind that the use of a different capsid
for redosing purposes is highly resource intensive as in
many ways it is equivalent to developing a new investi-
gational gene therapy product.
Chemical modification of AAV vectors and the co-
purification of AAV with exosomes have been shown to
provide some resistance to Nabs.
92
Both approaches are
less effective in evading high titer antibodies, thus being
potentially useful only in the context of pre-existing hu-
moral immunity to AAV. Also physical isolation of a
target organ, like the liver, with saline flushing has been
shown to be effective against low to moderate NAb ti-
ters.
93
Ultimately the clinical feasibility of the approach
needs to be established.
As a broadly available and safe, clinically established
technology, the use of plasmapheresis/plasma exchange
has been shown to be effective in lowering antibody titers
to AAV in preclinical studies and in human samples in the
context of natural immunity to AAV.
94
While no data are
available in the context of gene therapy trials, based on the
data available, the strategy is likely to be effective in the
context of low anti-AAV NAb titers. Lowering of high-
titer NAbs would likely require several cycles of plasma
absorption/exchange, due to the rapid rebound post-
procedure linked to antibodies residing in the extravas-
cular space.
More recently, it has been shown that the application of
immune-adsorption procedure enables successful read-
ministration of AAV5 in NHPs
95
and AAV-specific col-
umns have been developed to remove AAV-specific Nabs
from circulation.
96
In addition, empty AAV capsid decoys
have been explored as a potential way to absorb pre-
existing antibodies during vector administration.
74
Nevertheless, as empty AAV capsids alone have been
suggested to trigger innate immune responses,
10
the im-
pact of such strategies on overall immune activation will
need to be carefully assessed.
IgG-cleaving endopeptidases are both in development
and approved drugs for kidney transplant and other
antibody-mediated diseases.
97
Following a single enzyme
infusion in human trials, a rapid cleavage of IgG was ob-
served. In the context of gene transfer, the approach has
been shown to be effective in allowing for the rapid in-
activation of NAbs, allowing for vector administration in
the context of pre-existing immunity to AAV and, poten-
tially, readministration.
98–100
While only preclinical data
are available to date, gene therapy trials set to start soon
will have to confirm the safety and efficacy of the
approach.
A key question is the efficacy of the approach at high
NAb titers, as F(ab’)2 fragments released upon IgG
cleavage retain some neutralization activity that can
potentially affect vector transduction. Related to that, the
IMMUNE RESPONSES TO AAV GENE THERAPY 843
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
performance of current antibody detection methods needs
to be assessed carefully in the context of this technology.
Several endopeptidase variants are in development, in-
cluding a novel IgG and IgM cleaving enzyme.
101
Inhibitors of the neonatal Fc receptor are approved
drugs for antibody-mediated diseases such as myasthenia
gravis.
102
The chronic use of this class of drugs is safe, and
human studies showed lowering of antibody titers by up to
70% of baseline. Early preclinical studies show that the
approach can be useful alone or in combination with IgG-
cleaving endopeptidases to address Nabs.
103
More studies
are likely needed to define the efficacy of the approach,
along or in combination with other strategies, at different
NAb titers.
As efforts to tackle the issue of anti-AAV NAbs con-
tinue, more strategies are being developed to address the
issue. With clinical trials set to start, the issue of pre-
existing immunity to AAV, mostly characterized by low
NAb titers, is likely to be addressed effectively with some
of the approaches outlined above. Vector readministration
will likely require the combination of approaches, for
example immunomodulation at the time of the first ad-
ministration to lower antibody production followed by IgG
removal, or reduction of IgG half-life followed by IgG
removal. Of note, successful AAV administration in the
presence of humoral immunity to the capsid requires only
a temporally limited window of time to allow for the
vector to reach the target tissue.
104
Nevertheless, risk
benefit evaluation and development of suitable biomarkers
to monitor residual neutralization activity will be key to
gather much needed clinical learnings on this critical
challenge of in vivo gene transfer.
T cell immunity
The human immune system has evolved powerful cell-
mediated adaptive mechanisms to fight against and pre-
vent viral recurrent infections, which hampers the efficacy
of viral vector-based gene therapies. As nonpathogenic
viruses, and originally not associated with T cell activation
in mouse studies, AAVs were put forward as ideal viral
vectors invisible to the adaptive T cell surveillance.
However, the host cellular immune responses to AAV
have been challenging to elucidate due to several factors:
(1) AAV vector immunogenicity depends on serotype,
expression cassette, route of administration, target tissue,
and dose
105
; (2) differences in T cell responses between
animal models and humans
106
; and (3) low sensitivity of
AAV-reactive lymphocyte detection assays in peripheral
blood as well as lack of correlation with humoral
response.
81,107
In fact, most of what is known about the T cell re-
sponses to AAV come from lessons learned from gene
therapy clinical trials. As a general mechanism, upon
AAV systemic administration, transduced APC can rec-
ognize and present capsid-derived antigens to lympho-
cytes. Antigen presentation via major histocompatibility
complex (MHC) class I will activate cytotoxic CD8
+
T
cells that will eliminate transduced cells, whereas AAV
antigen presentation via MHC class II will activate CD4
+
T helper cells. Helper cells will amplify and expand both
the cellular and humoral immune response (Fig. 1) and
increase the anti-AAV antibody secretion by B cells.
108
Several assays have been adapted to detect AAV-specific
T cell responses, such as lymphocyte proliferation assays
in response to the AAV capsid or capsid-derived pep-
tides,
109
IFN-cELISPOT,
110,111
or activated lymphocyte
detection by flow cytometry.
13,110
Preclinical data and natural AAV infections
in human population
Importantly, most of the AAV immunological studies
performed in mice have shown that, upon AAV adminis-
tration, mouse CD8
+
T cells do not proliferate in vivo when
re-exposed to AAV-derived antigens (capsid or trans-
gene).
112,113
Furthermore, mouse CD8
+
cells specific for
AAV-derived epitopes fail to eliminate AAV-transduced
cells.
114
Interestingly however, ex vivo expanded capsid-
specific CD8
+
T cells can kill AAV transducer murine and
human hepatocytes in vitro (and to some degree in vivo in
mice upon adoptive transfer).
115–117
Despite the similarity
of NHP to humans in many translational aspects, it has
been also reported that AAV-specific CD8
+
cytotoxic T
cells obtained from NHP do not destroy AAV-transduced
cells
77,118
despite most of them being effector CD8
+
rather
than memory T cells.
108
AAV-specific CD4
+
T cells can
also be detected in NHP and their prevalence rates are
close to 50% in the NHP-screened population.
108
Regarding the healthy human population exposed to
natural AAV infections, Li et al., reported detection of AAV
capsid-specific CD8
+
and/or CD4
+
T cells in 50% of the
screened peripheral blood samples.
108
Half of the responsive
cells were identified as CD8
+
memory cells, 25% as CD8
+
effector cells, and the other 25% as effector memory cells, as
conrmedforAAV1inhealthydonors.
81
Of note, different
distributions were observed in NHP as humans have twice
the proportion of CD8
+
memory cells than primates. These
comprehensive studies also revealed that human CD4
+
T
cells were mainly central memory cells.
108
Clinical trials and approved AAV therapies
The development of a cellular immune response to
AAV was first described after FIX expression was lost in
one subject enrolled in the high-dose group in the first
AAV-mediated liver-directed clinical trial.
37
In this
landmark study, the vector was administered through the
hepatic artery. Unexpectedly, a loss of FIX expression in
plasma was detected around 4 weeks after AAV dosing,
overlapping with an asymptomatic rise in liver transami-
nases. For the second hemophilia B gene therapy trial, the
AAV8 genome was changed to a self-complementary
844 COSTA-VERDERA ET AL.
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
form and administered systemically.
82,119
Four out of six
patients from the high-dose cohort (2 ·10
12
vg/kg)
showed the same transient increase of transaminases.
However, in this case, a short course of prednisolone
treatment was able to control the cytotoxic CD8
+
response
allowing for long-term FIX expression.
119
Notably, this
response was not observed at the lower vector doses tested
in the trial, which suggested that the immune response was
dose dependent. Data collected from a hemophilia A
clinical trial did not evidence a consistent increase in
capsid-specific T cells in peripheral blood,
120
although
few patients dosed with Roctavian (Valoctocogene rox-
aparvovec) showed sporadic T cell-positive responses by
IFNcELISPOT.
111
Across trials, however, increase in li-
ver enzymes has been consistently observed post-AAV
infusion, in some cases requiring the administration of
prolonged immunomodulatory regimens.
121
To date, no
transgene-specific CD8
+
responses have been reported
upon liver-directed AAV administration.
122
The route of administration is an important factor in
mounting a cell-mediated immune response. For instance,
the muscle has the potential to initiate inflammation as well
as promote infiltration of circulating immune cells, which
has been widely exploited for efficient host vaccina-
tion,
123,124
including AAV-based vaccines.
125,126
Never-
theless, data from gene therapy clinical trials with
intramuscular administration, including those for the
treatment of lipoprotein lipase (LPL) deficiency, alpha-1
antitrypsin (AAT) deficiency, and Pompe disease and Du-
chenne muscular dystrophy,
127–130
did not show a clearance
of transduced cells despite the detection of capsid-specific
T cells. This is in contrast with liver-directed clinical
studies. Furthermore, CD4
+
FoxP3 T cells were found both
in the LPL and the AAT deficiency trials, suggesting an
induction of tolerance in the administration site, preventing
the elimination of transduced muscle cells.
131,132
To treat inborn and acquired neurological diseases
AAV vectors are usually directly administered into the
central nervous system.
133–137
Because it is a local route of
administration into a site long considered im-
munopriviledged, many trials have not included T cell
response assays. Therefore, information in this regard is
limited. In some of these trials, the anti-AAV antibody
response in circulation was tested and an increase in NAbs
was detected. This suggests that vector leakage from the
injection site can trigger a peripheral B cell response and
potentially T cell responses as well. Noteworthy, in a recent
clinical trial for Tay-Sachs disease, mild T cells responses
were observed upon AAV8 delivery to the cisterna magna
and the thoracolumbar junction.
138
A different strategy was
used for the treatment of SMA. Zolgensma, an US Food and
Drug Administration (FDA) and European Medicines
Agency (EMA) approved AAV9-based vector, is adminis-
tered systemically together with prednisolone (https://www
.fda.gov/media/126109/download).
In the Clinical Review Document (https://www.fda
.gov/media/128116), it is noted that T cell responses to
AAV9 were observed in some cases, whereas the Euro-
pean Public Assessment Report includes that AAV9-
specific responses were detected in all patients by
ELISPOT (https://www.ema.europa.eu/en/documents/
assessment-report/zolgensma-epar-public-assessment-
report_en.pdf). However, none of the studies showed a
correlation between ELISPOT-positive results and drug
efficacy. Noteworthy, results from ocular genetic disease
clinical trials have shown that both intravitreal and sub-
retinal AAV administration can recruit T cells to the
eye,
139–141
in contrast to what was originally thought about
the immune privilege of the eye. Of note, however, lym-
phocyte infiltration does not always result in loss of ex-
pression.
140,141
Modulation and mitigation strategies
Thanks to these valuable clinical experiences, many
AAV-based clinical protocols include at least a corticoste-
roid regimen at the time of vector dosing, or in the event of a
rise in transaminases (https://doi.org/10.1182/blood-2019-
124091, https://www.fda.gov/media/164094/download,
https://www.fda.gov/media/126109/download). It should
be noted that disease-associated underlying liver patholo-
gies in conjunction with high AAV vector doses have been
associated with severe liver toxicities despite im-
munomodulation.
29
Additional forms of immunosuppres-
sion to control T cell-specific responses to the AAV capsid
include the use of cyclosporin, tacrolimus, or sirolimus/
rapamycin.
142
Most of these drugs are directed toward T
cell suppression, however, rapamycin has been shown to
have a beneficial effect on Treg cells and therefore capable
of inducing tolerance to the AAV capsid.
143–145
Recently, nanoparticles loaded with rapamycin have
shown a durable tolerogenic response to AAV vectors in
preclinical studies, even allowing for vector read-
ministration.
87,146
Other mitigation strategies focusing on
Treg cell modulation include the incorporation of IgG-
derived MHC class II epitopes into the therapeutic ex-
pression cassette to reduce AAV immunity
147
or the use of
AAV-specific CAR-Tregs and polyclonal Tregs to allow
for successful transgene expression in the presence of
capsid-specific T cell responses.
148
In the recent studies described above,
88,89
the specific
T cell response to AAV8 empty capsids (dose *2·10
12
capsid particles/kg) was studied in healthy volunteers, as
well as the potential of an immunotolerant treatment to
control it (ImmTOR) as reported by Gordon et al., at the
25th ASGCT meeting.
149
As in previous studies on nat-
ural AAV infections, patient-specific differences in T cell
induction were observed, together with asymptomatic
alanine transaminase increases. Notably, a major contri-
bution of CD4
+
T helper cells was reported. In addition, an
increase in IFNclevels was observed both in the group
IMMUNE RESPONSES TO AAV GENE THERAPY 845
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
administered with the empty capsids only as well as in the
cohort with ImmTOR. However, the increase of IFNcwas
delayed in the latter.
These results confirmed that cell-mediated responses in
patients can be variable and suggest empty capsid removal
from AAV preparations as a mitigation strategy to reduce
T cell immunity to AAV. Finally, capsid engineering and
alternate AAV serotype readministration have been
proved as other promising strategies to evade T cell
responses.
150
CONCLUSIONS
AAV vector-mediated gene transfer has shown great
promise in the clinic in many applications of the tech-
nology, resulting in several regulatory approvals of gene
therapy drugs based on the platform. Immunogenicity of
AAV has limited the success of in vivo gene therapy af-
fecting safety, efficacy, and durability of gene transfer and
resulting in significant variability in clinical trials.
Therefore, controlling host immune responses to AAV
vectors is key to ensure long-term gene therapy efficiency
and avoid undesired related toxicities. In this regard, in-
sight from preclinical animal models has been instru-
mental for gaining understanding of the mechanisms of
AAV immunogenicity, as exemplified, although not ex-
haustively, in Table 1. Nevertheless, there are non-
negligible differences between these models and the
human immune system prompting additional efforts to be
put in gathering more data from clinical trials, including
time-course studies of AAV immune responses, possibly
also tracking of tissue infiltrating lymphocytes.
In summary, the immunology studies conducted in both
preclinical models as well as in human trials over more
than two decades have enabled the field to accumulate a
significant body of knowledge that is being applied back to
improve the AAV-based gene therapy technology. We
now know that AAV vectors can be designed to engage
less with the innate immune system that manufacturing
should aim at reducing process-related impurities such as
empty AAV capsids. We also learned how to manage
immune responses in the clinic using various immuno-
modulatory regimens, and potential solutions to the issue
of NAbs to AAV will be tested in the clinic in the near
future. As the work continues to progress, more solutions
will be added to the toolbox of gene therapy researchers,
enabling the continuing success of this promising thera-
peutic modality.
Table 1. Immune responses against transgene product documented in preclinical studies of adeno-associated viral-mediated
gene transfer
Species Pathological Model Transgene Product Target Tissues Immune Responses Observed Ref
Mouse HSV-2 gB SM Cellular and humoral
152
Mouse HA SM Cellular
153
Mouse Ova SM Cellular and humoral
154–156
Mouse b-gal SM Cellular
157
Mouse HIV-1 gag SM Cellular and humoral
18
Mouse Cas9 Liver Cellular
158
Mouse (F8
-/-
) Hemophilia A hFVIII Liver Humoral
159–161
Mouse (F9
-/-
) Hemophilia B hFIX Liver Humoral
162
Mouse (a-scg
-/-
) LGMD a-sg SM Cellular
163
Mouse (mdx) DMD b-gal SM Humoral
164
Mouse (ASM
-/-
) Niemann-Pick disease hASM CNS Humoral
165
Mouse (GAA
-/-
) Pompe disease GAA Heart, diaphragm, spinal cord Humoral
166
Dog Hemophilia A FVIII SM Humoral
167
Dog Hemophilia B FIX SM Humoral
168
Dog Batten disease TPP1 CNS Humoral
169
Dog DMD b-gal SM Cellular and humoral
170
Dog DMD Cas9 SM Cellular and humoral
171
Dog MPS I IDUA CNS Humoral
172
Cat MPS I IDUA CNS Humoral
173
Cat LPL deficiency LPL
S447X
SM Humoral
174
NHP hAAT SM Humoral
175
NHP EPO SM Humoral
176
NHP GFP Liver Cellular
157
NHP UGT1A1 Liver Cellular
177
NHP Sulfamidase CNS Cellular
178
NHP FVIII SM Cellular and humoral
179
a-scg, a-sarcoglycan; ASM, acid sphingomyelinase; b-gal, b-galactosidase; CNS, central nervous system; DMD, Duchenne muscular dystrophy; EPO,
erythropoietin; GAA, acid-a-glucosidase; GFP, green fluorescent protein; HA, hemagglutinin; hAAT, human a1-antitrypsin; hFIX, human clotting Factor IX; hFVIII,
human clotting Factor VIII; HIV1, human immunodeficiency virus 1; HSV2-gB, herpes simplex virus type 2 glycoprotein B; IDUA, a-L-iduronidase; LGMD, limb-
girdle muscular dystrophy; LPL, lipoprotein lipase; MPS I, type 1 mucopolysaccharidosis; NHP, non-human primates; Ova, ovalbumin; SM, skeletal muscle;
TPP1, tripeptidyl peptidase 1; UGT1A1, uridine diphosphate glucuronosyl transferase family 1 member A1.
846 COSTA-VERDERA ET AL.
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
AUTHOR DISCLOSURE
E.V. and A.K.-R. are inventors on pending and issued
patents on lentiviral gene transfer filed by the Telethon
Foundation and the San Raffaele Scientific Institute. F.M.
is employee of Spark Therapeutics, a Roche company, and
equity holder of Roche. He is also an inventor in pending
and issued patents on the AAV vector platform and
methods to overcome immunity to AAV vectors. G.G.A. is
cofounder, shareholder, and employee of Vivet Ther-
apeutics and issued patents on the AAV vector platform.
FUNDING INFORMATION
This work was supported by grants from the European
Research Council (ERC-CoG 819815-ImmunoStem) and
the Telethon Foundation (TELE22-AK) to A.K.-R.
REFERENCES
1. Decout A, Katz JD, Venkatraman S, et al. The
cGAS-STING pathway as a therapeutic target in
inflammatory diseases. Nat Rev Immunol 2021;
21(9):548–569; doi: 10.1038/s41577-021-00524-z
2. Motwani M, Pesiridis S, Fitzgerald KA. DNA
sensing by the cGAS-STING pathway in health
and disease. Nat Rev Genet 2019;20(11):657–
674; doi: 10.1038/s41576-019-0151-1
3. Rehwinkel J, Gack MU. RIG-I-like receptors:
Their regulation and roles in RNA sensing. Nat
Rev Immunol 2020;20(9):537–551; doi: 10.1038/
s41577-020-0288-3
4. Piras F, Kajaste-Rudnitski A. Antiviral immunity
and nucleic acid sensing in haematopoietic stem
cell gene engineering. Gene Ther 2021;28(1–2):
16–28; doi: 10.1038/s41434-020-0175-3
5. Annoni A, Gregori S, Naldini L, et al. Modulation
of immune responses in lentiviral vector-
mediated gene transfer. Cell Immunol 2019;342:
103802; doi: 10.1016/j.cellimm.2018.04.012
6. de Braganca L, Ferguson GJ, Luis Santos J, et al.
Adverse immunological responses against non-
viral nanoparticle (NP) delivery systems in the
lung. J Immunotoxicol 2021;18(1):61–73; doi: 10
.1080/1547691X.2021.1902432
7. Dudek AM, Porteus MH. Answered and un-
answered questions in early-stage viral vector
transduction biology and innate primary cell
toxicity for ex-vivo gene editing. Front Immunol
2021;12:660302; doi: 10.3389/fimmu.2021
.660302
8. Verbeke R, Hogan MJ, Lore K, et al. Innate
immune mechanisms of mRNA vaccines. Im-
munity 2022;55(11):1993–2005; doi: 10.1016/j
.immuni.2022.10.014
9. Verdera HC, Kuranda K, Mingozzi F. AAV vector
immunogenicity in humans: A long journey to
successful gene transfer. Mol Ther 2020;28(3):
723–746; doi: 10.1016/j.ymthe.2019.12.010
10. Hosel M, Broxtermann M, Janicki H, et al. Toll-
like receptor 2-mediated innate immune re-
sponse in human nonparenchymal liver cells
toward adeno-associated viral vectors. Hepatol-
ogy 2012;55(1):287–297; doi: 10.1002/hep.24625
11. Rogers GL, Suzuki M, Zolotukhin I, et al. Unique
roles of TLR9- and MyD88-dependent and -in-
dependent pathways in adaptive immune re-
sponses to AAV-mediated gene transfer. J
Innate Immun 2015;7(3):302–314; doi: 10.1159/
000369273
12. Rogers GL, Shirley JL, Zolotukhin I, et al. Plas-
macytoid and conventional dendritic cells coop-
erate in crosspriming AAV capsid-specific CD8(+)
T cells. Blood 2017;129(24):3184–3195; doi: 10
.1182/blood-2016-11-751040
13. Kuranda K, Jean-Alphonse P, Leborgne C, et al.
Exposure to wild-type AAV drives distinct capsid
immunity profiles in humans. J Clin Invest 2018;
128(12):5267–5279; doi: 10.1172/JCI122372
14. Zhu J, Huang X, Yang Y. The TLR9-MyD88
pathway is critical for adaptive immune re-
sponses to adeno-associated virus gene therapy
vectors in mice. J Clin Invest 2009;119(8):2388–
2398; doi: 10.1172/JCI37607
15. Ashley SN, Somanathan S, Giles AR, et al. TLR9
signaling mediates adaptive immunity following
systemic AAV gene therapy. Cell Immunol 2019;
346:103997; doi: 10.1016/j.cellimm.2019.103997
16. Faust SM, Bell P, Cutler BJ, et al. CpG-depleted
adeno-associated virus vectors evade immune
detection. J Clin Invest 2013;123(7):2994–3001;
doi: 10.1172/JCI68205
17. Martino AT, Suzuki M, Markusic DM, et al. The
genome of self-complementary adeno-
associated viral vectors increases Toll-like re-
ceptor 9-dependent innate immune responses in
the liver. Blood 2011;117(24):6459–6468; doi: 10
.1182/blood-2010-10-314518
18. Wu T, Topfer K, Lin SW, et al. Self-
complementary AAVs induce more potent
transgene product-specific immune responses
compared to a single-stranded genome. Mol
Ther 2012;20(3):572–579; doi: 10.1038/mt.2011
.280
19. Rogers GL, Martino AT, Zolotukhin I, et al. Role
of the vector genome and underlying factor IX
mutation in immune responses to AAV gene
therapy for hemophilia B. J Transl Med 2014;12:
25; doi: 10.1186/1479-5876-12-25
20. Bertolini TB, Shirley JL, Zolotukhin I, et al. Effect
of CpG depletion of vector genome on CD8(+)T
cell responses in AAV gene therapy. Front Im-
munol 2021;12:672449; doi: 10.3389/fimmu.2021
.672449
21. Konkle BA, Walsh CE, Escobar MA, et al. BAX
335 hemophilia B gene therapy clinical trial re-
sults: Potential impact of CpG sequences on
gene expression. Blood 2021;137(6):763–774;
doi: 10.1182/blood.2019004625
22. Xiang Z, Kurupati RK, Li Y, et al. The Effect of
CpG sequences on capsid-specific CD8(+) T cell
responses to AAV vector gene transfer. Mol Ther
2020;28(3):771–783; doi: 10.1016/j.ymthe.2019
.11.014
23. Wright JF. Codon modification and PAMPs in
clinical AAV vectors: The tortoise or the hare?
Mol Ther 2020;28(3):701–703; doi: 10.1016/j
.ymthe.2020.01.026
24. Piras F, Riba M, Petrillo C, et al. Lentiviral vec-
tors escape innate sensing but trigger p53 in
human hematopoietic stem and progenitor cells.
EMBO Mol Med 2017;9(9):1198–1211; doi: 10
.15252/emmm.201707922
25. Ferrari S, Jacob A, Cesana D, et al. Choice of
template delivery mitigates the genotoxic risk
and adverse impact of editing in human hema-
topoietic stem cells. Cell Stem Cell 2022;29(10):
1428.e9–1444.e9; doi: 10.1016/j.stem.2022.09
.001
26. Ferrari S, Jacob A, Beretta S, et al. Efficient
gene editing of human long-term hematopoietic
stem cells validated by clonal tracking. Nat
Biotechnol 2020;38(11):1298–1308; doi: 10
.1038/s41587-020-0551-y
27. Schiroli G, Conti A, Ferrari S, et al. Precise gene
editing preserves hematopoietic stem cell func-
tion following transient p53-mediated DNA
damage response. Cell Stem Cell 2019;24(4):
551.e8–565.e8; doi: 10.1016/j.stem.2019.02.019
28. Day JW, Mendell JR, Mercuri E, et al. Clinical
trial and postmarketing safety of onasemnogene
abeparvovec therapy. Drug Saf 2021;44(10):
1109–1119; doi: 10.1007/s40264-021-01107-6
29. Chand D, Mohr F, McMillan H, et al. Hepato-
toxicity following administration of onasemno-
gene abeparvovec (AVXS-101) for the treatment
of spinal muscular atrophy. J Hepatol 2021;74(3):
560–566; doi: 10.1016/j.jhep.2020.11.001
IMMUNE RESPONSES TO AAV GENE THERAPY 847
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
30. Gavriilaki E, Anagnostopoulos A, Mastellos DC.
Complement in thrombotic microangiopathies:
Unraveling Ariadne’s thread into the labyrinth of
complement therapeutics. Front Immunol 2019;
10:337; doi: 10.3389/fimmu.2019.00337
31. Guillou J, de Pellegars A, Porcheret F, et al. Fatal
thrombotic microangiopathy case following
adeno-associated viral SMN gene therapy. Blood
Adv 2022;6(14):4266–4270; doi: 10.1182/
bloodadvances.2021006419
32. Mendell JR, Al-Zaidy SA, Rodino-Klapac LR,
et al. Current clinical applications of in vivo gene
therapy with AAVs. Mol Ther 2021;29(2):464–
488; doi: 10.1016/j.ymthe.2020.12.007
33. Zaiss AK, Cotter MJ, White LR, et al. Comple-
ment is an essential component of the immune
response to adeno-associated virus vectors. J
Virol 2008;82(6):2727–2740; doi: 10.1128/JVI
.01990-07
34. West C, Federspiel J, Rogers K, et al. Comple-
ment activation by AAV-neutralizing antibody
complexes. Hum Gene Ther 2023;34(11–12):554–
566; doi: 10.1089/hum.2023.018
35. Smith CJ, Ross N, Kamal A, et al. Pre-existing
humoral immunity and complement pathway
contribute to immunogenicity of adeno-
associated virus (AAV) vector in human blood.
Front Immunol 2022;13:999021; doi: 10.3389/
fimmu.2022.999021
36. Chand DH, Zaidman C, Arya K, et al. Thrombotic
microangiopathy following onasemnogene abe-
parvovec for spinal muscular atrophy: A case
series. J Pediatr 2021;231:265–268; doi: 10
.1016/j.jpeds.2020.11.054
37. Manno CS, Pierce GF, Arruda VR, et al. Suc-
cessful transduction of liver in hemophilia by
AAV-Factor IX and limitations imposed by the
host immune response. Nat Med 2006;12(3):
342–347; doi: 10.1038/nm1358
38. Majowicz A, Nijmeijer B, Lampen MH, et al.
Therapeutic hFIX activity achieved after single
AAV5-hFIX treatment in hemophilia B patients
and NHPs with pre-existing anti-AAV5 NABs.
Mol Ther Methods Clin Dev 2019;14:27–36; doi:
10.1016/j.omtm.2019.05.009
39. ASGCT. Presidential symposium and presenta-
tion of top abstracts. Mol Ther 2023;31(4):1–794;
doi: 10.1016/j.ymthe.2023.04.017
40. Shao W, Earley LF, Chai Z, et al. Double-
stranded RNA innate immune response activa-
tion from long-term adeno-associated virus
vector transduction. JCI Insight 2018;3(12):
e120474; doi: 10.1172/jci.insight.120474
41. Hinderer C, Katz N, Buza EL, et al. Severe toxicity
in nonhuman primates and piglets following
high-dose intravenous administration of an
adeno-associated virus vector expressing human
SMN. Hum Gene Ther 2018;29(3):285–298; doi:
10.1089/hum.2018.015
42. Hordeaux J, Buza EL, Jeffrey B, et al. MicroRNA-
mediated inhibition of transgene expression
reduces dorsal root ganglion toxicity by AAV
vectors in primates. Sci Transl Med 2020;
12(569):eaba9188; doi: 10.1126/scitranslmed
.aba9188
43. Hordeaux J, Buza EL, Dyer C, et al. Adeno-
associated virus-induced dorsal root ganglion
pathology. Hum Gene Ther 2020;31(15–16):808–
818; doi: 10.1089/hum.2020.167
44. Lek A, Wong B, Keeler A, et al. Unexpected
death of a duchenne muscular dystrophy patient
in an N-of-1 trial of rAAV9-delivered CRISPR-
transactivator. medRxiv 2023;
2023.05.16.23289881; doi: 10.1101/2023.05.16
.23289881
45. Lazear HM, Schoggins JW, Diamond MS. Shared
and distinct functions of type I and type III in-
terferons. Immunity 2019;50(4):907–923; doi: 10
.1016/j.immuni.2019.03.025
46. Holscher C, Sonntag F, Henrich K, et al. The
SUMOylation pathway restricts gene transduc-
tion by adeno-associated viruses. PLoS Pathog
2015;11(12):e1005281; doi: 10.1371/journal.ppat
.1005281
47. Chen Q, Njenga R, Leuchs B, et al. SUMOylation
targets adeno-associated virus capsids but
mainly restricts transduction by cellular mecha-
nisms. J Virol 2020;94(19):e00871-20; doi: 10
.1128/JVI.00871-20
48. Madigan VJ, Tyson TO, Yuziuk JA, et al. A
CRISPR screen identifies the cell polarity deter-
minant crumbs 3 as an adeno-associated virus
restriction factor in hepatocytes. J Virol 2019;
93(21):e00943-19; doi: 10.1128/JVI.00943-19
49. Qing K, Hansen J, Weigel-Kelley KA, et al.
Adeno-associated virus type 2-mediated gene
transfer: Role of cellular FKBP52 protein in
transgene expression. J Virol 2001;75(19):8968–
8976; doi: 10.1128/JVI.75.19.8968-8976.2001
50. Zhao W, Zhong L, Wu J, et al. Role of cellular
FKBP52 protein in intracellular trafficking of re-
combinant adeno-associated virus 2 vectors.
Virology 2006;353(2):283–293; doi: 10.1016/j
.virol.2006.04.042
51. Schreiber CA, Sakuma T, Izumiya Y, et al. An
siRNA screen identifies the U2 snRNP spliceo-
some as a host restriction factor for recombinant
adeno-associated viruses. PLoS Pathog 2015;
11(8):e1005082; doi: 10.1371/journal.ppat
.1005082
52. Buning H, Srivastava A. Capsid modifications for
targeting and improving the efficacy of AAV
vectors. Mol Ther Methods Clin Dev 2019;12:
248–265; doi: 10.1016/j.omtm.2019.01.008
53. Zhong L, Zhao W, Wu J, et al. A dual role of
EGFR protein tyrosine kinase signaling in ubi-
quitination of AAV2 capsids and viral second-
strand DNA synthesis. Mol Ther 2007;15(7):
1323–1330; doi: 10.1038/sj.mt.6300170
54. Yan Z, Zak R, Luxton GW, et al. Ubiquitination of
both adeno-associated virus type 2 and 5 capsid
proteins affects the transduction efficiency of
recombinant vectors. J Virol 2002;76(5):2043–
2053; doi: 10.1128/jvi.76.5.2043-2053.2002
55. Rossi A, Dupaty L, Aillot L, et al. Vector un-
coating limits adeno-associated viral vector-
mediated transduction of human dendritic cells
and vector immunogenicity. Sci Rep 2019;9(1):
3631; doi: 10.1038/s41598-019-40071-1
56. Das A, Vijayan M, Walton EM, et al. Epigenetic
silencing of recombinant adeno-associated virus
genomes by NP220 and the HUSH complex. J
Virol 2022;96(4):e0203921; doi: 10.1128/JVI
.02039-21
57. Mano M, Ippodrino R, Zentilin L, et al. Genome-
wide RNAi screening identifies host restriction
factors critical for in vivo AAV transduction. Proc
Natl Acad Sci U S A 2015;112(36):11276–11281;
doi: 10.1073/pnas.1503607112
58. Ngo AM, Puschnik AS. Genome-scale analysis of
cellular restriction factors that inhibit transgene
expression from adeno-associated virus vectors.
J Virol 2023;97(4):e0194822; doi: 10.1128/jvi
.01948-22
59. Meisen WH, Nejad ZB, Hardy M, et al. Pooled
screens identify GPR108 and TM9SF2 as host
cell factors critical for AAV transduction. Mol
Ther Methods Clin Dev 2020;17:601–611; doi: 10
.1016/j.omtm.2020.03.012
60. Chan YK, Wang SK, Chu CJ, et al. Engineering
adeno-associated viral vectors to evade innate
immune and inflammatory responses. Sci Transl
Med 2021;13(580):eabd3438; doi: 10.1126/
scitranslmed.abd3438
61. Li N, Bertolini T, Herzog RW. AAV vector dose
determines TLR9 dependence of CD8+T cell
response to transgene product. Blood 2020;136:
3; doi: 10.1182/blood-2020-141054
62. Chan YK, Wang S, Letizia A, et al. Reducing
AAV-Mediated Immune Responses and Pathol-
ogy in a Subretinal Pig Model by Engineering the
Vector Genome. Cell Press: Cambridge, MA,
USA; 2019.
63. Kumar S, Biswas M, Pineros AR, et al. Cross
priming of transgene product-specific CD8+T
cells in hepatic AAV gene transfer depends on
IL-1 receptor and XCR1+dendritic cells but not
TLR9. Blood 2020;136:2–3; doi: 10.1182/blood-
2020-142926
64. Kumar S, Moanaro B, Sreevani A, et al. TLR9-
independent CD8 T cell responses in hepatic
AAV gene transfer through IL-1R1-MyD88 sig-
naling. Immunity 2023; Sneak Peak; doi: 10
.2139/ssrn.4418617
65. Pfizer. Pfizer’s new phase 1b results of gene
therapy in ambulatory boys with duchenne
muscular dystrophy (DMD) support advancement
into pivotal phase 3 study [press release]. 2020.
Available from: https://www.pfizer.com/news/
press-release/press-release-detail/pfizers-new-
phase-1b-results-gene-therapy-ambulatory-boys
66. Solid. Solid Biosciences Announces FDA Lifts
Clinical Hold on IGNITE DMD Clinical Trial [press
848 COSTA-VERDERA ET AL.
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
release]. 2020. Available from: https://www
.solidbio.com/about/media/press-releases/solid-
biosciences-announces-fda-lifts-clinical-hold-on-
ignite-dmd-clinical-trial
67. Qiao C, Yuan Z, Li J, et al. Liver-specific
microRNA-122 target sequences incorporated in
AAV vectors efficiently inhibits transgene ex-
pression in the liver. Gene Ther 2011;18(4):403–
410; doi: 10.1038/gt.2010.157
68. Majowicz A, Maczuga P, Kwikkers KL, et al. Mir-
142-3p target sequences reduce transgene-
directed immunogenicity following intramuscular
adeno-associated virus 1 vector-mediated gene
delivery. J Gene Med 2013;15(6–7):219–232;
doi: 10.1002/jgm.2712
69. Xiao Y, Muhuri M, Li S, et al. Circumventing
cellular immunity by miR142-mediated regulation
sufficiently supports rAAV-delivered OVA ex-
pression without activating humoral immunity.
JCI Insight 2019;5(13):e99052; doi: 10.1172/jci
.insight.99052
70. Eoli M, Farina F, Gentner B, et al. CLRM-08
Targeting immune-payload to the glioblastoma
tumor microenvironment using a macrophage-
based treatment relying on autologous, geneti-
cally modified, hematopoietic stem cell-based
therapy: The TEM-GBM study (NCT03866109).
Neurooncol Adv 2022;4(Suppl. 1):i7; doi: 10
.1093/noajnl/vdac078.028
71. Mitchell AM, Samulski RJ. Mechanistic insights
into the enhancement of adeno-associated virus
transduction by proteasome inhibitors. J Virol
2013;87(23):13035–13041; doi: 10.1128/JVI
.01826-13
72. Kai Chan Y, Gack MU. Viral evasion of intra-
cellular DNA and RNA sensing. Nat Publish
Group 2016;14(6):360–373; doi: 10.1038/nrmicro
.2016.45
73. Simon V, Bloch N, Landau NR. Intrinsic host
restrictions to HIV-1 and mechanisms of viral
escape. Nat Immunol 2015;16(6):546–553; doi:
10.1038/ni.3156
74. Mingozzi F, Anguela XM, Pavani G, et al. Over-
coming preexisting humoral immunity to AAV
using capsid decoys. Sci Transl Med 2013;5(194):
194ra92; doi: 10.1126/scitranslmed.3005795
75. George LA, Sullivan SK, Giermasz A, et al. He-
mophilia B gene therapy with a high-specific-
activity factor IX variant. N Engl J Med 2017;
377(23):2215–2227; doi: 10.1056/NEJMoa
1708538
76. Jiang H, Lillicrap D, Patarroyo-White S, et al.
Multiyear therapeutic benefit of AAV serotypes
2, 6, and 8 delivering factor VIII to hemophilia A
mice and dogs. Blood 2006;108(1):107–115; doi:
10.1182/blood-2005-12-5115
77. Jiang H, Couto LB, Patarroyo-White S, et al.
Effects of transient immunosuppression on
adenoassociated, virus-mediated, liver-directed
gene transfer in rhesus macaques and implica-
tions for human gene therapy. Blood 2006;
108(10):3321–3328; doi: 10.1182/blood-2006-04-
017913
78. Li C, Narkbunnam N, Samulski RJ, et al. Neu-
tralizing antibodies against adeno-associated
virus examined prospectively in pediatric pa-
tients with hemophilia. Gene Ther 2012;19(3):
288–294; doi: 10.1038/gt.2011.90
79. Fitzpatrick Z, Leborgne C, Barbon E, et al. Influ-
ence of pre-existing anti-capsid neutralizing and
binding antibodies on AAV vector transduction.
Mol Ther Methods Clin Dev 2018;9:119–129;
doi: 10.1016/j.omtm.2018.02.003
80. Murphy SL, Li H, Mingozzi F, et al. Diverse IgG
subclass responses to adeno-associated virus
infection and vector administration. J Med Virol
2009;81(1):65–74; doi: 10.1002/jmv.21360
81. Veron P, Leborgne C, Monteilhet V, et al. Hu-
moral and cellular capsid-specific immune re-
sponses to adeno-associated virus type 1 in
randomized healthy donors. J Immunol 2012;
188(12):6418–6424; doi: 10.4049/jimmunol
.1200620
82. Nathwani AC, Tuddenham EG, Rangarajan S,
et al. Adenovirus-associated virus vector-
mediated gene transfer in hemophilia B. N Engl
J Med 2011;365(25):2357–2365; doi: 10.1056/
NEJMoa1108046
83. George LA, Ragni MV, Rasko JEJ, et al. Long-
term follow-up of the first in human intravascular
delivery of AAV for gene transfer: AAV2-hFIX16
for severe hemophilia B. Mol Ther 2020;28(9):
2073–2082; doi: 10.1016/j.ymthe.2020.06.001
84. Gorovits B, Azadeh M, Buchlis G, et al. Evalua-
tion of the humoral response to adeno-
associated virus-based gene therapy modalities
using total antibody assays. AAPS J 2021;23(6):
108; doi: 10.1208/s12248-021-00628-3
85. Gorovits B, Fiscella M, Havert M, et al. Re-
commendations for the development of cell-
based anti-viral vector neutralizing antibody
assays. AAPS J 2020;22(2):24; doi: 10.1208/
s12248-019-0403-1
86. Corti M, Elder M, Falk D, et al. B-cell depletion is
protective against anti-AAV capsid immune re-
sponse: A human subject case study. Mol Ther
Methods Clin Dev 2014;1:14033; doi: 10.1038/
mtm.2014.33
87. Meliani A, Boisgerault F, Hardet R, et al. Antigen-
selective modulation of AAV immunogenicity
with tolerogenic rapamycin nanoparticles enables
successful vector re-administration. Nat Com-
mun 2018;9(1):4098; doi: 10.1038/s41467-018-
06621-3
88. Gojanovich GS, Hasan MM, Kuszlewicz B, et al.
Using a systems biology approach to unravel the
Immunogenicity of AAV8 empty capsids in
healthy volunteers. In: Human Gene Therapy:
Edinburgh International Conference Centre;
2022.
89. Schroeder H, Johnston L, Clyde E, et al. Func-
tional assessment of T cell responses to AAV8
empty capsids in healthy volunteers. Mol Ther
2022;30(4):19–20.
90. Tse LV, Klinc KA, Madigan VJ, et al. Structure-
guided evolution of antigenically distinct adeno-
associated virus variants for immune evasion.
Proc Natl Acad Sci U S A 2017;114(24):E4812–
E4821; doi: 10.1073/pnas.1704766114
91. Majowicz A, Salas D, Zabaleta N, et al. Suc-
cessful repeated hepatic gene delivery in mice
and non-human primates achieved by sequential
administration of AAV5(ch) and AAV1. Mol Ther
2017;25(8):1831–1842; doi: 10.1016/j.ymthe
.2017.05.003
92. Hudry E, Martin C, Gandhi S, et al. Exosome-
associated AAV vector as a robust and conve-
nient neuroscience tool. Gene Ther 2016;23(4):
380–392; doi: 10.1038/gt.2016.11
93. Mimuro J, Mizukami H, Hishikawa S, et al.
Minimizing the inhibitory effect of neutralizing
antibody for efficient gene expression in the liver
with adeno-associated virus 8 vectors. Mol Ther
2013;21(2):318–323; doi: 10.1038/mt.2012.258
94. Monteilhet V, Saheb S, Boutin S, et al. A 10
patient case report on the impact of plasma-
pheresis upon neutralizing factors against
adeno-associated virus (AAV) types 1, 2, 6, and
8. Mol Ther 2011;19(11):2084–2091; doi: 10
.1038/mt.2011.108
95. Salas D, Kwikkers KL, Zabaleta N, et al. Im-
munoadsorption enables successful rAAV5-
mediated repeated hepatic gene delivery in
nonhuman primates. Blood Adv 2019;3(17):
2632–2641; doi: 10.1182/bloodadvances
.2019000380
96. Bertin B, Veron P, Leborgne C, et al. Capsid-
specific removal of circulating antibodies to
adeno-associated virus vectors. Sci Rep 2020;
10(1):864; doi: 10.1038/s41598-020-57893-z
97. Jordan SC, Lorant T, Choi J, et al. IgG endo-
peptidase in highly sensitized patients under-
going transplantation. N Engl J Med 2017;
377(5):442–453; doi: 10.1056/NEJMoa1612567
98. Leborgne C, Barbon E, Alexander JM, et al. IgG-
cleaving endopeptidase enables in vivo gene
therapy in the presence of anti-AAV neutralizing
antibodies. Nat Med 2020;26(7):1096–1101; doi:
10.1038/s41591-020-0911-7
99. Elmore ZC, Oh DK, Simon KE, et al. Rescuing
AAV gene transfer from neutralizing antibodies
with an IgG-degrading enzyme. JCI Insight 2020;
5(19):e139881; doi: 10.1172/jci.insight.139881
100. Ros-Ganan I, Hommel M, Trigueros-Motos L,
et al. Optimising the IgG-degrading enzyme
treatment regimen for enhanced adeno-
associated virus transduction in the presence of
neutralising antibodies. Clin Transl Immunology
2022;11(2):e1375; doi: 10.1002/cti2.1375
101. Smith T, Hull JA, Fusco RM, et al. Newly en-
gineered IgM and IgG cleaving enzymes for AAV
gene therapy. In: Molecular Therapy. Los An-
IMMUNE RESPONSES TO AAV GENE THERAPY 849
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
geles Convention Center: Los Angeles, CA, USA;
2023.
102. Howard JF, Jr., Bril V, Burns TM, et al. Rando-
mized phase 2 study of FcRn antagonist efgar-
tigimod in generalized myasthenia gravis.
Neurology 2019;92(23):e2661–e2673; doi: 10
.1212/WNL.0000000000007600
103. Horiuchi M, Hinderer CJ, Shankle HN, et al.
Neonatal Fc receptor inhibition enables adeno-
associated virus gene therapy despite pre-
existing humoral immunity. Hum Gene Ther
2023; doi: 10.1089/hum.2022.216
104. Murphy SL, Li H, Zhou S, et al. Prolonged sus-
ceptibility to antibody-mediated neutralization
for adeno-associated vectors targeted to the li-
ver. Mol Ther 2008;16(1):138–145; doi: 10.1038/
sj.mt.6300334
105. Ronzitti G, Gross DA, Mingozzi F. Human immune
responses to adeno-associated virus (AAV) vec-
tors. Front Immunol 2020;11:670; doi: 10.3389/
fimmu.2020.00670
106. Herzog RW. Immune responses to AAV capsid:
Are mice not humans after all? Mol Ther 2007;
15(4):649–650; doi: 10.1038/sj.mt.6300123
107. Mingozzi F, Maus MV, Hui DJ, et al. CD8(+)
T-cell responses to adeno-associated virus cap-
sid in humans. Nat Med 2007;13(4):419–422;
doi: 10.1038/nm1549
108. Li H, Lasaro MO, Jia B, et al. Capsid-specific
T-cell responses to natural infections with
adeno-associated viruses in humans differ from
those of nonhuman primates. Mol Ther 2011;
19(11):2021–2030; doi: 10.1038/mt.2011.81
109. Unzu C, Hervas-Stubbs S, Sampedro A, et al.
Transient and intensive pharmacological immu-
nosuppression fails to improve AAV-based liver
gene transfer in non-human primates. J Transl
Med 2012;10:122; doi: 10.1186/1479-5876-10-
122
110. Hui DJ, Edmonson SC, Podsakoff GM, et al. AAV
capsid CD8+T-cell epitopes are highly con-
served across AAV serotypes. Mol Ther Methods
Clin Dev 2015;2:15029; doi: 10.1038/mtm.2015
.29
111. Patton KS, Harrison MT, Long BR, et al. Mon-
itoring cell-mediated immune responses in AAV
gene therapy clinical trials using a validated IFN-
gamma ELISpot method. Mol Ther Methods Clin
Dev 2021;22:183–195; doi: 10.1016/j.omtm.2021
.05.012
112. Lin SW, Hensley SE, Tatsis N, et al. Recombinant
adeno-associated virus vectors induce function-
ally impaired transgene product-specific CD8+T
cells in mice. J Clin Invest 2007;117(12):3958–
3970; doi: 10.1172/JCI33138
113. Lin J, Zhi Y, Mays L, et al. Vaccines based on
novel adeno-associated virus vectors elicit ab-
errant CD8+T-cell responses in mice. J Virol
2007;81(21):11840–11849; doi: 10.1128/JVI
.01253-07
114. Wang L, Figueredo J, Calcedo R, et al. Cross-
presentation of adeno-associated virus serotype
2 capsids activates cytotoxic T cells but does not
render hepatocytes effective cytolytic targets.
Hum Gene Ther 2007;18(3):185–194; doi: 10
.1089/hum.2007.001
115. Martino AT, Basner-Tschakarjan E, Markusic
DM, et al. Engineered AAV vector minimizes
in vivo targeting of transduced hepatocytes by
capsid-specific CD8+T cells. Blood 2013;121(12):
2224–2233; doi: 10.1182/blood-2012-10-460733
116. Pien GC, Basner-Tschakarjan E, Hui DJ, et al.
Capsid antigen presentation flags human hepa-
tocytes for destruction after transduction by
adeno-associated viral vectors. J Clin Invest
2009;119(6):1688–1695; doi: 10.1172/JCI36891
117. Palaschak B, Marsic D, Herzog RW, et al. An
immune-competent murine model to study
elimination of AAV-transduced hepatocytes by
capsid-specific CD8(+) T cells. Mol Ther Methods
Clin Dev 2017;5:142–152; doi: 10.1016/j.omtm
.2017.04.004
118. Li C, Hirsch M, Asokan A, et al. Adeno-
associated virus type 2 (AAV2) capsid-specific
cytotoxic T lymphocytes eliminate only vector-
transduced cells coexpressing the AAV2 capsid
in vivo. J Virol 2007;81(14):7540–7547; doi: 10
.1128/JVI.00529-07
119. Nathwani AC, Reiss UM, Tuddenham EG, et al.
Long-term safety and efficacy of factor IX gene
therapy in hemophilia B. N Engl J Med 2014;
371(21):1994–2004; doi: 10.1056/
NEJMoa1407309
120. Rangarajan S, Walsh L, Lester W, et al. AAV5-
factor VIII gene transfer in severe hemophilia A.
N Engl J Med 2017;377(26):2519–2530; doi: 10
.1056/NEJMoa1708483
121. Ozelo MC, Mahlangu J, Pasi KJ, et al. Valocto-
cogene roxaparvovec gene therapy for hemo-
philia A. N Engl J Med 2022;386(11):1013–1025;
doi: 10.1056/NEJMoa2113708
122. Ertl HCJ. T cell-mediated immune responses to
AAV and AAV vectors. Front Immunol 2021;12:
666666; doi: 10.3389/fimmu.2021.666666
123. Liang F, Lore K. Local innate immune responses
in the vaccine adjuvant-injected muscle. Clin
Transl Immunology 2016;5(4):e74; doi: 10.1038/
cti.2016.19
124. Zuckerman JN. The importance of injecting
vaccines into muscle. Different patients need
different needle sizes. BMJ 2000;321(7271):
1237–1238; doi: 10.1136/bmj.321.7271.1237
125. Krotova K, Kuoch Yoshitomi H, Caine C, et al.
Tumor antigen-loaded AAV vaccine drives pro-
tective immunity in a melanoma animal model.
Mol Ther Methods Clin Dev 2023;28:301–311;
doi: 10.1016/j.omtm.2023.01.006
126. Zabaleta N, Dai W, Bhatt U, et al. An AAV-
based, room-temperature-stable, single-dose
COVID-19 vaccine provides durable immunoge-
nicity and protection in non-human primates. Cell
Host Microbe 2021;29(9):1437.e8–1453.e8; doi:
10.1016/j.chom.2021.08.002
127. Mingozzi F, Meulenberg JJ, Hui DJ, et al. AAV-1-
mediated gene transfer to skeletal muscle in
humans results in dose-dependent activation of
capsid-specific T cells. Blood 2009;114(10):2077–
2086; doi: 10.1182/blood-2008-07-167510
128. Brantly ML, Chulay JD, Wang L, et al. Sustained
transgene expression despite T lymphocyte re-
sponses in a clinical trial of rAAV1-AAT gene
therapy. Proc Natl Acad Sci U S A 2009;106(38):
16363–16368; doi: 10.1073/pnas.0904514106
129. Smith BK, Collins SW, Conlon TJ, et al. Phase I/II
trial of adeno-associated virus-mediated alpha-
glucosidase gene therapy to the diaphragm for
chronic respiratory failure in Pompe disease:
Initial safety and ventilatory outcomes. Hum
Gene Ther 2013;24(6):630–640; doi: 10.1089/
hum.2012.250
130. Bowles DE, McPhee SW, Li C, et al. Phase 1
gene therapy for Duchenne muscular dystrophy
using a translational optimized AAV vector. Mol
Ther 2012;20(2):443–455; doi: 10.1038/mt.2011
.237
131. Ferreira V, Petry H, Salmon F. Immune responses
to AAV-vectors, the glybera example from bench
to bedside. Front Immunol 2014;5:82; doi: 10
.3389/fimmu.2014.00082
132. Mueller C, Chulay JD, Trapnell BC, et al. Human
Treg responses allow sustained recombinant
adeno-associated virus-mediated transgene ex-
pression. J Clin Invest 2013;123(12):5310–5318;
doi: 10.1172/JCI70314
133. Bankiewicz KS, Forsayeth J, Eberling JL, et al.
Long-term clinical improvement in MPTP-
lesioned primates after gene therapy with AAV-
hAADC. Mol Ther 2006;14(4):564–570; doi: 10
.1016/j.ymthe.2006.05.005
134. Chien YH, Lee NC, Tseng SH, et al. Efficacy and
safety of AAV2 gene therapy in children with
aromatic L-amino acid decarboxylase deficiency:
An open-label, phase 1/2 trial. Lancet Child
Adolesc Health 2017;1(4):265–273; doi: 10.1016/
S2352-4642(17)30125-6
135. Marks WJ, Jr., Bartus RT, Siffert J, et al. Gene
delivery of AAV2-neurturin for Parkinson’s dis-
ease: A double-blind, randomised, controlled
trial. Lancet Neurol 2010;9(12):1164–1172; doi:
10.1016/S1474-4422(10)70254-4
136. Rosenberg JB, Kaplitt MG, De BP, et al.
AAVrh.10-mediated APOE2 central nervous sys-
tem gene therapy for APOE4-associated Alzhei-
mer’s disease. Hum Gene Ther Clin Dev 2018;
29(1):24–47; doi: 10.1089/humc.2017.231
137. Zerah M, Piguet F, Colle MA, et al. Intracerebral
gene therapy using AAVrh.10-hARSA recombi-
nant vector to treat patients with early-onset
forms of metachromatic leukodystrophy: Pre-
clinical feasibility and safety assessments in
nonhuman primates. Hum Gene Ther Clin Dev
850 COSTA-VERDERA ET AL.
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
2015;26(2):113–124; doi: 10.1089/humc.2014
.139
138. Flotte TR, Cataltepe O, Puri A, et al. AAV gene
therapy for Tay-Sachs disease. Nat Med 2022;
28(2):251–259; doi: 10.1038/s41591-021-01664-
4
139. Dimopoulos IS, Hoang SC, Radziwon A, et al.
Two-year results after AAV2-mediated gene
therapy for choroideremia: The alberta experi-
ence. Am J Ophthalmol 2018;193:130–142; doi:
10.1016/j.ajo.2018.06.011
140. Maguire AM, Russell S, Wellman JA, et al. Ef-
ficacy, safety, and durability of voretigene
neparvovec-rzyl in RPE65 mutation-associated
inherited retinal dystrophy: Results of phase 1
and 3 trials. Ophthalmology 2019;126(9):1273–
1285; doi: 10.1016/j.ophtha.2019.06.017
141. Maguire AM, High KA, Auricchio A, et al. Age-
dependent effects of RPE65 gene therapy for
Leber’s congenital amaurosis: A phase 1 dose-
escalation trial. Lancet 2009;374(9701):1597–
1605; doi: 10.1016/S0140-6736(09)61836-5
142. Schreiber SL, Crabtree GR. The mechanism of
action of cyclosporin A and FK506. Immunol
Today 1992;13(4):136–142; doi: 10.1016/0167-
5699(92)90111-J
143. Nayak S, Sarkar D, Perrin GQ, et al. Prevention
and reversal of antibody responses against fac-
tor IX in gene therapy for hemophilia B. Front
Microbiol 2011;2:244; doi: 10.3389/fmicb.2011
.00244
144. Fanigliulo D, Lazzerini PE, Capecchi PL, et al.
Clinically-relevant cyclosporin and rapamycin
concentrations enhance regulatory T cell func-
tion to a similar extent but with different
mechanisms: An in-vitro study in healthy hu-
mans. Int Immunopharmacol 2015;24(2):276–284;
doi: 10.1016/j.intimp.2014.12.021
145. Araki K, Turner AP, Shaffer VO, et al. mTOR
regulates memory CD8 T-cell differentiation.
Nature 2009;460(7251):108–112; doi: 10.1038/
nature08155
146. Weber ND, Odriozola L, Ros-Ganan I, et al.
Rescue of infant progressive familial intrahepatic
cholestasis type 3 mice by repeated dosing of
AAV gene therapy. JHEP Rep 2023;5(5):100713;
doi: 10.1016/j.jhepr.2023.100713
147. Hui DJ, Basner-Tschakarjan E, Chen Y, et al.
Modulation of CD8+T cell responses to AAV
vectors with IgG-derived MHC class II epitopes.
Mol Ther 2013;21(9):1727–1737; doi: 10.1038/mt
.2013.166
148. Arjomandnejad M, Sylvia K, Blackwood M, et al.
Modulating immune responses to AAV by ex-
panded polyclonal T-regs and capsid specific
chimeric antigen receptor T-regulatory cells. Mol
Ther Methods Clin Dev 2021;23:490–506; doi: 10
.1016/j.omtm.2021.10.010
149. ASGCT. 2022 ASGCT Annual Meeting Abstracts.
Mol Ther 2022;30(4, Supplement 1):1–592; doi:
10.1016/j.ymthe.2022.04.017
150. Barnes C, Scheideler O, Schaffer D. Engineering
the AAV capsid to evade immune responses.
Curr Opin Biotechnol 2019;60:99–103; doi: 10
.1016/j.copbio.2019.01.002
151. Boutin S, Monteilhet V, Veron P, et al. Pre-
valence of serum IgG and neutralizing factors
against adeno-associated virus (AAV) types 1, 2,
5, 6, 8, and 9 in the healthy population: Im-
plications for gene therapy using AAV vectors.
Hum Gene Ther 2010;21(6):704–712; doi: 10
.1089/hum.2009.182
152. Manning WC, Zhou S, Bland MP, et al. Transient
immunosuppression allows transgene expression
following readministration of adeno-associated
viral vectors. Hum Gene Ther 1998;9(4):477–485;
doi: 10.1089/hum.1998.9.4-477
153. Sarukhan A, Camugli S, Gjata B, et al. Successful
interference with cellular immune responses to
immunogenic proteins encoded by recombinant
viral vectors. J Virol 2001;75(1):269–277; doi: 10
.1128/JVI.75.1.269-277.2001
154. Hardet R, Chevalier B, Dupaty L, et al. Oral-
tolerization prevents immune responses and
improves transgene persistence following gene
transfer mediated by adeno-associated viral
vector. Mol Ther 2016;24(1):87–95; doi: 10.1038/
mt.2015.146
155. Wang L, Dobrzynski E, Schlachterman A, et al.
Systemic protein delivery by muscle-gene
transfer is limited by a local immune response.
Blood 2005;105(11):4226–4234; doi: 10.1182/
blood-2004-03-0848
156. Kumar SRP, Hoffman BE, Terhorst C, et al. The
balance between CD8(+) T cell-mediated clear-
ance of AAV-encoded antigen in the liver and
tolerance is dependent on the vector dose. Mol
Ther 2017;25(4):880–891; doi: 10.1016/j.ymthe
.2017.02.014
157. Gao G, Wang Q, Calcedo R, et al. Adeno-
associated virus-mediated gene transfer to
nonhuman primate liver can elicit destructive
transgene-specific T cell responses. Hum Gene
Ther 2009;20(9):930–942; doi: 10.1089/hum.2009
.060
158. Li A, Tanner MR, Lee CM, et al. AAV-CRISPR
gene editing is negated by pre-existing immunity
to Cas9. Mol Ther 2020;28(6):1432–1441; doi: 10
.1016/j.ymthe.2020.04.017
159. Bunting S, Zhang L, Xie L, et al. Gene therapy
with BMN 270 results in therapeutic levels of
FVIII in mice and primates and normalization of
bleeding in hemophilic mice. Mol Ther 2018;
26(2):496–509; doi: 10.1016/j.ymthe.2017.12.009
160. Sack BK, Wang X, Sherman A, et al. Immune
responses to human factor IX in haemophilia B
mice of different genetic backgrounds are dis-
tinct and modified by TLR4. Haemophilia 2015;
21(1):133–139; doi: 10.1111/hae.12522
161. Greig JA, Wang Q, Reicherter AL, et al. Char-
acterization of adeno-associated viral vector-
mediated human factor VIII gene therapy in
hemophilia A mice. Hum Gene Ther 2017;28(5):
392–402; doi: 10.1089/hum.2016.128
162. Nathwani AC, Davidoff A, Hanawa H, et al.
Factors influencing in vivo transduction by re-
combinant adeno-associated viral vectors ex-
pressing the human factor IX cDNA. Blood 2001;
97(5):1258–1265; doi: 10.1182/blood.v97.5.1258
163. Dressman D, Araishi K, Imamura M, et al. De-
livery of alpha- and beta-sarcoglycan by recom-
binant adeno-associated virus: Efficient rescue of
muscle, but differential toxicity. Hum Gene Ther
2002;13(13):1631–1646; doi: 10.1089/10430
340260201725
164. Yuasa K, Sakamoto M, Miyagoe-Suzuki Y, et al.
Adeno-associated virus vector-mediated gene
transfer into dystrophin-deficient skeletal mus-
cles evokes enhanced immune response against
the transgene product. Gene Ther 2002;9(23):
1576–1588; doi: 10.1038/sj.gt.3301829
165. Passini MA, Bu J, Fidler JA, et al. Combination
brain and systemic injections of AAV provide
maximal functional and survival benefits in the
Niemann-Pick mouse. Proc Natl Acad Sci U S A
2007;104(22):9505–9510; doi: 10.1073/pnas
.0703509104
166. Falk DJ, Soustek MS, Todd AG, et al. Com-
parative impact of AAV and enzyme replacement
therapy on respiratory and cardiac function in
adult Pompe mice. Mol Ther Methods Clin Dev
2015;2:15007; doi: 10.1038/mtm.2015.7
167. Sabatino DE, Lange AM, Altynova ES, et al. Ef-
ficacy and safety of long-term prophylaxis in
severe hemophilia A dogs following liver gene
therapy using AAV vectors. Mol Ther 2011;19(3):
442–449; doi: 10.1038/mt.2010.240
168. Arruda VR, Schuettrumpf J, Herzog RW, et al.
Safety and efficacy of factor IX gene transfer to
skeletal muscle in murine and canine hemophilia
B models by adeno-associated viral vector se-
rotype 1. Blood 2004;103(1):85–92; doi: 10.1182/
blood-2003-05-1446
169. Katz ML, Tecedor L, Chen Y, et al. AAV gene
transfer delays disease onset in a TPP1-deficient
canine model of the late infantile form of Batten
disease. Sci Transl Med 2015;7(313):313ra180;
doi: 10.1126/scitranslmed.aac6191
170. Yuasa K, Yoshimura M, Urasawa N, et al. In-
jection of a recombinant AAV serotype 2 into
canine skeletal muscles evokes strong immune
responses against transgene products. Gene
Ther 2007;14(17):1249–1260; doi: 10.1038/sj.gt
.3302984
171. Hakim CH, Kumar SRP, Perez-Lopez DO, et al.
Cas9-specific immune responses compromise
local and systemic AAV CRISPR therapy in mul-
tiple dystrophic canine models. Nat Commun
2021;12(1):6769; doi: 10.1038/s41467-021-
26830-7
172. Hinderer C, Bell P, Louboutin JP, et al. Neonatal
systemic AAV induces tolerance to CNS gene
therapy in MPS I dogs and nonhuman primates.
IMMUNE RESPONSES TO AAV GENE THERAPY 851
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
Mol Ther 2015;23(8):1298–1307; doi: 10.1038/mt
.2015.99
173. Hinderer C, Bell P, Gurda BL, et al. Intrathecal
gene therapy corrects CNS pathology in a feline
model of mucopolysaccharidosis I. Mol Ther
2014;22(12):2018–2027; doi: 10.1038/mt.2014
.135
174. Ross CJ, Twisk J, Bakker AC, et al. Correction of
feline lipoprotein lipase deficiency with adeno-
associated virus serotype 1-mediated gene
transfer of the lipoprotein lipase S447X benefi-
cial mutation. Hum Gene Ther 2006;17(5):487–
499; doi: 10.1089/hum.2006.17.487
175. Song S, Scott-Jorgensen M, Wang J, et al. In-
tramuscular administration of recombinant
adeno-associated virus 2 alpha-1 antitrypsin
(rAAV-SERPINA1) vectors in a nonhuman primate
model: Safety and immunologic aspects. Mol
Ther 2002;6(3):329–335; doi: 10.1006/mthe.2002
.0673
176. Gao G, Lebherz C, Weiner DJ, et al. Ery-
thropoietin gene therapy leads to autoimmune
anemia in macaques. Blood 2004;103(9):3300–
3302; doi: 10.1182/blood-2003-11-3852
177. Greig JA, Calcedo R, Kuri-Cervantes L, et al.
AAV8 gene therapy for Crigler-Najjar syndrome
in macaques elicited transgene T cell responses
that are resident to the liver. Mol Ther Methods
Clin Dev 2018;11:191–201; doi: 10.1016/j.omtm
.2018.10.012
178. Haurigot V, Marco S, Ribera A, et al. Whole body
correction of mucopolysaccharidosis IIIA by in-
tracerebrospinal fluid gene therapy. J Clin Invest
2013;123(8):3254–3271; doi: 10.1172/JCI66778
179. McIntosh J, Lenting PJ, Rosales C, et al. Ther-
apeutic levels of FVIII following a single pe-
ripheral vein administration of rAAV vector
encoding a novel human factor VIII variant. Blood
2013;121(17):3335–3344; doi: 10.1182/blood-
2012-10-462200
Received for publication July 21, 2023;
accepted after revision September 2, 2023.
Published online: September 4, 2023.
852 COSTA-VERDERA ET AL.
Downloaded by 190.129.165.51 from www.liebertpub.com at 01/12/24. For personal use only.
Preprint
Full-text available
Adeno-associated viral (AAV) vector-based gene therapy is gaining foothold as a treatment option for a variety of genetic neurological diseases with encouraging clinical results. Nonetheless, dose-dependent toxicities and severe adverse events have emerged in recent clinical trials through mechanisms that remain unclear. We have modelled here the impact of AAV transduction in the context of cell models of the human central nervous system (CNS), taking advantage of induced pluripotent stem cell-based technologies. Our work uncovers vector-induced cell-intrinsic innate immune mechanisms that contribute to apoptosis in 2D and 3D models. While empty AAV capsids were well tolerated, the AAV genome triggered p53-dependent DNA damage responses across CNS cell types followed by induction of IL-1R- and STING-dependent inflammatory responses. In addition, transgene expression led to MAVS-dependent signaling and activation of type I interferon (IFN) responses. Cell-intrinsic and paracrine apoptosis onset could be prevented by inhibiting p53 or acting downstream of STING- and IL-1R-mediated responses. Activation of DNA damage, type I IFN and CNS inflammation were confirmed in vivo , in a mouse model. Together, our work identifies the cell-autonomous innate immune mechanisms of vector DNA sensing that can potentially contribute to AAV-associated neurotoxicity.
Article
Gene therapy clinical trials are rapidly expanding for inherited metabolic liver diseases whilst two gene therapy products have now been approved for liver based monogenic disorders. Liver‐directed gene therapy has recently become an option for treatment of haemophilias and is likely to become one of the favoured therapeutic strategies for inherited metabolic liver diseases in the near future. In this review, we present the different gene therapy vectors and strategies for liver‐targeting, including gene editing. We highlight the current development of viral and nonviral gene therapy for a number of inherited metabolic liver diseases including urea cycle defects, organic acidaemias, Crigler–Najjar disease, Wilson disease, glycogen storage disease Type Ia, phenylketonuria and maple syrup urine disease. We describe the main limitations and open questions for further gene therapy development: immunogenicity, inflammatory response, genotoxicity, gene therapy administration in a fibrotic liver. The follow‐up of a constantly growing number of gene therapy treated patients allows better understanding of its benefits and limitations and provides strategies to design safer and more efficacious treatments. Undoubtedly, liver‐targeting gene therapy offers a promising avenue for innovative therapies with an unprecedented potential to address the unmet needs of patients suffering from inherited metabolic diseases.
Preprint
Full-text available
An N-of-1 trial was developed to deliver a dCas9-VP64 transgene designed to upregulate the cortical dystrophin as a custom therapy for a Duchenne muscular dystrophy (DMD) patient. After showing signs of mild cardiac dysfunction and pericardial effusion, the patient acutely decompensated and sustained cardiac arrest six-days after dosing and succumbed two-days later. Post-mortem examination revealed severe acute-respiratory distress syndrome with diffuse alveolar damage. Vector biodistribution data was obtained and revealed minimal expression of transgene in liver. There was no evidence of AAV9 antibodies nor of effector T cell reactivity. These findings demonstrate innate immune signaling with capillary leak as a form of toxicity in an advanced DMD case treated with high-dose rAAV gene therapy.
Article
Full-text available
Treatment of monogenetic disorders using adeno-associated viral vectors (AAV) is an area of intense interest. AAV is a human pathogen and pre-existing capsid antibodies are prevalent in the population posing a challenge to safety and efficacy of AAV-mediated gene therapies. Here we investigated the risk of AAV-mediated complement activation when sera from a cohort of human donors was exposed to AAV9 capsid. Seropositive donor sera carrying neutralizing antibodies from a previous environmental exposure activated complement when admixed with AAV9 capsids and complement-activation was associated with donors who had higher levels of ant-AAV IgG1 antibodies. These findings were consistent with Mass spectrometry analysis that identified increased binding of immunoglobulins and complement factors when AAV9 capsids were admixed with seropositive sera. Finally, complement activation was abrogated after IgG-depletion using affinity columns or serum pre-treatment with an IgG degrading enzyme. Overall, these results demonstrate an important role of pre-existing neutralizing antibodies in activating complement; a risk that can be mitigated by employing adequate immunosuppression strategies when dosing seropositive patients with vector.
Article
Full-text available
Adeno-associated virus (AAV) vectors are one of the leading platforms for gene delivery for the treatment of human genetic diseases, but the antiviral cellular mechanisms that interfere with optimal transgene expression are incompletely understood. Here, we performed two genome-scale CRISPR screens to identify cellular factors that restrict transgene expression from recombinant AAV vectors. Our screens revealed several components linked to DNA damage response, chromatin remodeling, and transcriptional regulation. Inactivation of the Fanconi anemia gene FANCA; the human silencing hub (HUSH)-associated methyltransferase SETDB1; and the gyrase, Hsp90, histidine kinase, and MutL (GHKL)-type ATPase MORC3 led to increased transgene expression. Moreover, SETDB1 and MORC3 knockout improved transgene levels of several AAV serotypes as well as other viral vectors, such as lentivirus and adenovirus. Finally, we demonstrated that the inhibition of FANCA, SETDB1, or MORC3 also enhanced transgene expression in human primary cells, suggesting that they could be physiologically relevant pathways that restrict AAV transgene levels in therapeutic settings. IMPORTANCE Recombinant AAV (rAAV) vectors have been successfully developed for the treatment of genetic diseases. The therapeutic strategy often involves the replacement of a defective gene by the expression of a functional copy from the rAAV vector genome. However, cells possess antiviral mechanisms that recognize and silence foreign DNA elements thereby limiting transgene expression and its therapeutic effect. Here, we utilize a functional genomics approach to uncover a comprehensive set of cellular restriction factors that inhibit rAAV-based transgene expression. Genetic inactivation of selected restriction factors increased rAAV transgene expression. Hence, modulation of identified restriction factors has the potential to enhance AAV gene replacement therapies.
Article
Full-text available
We previously described therapeutic opportunities provided by capsid- and expression cassette-optimized adeno-associated virus serotype 6 (AAV6) vectors to suppress tumor growth in both solid and metastatic mouse models by using artificial ovalbumin (OVA) immunogen. In the current study, we further elucidated the mechanism of function of a novel AAV-based vaccine loaded with the melanoma tumor-associated antigens premelanosome protein gp100, tyrosinase (Tyr), tyrosinase-related protein 1 (TRP1), and dopachrome tautomerase (TRP2). We showed that the AAV6-based vaccine creates cellular and humoral antigen-specific responses, while antigen expression at the site of vaccine injection was temporal, and the clearance of antigen coincided with T cell infiltration. Our data revealed the superior protective immune response of optimized AAV6-TRP1 compared with other self-antigens in a disease-free mouse model. We further assessed the ability of AAV6-TRP1 to protect animals from metastatic spread in the lungs and to extend animal survival by inhibiting solid tumor growth. Flow cytometry-based analysis indicated significant infiltration of CD8+ T cells and natural killer (NK) cells in the tumor site, as well as changes in the polarization of intratumoral macrophages. Altogether, our data strongly support the use of optimized AAV vectors for cancer vaccine development.
Article
Full-text available
Long-range gene editing by homology-directed repair (HDR) in hematopoietic stem/progenitor cells (HSPCs) often relies on viral transduction with recombinant adeno-associated viral vector (AAV) for template delivery. Here, we uncover unexpected load and prolonged persistence of AAV genomes and their fragments, which trigger sustained p53-mediated DNA damage response (DDR) upon recruiting the MRE11-RAD50-NBS1 (MRN) complex on the AAV inverted terminal repeats (ITRs). Accrual of viral DNA in cell-cycle-arrested HSPCs led to its frequent integration, predominantly in the form of transcriptionally competent ITRs, at nuclease on- and off-target sites. Optimized delivery of integrase-defective lentiviral vector (IDLV) induced lower DNA load and less persistent DDR, improving clonogenic capacity and editing efficiency in long-term repopulating HSPCs. Because insertions of viral DNA fragments are less frequent with IDLV, its choice for template delivery mitigates the adverse impact and genotoxic burden of HDR editing and should facilitate its clinical translation in HSPC gene therapy.
Article
Recombinant Adeno Associated Viruses (AAV) are a commonly used gene delivery viral vector system to effectively introduce foreign genes into a broad range of cells and tissues. These gene delivery tools are widely explored in novel strategies for treating human diseases associated with blood, central nervous system and neuromuscular disorders or cancer. In vivo gene therapies approved in the US and Europe based on this approach have yet indicated minimal immunogenicity, minimal genome integration, and no pathogenicity. Previous exposure to naturally occurring AAVs results in pre-existing immunity that limits the efficacy of gene therapy and represents a potential safety concern. Therefore, it is important to screen for pre- existing AAV antibodies prior to gene transfer therapies. For most AAV gene therapy candidates currently in development neutralization assays are used to determine anti AAV antibody levels. These cell- based assays pose challenges such as lack of reproducibility, time consumption, and limitations in screening large sample numbers. An ELISA based method is an alternative that can be developed to measure total antibody levels against a specific serotype. In general, these types of in vitro assays are sensitive, reproducible, rapid, and can be validated to provide precise and accurate measurements even at high throughput demands. Therefore, we have developed an ELISA based anti-AAV antibody detection method using the automated GyroLab immunoassay platform that requires minimum sample volume while ensuring precision and accuracy of measurements. In this method, AAV capsid serotypes are immobilized on streptavidin- coated bead beds within the microcolumn of a GyroLab BioaffyTM CD using AAV specific biotinylated ligands. The ratio of biotinylated ligands to form each specific complex is optimized. Primary antibodies against specific serotypes are captured by these immobilized AAV particles and detected by an Alexa Fluor 647-labeled secondary IgG antibody of the intended testing species. The method was qualified to detect anti-AAV8 antibodies between 7.8 ng/mL and 2000 ng/mL, anti-AAV9 antibodies between 16 ng/mL and 10000 ng/mL, and anti- AAV2 antibodies between 90 ng/mL and 25000 ng/mL, respectively. Anti-AAV antibody levels can be determined with ±20% accuracy of the nominal concentration and with ≤20% precision (%CV). These GyroLab immunoassays are capable of detecting and quantifying AAV antibodies in different matrices from multiple species such as human, mouse and non-human primates. Assay performance was determined using AAV positive and negative non-human primate sera obtained from a commercial vendor. Both calibrator standards and quality control (QC) samples are used for sample data acceptance within a single run. This established in vitro method can be utilized in preclinical studies to detect and quantify both pre-existing AAV antibodies as well as potential immunogenicity due to an AAV based gene therapy.
Article
Background & aims: Gene therapy using recombinant adeno-associated virus (rAAV) vector carrying multidrug resistance protein 3 (MDR3) coding sequence (AAV8-MDR3) represents a potential curative treatment for progressive familial intrahepatic cholestasis type 3 (PFIC3), which presents in early childhood. However, patients with the severest form of PFIC3 should receive treatment early after detection to prevent irreversible hepatic fibrosis leading ultimately to liver transplantation or death. This represents a challenge for rAAV-based gene therapy because therapeutic efficacy is expected to wane as rAAV genomes are lost owing to hepatocyte division, and the formation of AAV-specific neutralising antibodies precludes re-administration. Here, we tested a strategy of vector re-administration in infant PFIC3 mice with careful evaluation of its oncogenicity - a particular concern surrounding rAAV treatment. Methods: AAV8-MDR3 was re-administered to infant Abcb4 -/- mice 2 weeks after a first dose co-administered with tolerogenic nanoparticles carrying rapamycin (ImmTOR) given at 2 weeks of age. Eight months later, long-term therapeutic efficacy and safety were assessed with special attention paid to the potential oncogenicity of rAAV treatment. Results: Co-administration with ImmTOR mitigated the formation of rAAV-specific neutralising antibodies and enabled an efficacious second administration of AAV8-MDR3, resulting in stable correction of the disease phenotype, including a restoration of bile phospholipid content and healthy liver function, as well as the prevention of liver fibrosis, hepatosplenomegaly, and gallstones. Furthermore, efficacious repeat rAAV administration prevented the appearance of liver malignancies in an animal model highly prone to developing hepatocellular carcinoma. Conclusions: These outcomes provide strong evidence for rAAV redosing through co-administration with ImmTOR, as it resulted in a long-term therapeutic effect in a paediatric liver metabolic disorder, including the prevention of oncogenesis. Impact and implications: Redosing of gene therapy for inborn hepatobiliary disorders may be essential as effect wanes during hepatocyte division and renewal, particularly in paediatric patients, but the approach may carry long-term risks of liver cancer. Viral vectors carrying a therapeutic gene exerted a durable cure of progressive familial intrahepatic cholestasis type 3 in infant mice and reduced the risk of liver cancer only following a second administration.
Article
Advances in adeno-associated virus-based gene therapy are transforming our ability to treat rare genetic disorders and address other unmet medical needs. However, the natural prevalence of anti-adeno-associated virus neutralizing antibodies in humans currently limits the population who can benefit from adeno-associated virus-based gene therapies. Neonatal Fc receptor plays an essential role in the long half-life of IgG, a key neutralizing antibody. Researchers have developed several neonatal Fc receptor-inhibiting monoclonal antibodies to treat autoimmune diseases, as inhibiting the interaction between neonatal Fc receptor and IgG Fc can reduce circulating IgG levels to 20%-30% of the baseline. We evaluated the utility of one such monoclonal antibody, M281, to reduce pre-existing neutralizing antibody levels and to permit gene delivery to the liver and heart via systemic adeno-associated virus gene therapy in mice and nonhuman primates. M281 successfully reduced neutralizing antibody titers along with total IgG levels; it also enhanced gene delivery to the liver and other organs after intravenous administration of adeno-associated virus in neutralizing antibody-positive animals. These results indicate that mitigating pre-existing humoral immunity via disruption of the neonatal Fc receptor-IgG interaction may make adeno-associated virus-based gene therapies effective in neutralizing antibody-positive patients.
Article
The lipid nanoparticle (LNP)-encapsulated, nucleoside-modified mRNA platform has been used to generate safe and effective vaccines in record time against COVID-19. Here, we review the current understanding of the manner whereby mRNA vaccines induce innate immune activation and how this contributes to protective immunity. We discuss innate immune sensing of mRNA vaccines at the cellular and intracellular levels and consider the contribution of both the mRNA and the LNP components to their immunogenicity. A key message that is emerging from recent observations is that the LNP carrier acts as a powerful adjuvant for this novel vaccine platform. In this context, we highlight important gaps in understanding and discuss how new insight into the mechanisms underlying the effectiveness of mRNA-LNP vaccines may enable tailoring mRNA and carrier molecules to develop vaccines with greater effectiveness and milder adverse events in the future.