You are on page 1of 14

European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

Contents lists available at ScienceDirect

European Journal of Pharmaceutics and Biopharmaceutics


journal homepage: www.elsevier.com/locate/ejpb

Graphical Review

Activation of the complement system by nanoparticles and strategies for


complement inhibition
Hajira B. Haroon a, b, Elisha Dhillon a, Z. Shadi Farhangrazi c, d, Panagiotis N. Trohopoulos e,
Dmitri Simberg f, g, S. Moein Moghimi a, b, g, *
a
School of Pharmacy, Newcastle University, Newcastle upon Tyne NE1 7RU, UK
b
Translational and Clinical Research Institute, Faculty of Health and Medical Sciences, Newcastle University, Newcastle upon Tyne NE2 4HH, UK
c
S. M. Discovery Group Inc, CO, USA
d
S. M. Discovery Group Ltd, Durham, UK
e
CosmoPHOS Ltd, Thessaloniki, Greece
f
Translational Bio-Nanosciences Laboratory, Department of Pharmaceutical Sciences, Skaggs School of Pharmacy, University of Colorado Anschutz Medical Center,
Aurora, CO, USA
g
Colorado Center for Nanomedicine and Nanosafety, University of Colorado Anschutz Medical Center, Aurora, CO, USA

A R T I C L E I N F O A B S T R A C T

Keywords: The complement system is a multicomponent and multifunctional arm of the innate immune system. Comple­
Complement ment contributes to non-specific host defence and maintains homeostasis through multifaceted processes and
Complement inhibitors pathways, including crosstalk with the adaptive immune system, the contact (coagulation) and the kinin systems,
Complement regulators
and alarmin high-mobility group box 1. Complement is also present intracellularly, orchestrating a wide range of
Complement receptors
Complosome
housekeeping and physiological processes in both immune and nonimmune cells, thus showing its more so­
Nanoparticles phisticated roles beyond innate immunity, but its roles are still controversial. Particulate drug carriers and
Opsonisation nanopharmaceuticals typically present architectures and surface patterns that trigger complement system in
different ways, resulting in both beneficial and adverse responses depending on the extent of complement
activation and regulation as well as pathophysiological circumstances. Here we consider the role of complement
system and complement regulations in host defence and evaluate the mechanisms by which nanoparticles trigger
and modulate complement responses. Effective strategies for the prevention of nanoparticle-mediated comple­
ment activation are introduced and discussed.

1. Introduction accomplished through opsonisation, inflammatory, and lytic processes


[1,2]. Thus, complement contributes to robust non-specific host defence
1.1. The complement system by orchestrating the clearance of immune complexes, cell debris, and
foreign particles by phagocytic cells, and evoking controlled inflam­
The complement system, comprising more than 50 fluid-phase and matory and cytolytic responses to maintain homeostasis [1,2]. The
membrane-bound proteins that include pattern-recognition molecules, broader participation of the complement in homeostasis is reflected in
zymogens, regulators, and inhibitors (Table 1), is among the major ef­ crosstalk with the coagulation (contact) system, for instance, to resolve
fectors of the innate immune system [1,2]. Complement is present endothelial injury and facilitate clotting, with the kinin system to
mainly in the blood, but some complement proteins are also available in amplify inflammation, and alarmin high-mobility group box 1 in alarm
the interstitial and extracellular space, in the lymph, in tear fluid, in the raising and inflammation [3,4]. Intracellularly (within subcellular
vitreous, in the cerebrospinal fluid, in serous exudates on mucosal sur­ compartments and organelles), complosome (cell-autonomous comple­
faces (e.g., oral cavity and the airways) and intracellularly (complo­ ment) apparently serves as a regulator of many physiological processes,
some). Complement serves as a functional bridge between innate and including cell metabolism (e.g., glycolysis, mitochondrial respiration),
adaptive immunity and its effector functions are predominantly autophagy, and gene regulation [5]. However, the exact roles of

* Corresponding author at: School of Pharmacy, Newcastle University, Newcastle upon Tyne NE1 7RU, UK.
E-mail address: seyed.moghimi@ncl.ac.uk (S.M. Moghimi).

https://doi.org/10.1016/j.ejpb.2023.11.006
Received 3 October 2023; Received in revised form 31 October 2023; Accepted 6 November 2023
Available online 8 November 2023
0939-6411/© 2023 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

Table 1 step process (recognition, proteolytic activation, and expression of


Activation components, regulators and receptors of the complement system with biologic activities). There are three major complement activation
corresponding chromosome location. pathways: classical, lectin, and alternative (Fig. 1) [1,2]. Although each
Activation Chromosome Protein regulators & Chromosome pathway has some unique proteins and enzymes, the activities that
components receptors result are identical for all three pathways, including opsonisation,
C1q A chain 1p36.12 C1 inhibitor 11q12.1 inflammation, chemotaxis, and biomembrane lysis (Fig. 1) [1,2]. Acti­
C1q B chain 1p36.12 C4 binding protein α 1q32.2 vation of the classical pathway is initiated mainly by the binding of
C1q C chain 1p36.12 C4 binding protein β 1q32.1 complement pattern-recognition molecule C1q to immune complexes
C1r 12p13.31 Factor H 1q31.3
(but also to clusters of anionic charge or hydrophobic motifs on particles
C1s 12p13.31 Factor H related-1 1q31.3
C2 6p21.33 Factor H related-2 1q31.3 and cells as well as to surfaces with immobilised C-reactive protein)
C3 19p13.3 Factor H related-3 1q31.3 followed by sequential activation of two zymogens C1r and C1s,
C4A 6p21.33 Factor H related-4 1q31.3 respectively [6]. The lectin pathway is initiated by the binding of
C4B 6p21.33 Factor H related-5 1q31.3 pattern-recognition molecules mannose binding lectin (MBL), ficolins
C5 9q33.2 Factor I 4q25
C6 5p13.1 Clusterin (SP-40) 8p21.1
and collectin-10 and -11 (Fig. 1) to the pathogen/foreign particle-
C7 5p13.1 Vitronectin 17q11.2 associated molecular patterns (e.g., carbohydrate structures and acety­
C8 α chain 1p32.2 CD46 1q32.2 lated residues) as well as to the damage-associated molecular patterns
C8 β chain 1032.2 CD55 (DAF) c 1q32.2 on apoptotic, necrotic and malignant cells, resulting in activation of
C8 γ chain 9q34.3 CD59 (protectin) 11p13
MBL-associated serine proteases (MASPs) [1,2]. MBL is the most studied
C9 5p13.1 CD93 (C1qRp) d 20p11.21
Factor B 6p21.33 CR1 (CD35) e 1q32.2 pattern-recognition molecule of the lectin pathway and the ligand
Factor D 19p13.3 CR2 (CD21) 1q32.2 binding properties and the effector function of MBL is highly dependent
Properdin (P) Xp11.23 CR3A (CD11b) 16p11.2 on its oligomeric state [9,10]. Regardless of the initiation pathway,
MBL-2 a 10q21.1 CD11c 16p11.2 complement activation leads to proteolytic cascades and the assembly of
Ficolin 1 9q34.3 CD18 (LFA-1) f 21q22.3
convertase enzymes that converge toward proteolytic cleavage and
Ficolin 2 9q34.3 CRIg g Xq12
Ficolin 3 1p36.11 C3a receptor 12p13.31 activation of the third complement protein (C3) (Figs. 1 & 2) [1]. In
Collectin-10 8q24.12 C5a receptor 1 19q13.32 contrast to the classical and the lectin pathways, the alternative pathway
Collectin-11 2p25.3 CSMD1 h 8p23.2 maintains constant surveillance through the spontaneous hydrolysis of
MASP-1 b 3q27.3 CSMD2 h 1p35.1
C3 (the “tickover” hypothesis) at a rate of ~1 % of total C3 per hour,
MASP-2 b 1p36.22 CSMD3 h 8q23.3
Elastase 19p13.3 which generates C3b [11]. There are also bypass pathways that directly
Coagulation Factor II 11p11.2 cleave C3 (Fig. 1). Notwithstanding, C3 activation culminates in the
a generation of two highly potent anaphylatoxins (first C3a, and then C5a
MBL = Mannose binding lectin.
b through the subsequent assembly of C5 convertases), opsonisation of
MASP-1 and MASP-2 = MBL-associated serine protease-1 and -2; alternative
splicing of the pre-mRNA encoding MASP-1 generates MASP-3 and MAp44 intruders’ surfaces with C3b and iC3b, which aids their recognition by
(MAP-1). MAp44 competes with MASP-2 for binding to MBL and ficolins, complement receptors (CRs) of phagocytic cells (Fig. 2), and the
resulting in inhibition of complement activation through the lectin pathway. sequential assembly of the lytic membrane attack complex (MAC) [1,2].
c
DAF = Decay accelerating factor. At the core of all the three-complement pathways lies a highly efficient
d
C1qRp = C1q receptor, selectively expressed by cells with a myeloid lineage, C3 amplification loop: a C3 feedback cycle, which enhances amplifica­
endothelial cells, platelets and microglia. tion, and the C3 breakdown cycle, which down-regulates it by gener­
e
CR = Complement receptor. ating iC3b as its primary product (Fig. 1). The main function of the
f
LFA-1 = Lymphocyte function-associated antigen-1. amplification loop is to deposit large quantities of C3b on the surface of
g
CRIg = V-set and immunoglobulin domain-containing 4 (a receptor for C3b
the activating particle/pathogen [1,2].
and iC3b opsonised pathogens).
h
CSMD 1, 2 and 3 = CUB and Sushi multiple domains 1, 2 and 3, respectively
(transmembrane proteins with complement inhibitory properties). 3. Complement regulation and therapeutics

Complement activation is tightly regulated as to prevent uncon­


complosome are still controversial and not universally accepted [5]. All
trolled inflammation and minimising bystander tissue damage [1,2].
in all, complement contributes to protection, tissue development, and
There are many types of soluble and membrane-anchored complement
regeneration, metabolic regulation, and disease pathophysiology.
regulators and inhibitors (Table 1) to ensure complement regulation at
Similar to pathogens, particulate drug carriers and nano­
many steps along complement activation pathways [1,2]. Of particular
pharmaceuticals exhibit diverse architectural arrangements and surface
importance are regulators acting at the level of the convertases (by
patterns that could be sensed by complement, where responses could
affecting their assembly as well as their proteolytic activity) and during
dramatically affect biological and therapeutic performance of adminis­
the assembly of the MAC (Figs. 1 & 2). For example, the fluid-phase
tered nanoparticles [6–8]. On the other hand, if nanoparticles derail
complement regulator factor I cleaves C3b and generates inactive frag­
complement activation this might instigate inflammatory reactions [7].
ments iC3b and C3c, but to prevent non-specific C3b degradation, it
Thus, identification and systematic mapping of physicochemical and
requires cofactor activity for its proteolytic function provided by factor
biological parameters that trigger (or evade) complement activation by
H, membrane cofactor protein (MCP, CD46) and CR1 (Fig. 2) [1,2,12].
therapeutic nanoparticles could open the path to any number of possi­
On the other hand, complement regulators such as decay accelerating
bilities, which in conjunction with artificial intelligence and machine
factor (DAF or CD55) accelerates the decay of C3 and C5 convertases by
learning algorithms, portends to fast-track and new design of nano­
competing with factor B for binding to surface-bound C3b and by dis­
particle libraries with tuneable complement responses. Here we high­
placing Bb from convertases, whereas CD59 interferes with the assembly
light complement activation mechanisms by nanoparticles and discuss
of MAC [1,2,12].
current nanoengineering and therapeutic trends in tuning complement
Notwithstanding, erroneous activation, over-activation, and/or
responses to administered nanoparticles.
insufficient regulation of the complement might turn its destructive
actions against the host [12,13]. Atherosclerosis is a fine, yet, intricate
2. Complement activation
example, where complement exhibits both protective and proathero­
genic properties, and reviewed in detail elsewhere [14]. Here, the role of
Complement activation in the blood may be summarized as a three-
complement is not only driven by plasma-derived complement factors,

228
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

but also there are indications that distinct lesional macrophage subsets Complement over-activation by enzymatically-modified low-density li­
are potential sources of local complement production [14]. Considering poproteins and cholesterol crystals has also been suggested to contribute
the latter, recently, a complement regulatory axis in atherosclerotic to pro-inflammatory reactions, macrophage death, and smooth muscle
plaques was realised, where factor H non-redundantly regulates the C3- cell proliferation in atherosclerotic plaques [14]. Cancer is another
dependent efferocytosis capacity of lesional macrophages [15]. In example of delicately taking advantage of complement proteins and
particular, factor H derived from inflammatory phagocytes exacerbates complement cleavage products for progression. For instance, intra­
atherosclerosis by limiting lesional apoptotic cell clearance [15]. tumoural liberation of C5a (e.g., as a result of intratumoural coagulation

(caption on next page)

229
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

Fig. 1. Complement activation pathways. The scheme shows sequential enzymatic steps in classical, lectin and alternative complement activation pathways, key
fluid-phase complement regulators (e.g., C1-inhibitor, Factor H, Factor I) and selected complement activation products (C4a, C4b, C4d, C3a, C3b, iC3b, C3bc, Bb,
C5a, sC5b-9) that can be measured quantitatively. Antibodies such as IgM, IgG and IgA, depending on their isotype, glycosylation and conformation, can trigger
complement activation through the three pathways. The C-reactive protein (CRP, a pentameric acute-phase protein) on binding to nanoparticles might also trigger
complement activation through the C1 complex arm of the classical pathway. The binding of complement pattern recognition molecules C1q and collectins (e.g.,
MBL, ficolins and collectin-11) to surfaces usually triggers activation of their associated proenzymes (C1r and C1s in the case of C1q; MASPs in the case of collectins),
resulting in complement activation through classical and lectin pathways, respectively. Complement activation through classical and lectin pathways cleaves C4,
resulting in liberation of C4a. Classification of C4a as a complement anaphylatoxin is debatable [95]. Alternative pathway functionality is through autoactivation of
soluble C3 that undergoes slow spontaneous hydrolysis [C3(H2O)], or when nascent C3b undergoes nucleophilic attack for instance, by hydroxyl or amine moieties
on nanoparticle surfaces. The first critical stage in complement cascade is the formation of C3 convertases [C4b2b (formerly called C4b2a) of classical and lectin
pathways and C3bBb of the alternative pathway] that proteolytically cleave C3. This results in the liberation of anaphylatoxin C3a and the opsonic fragments C3b and
iC3b. The next step is the assembly of pathway-specific C5 convertases [C4b2b3b (formerly called C4b2a3b) and C3bBbC3b], which cleaves C5, resulting in liberation
of another anaphylatoxin (C5a) and activation of the terminal pathway of complement. On full complement activation the membrane attack complex C5b-9 is
formed. In soluble form, C5b-9 is bound to vitronectin (sC5b-9). The amplification loop of the alternative pathway is responsible for propagating the responses of all
three-complement pathways by producing large amounts of C3b. Apart from complement convertases, proteases of the coagulation system such as kallikrein and
thrombin can directly cleave C3 and C5 (not shown in the scheme), leading to additional sources of anaphylatoxins C3a and C5a. Similarly, MASP-1 may directly
cleave C3. MASP-3 activates the zymogen factor D of the alternative pathway. The scheme also shows selected natural complement regulators [factors I and H, C1
esterase inhibitor, MAP-1 (MAp44, see Table 1) and small MBL associated protein (sMAP or MAp19)] and sites where they exert their action. The figure and the
legend are reproduced with permission with slight modifications [8].

processes that cleaves C5) helps tumour growth and invasiveness in Finally, C1 esterase inhibitor is another complement inhibitor, which is
many ways; for instance, by promoting immunosuppression, since C5a is widely used at experimental levels acting on classical (inactivating C1r
chemoattractant to myeloid-derived suppressor cells [16,17]. Many tu­ and C1s) and lectin (inactivating MASP-1 and MASP-2) pathways [31].
mours also overexpress complement proteins such as C1q, which pro­ In addition, C1 esterase inhibitor is the inhibitor of the kinin, fibrino­
mote tumourigenesis by favouring angiogenesis and metastasis lytic, and contact activation pathway of the coagulation cascade [31]. It
independently of complement activation [18]. Complement dysregula­ is in this context, and not as a complement inhibitor that the substitution
tion can occur due to genetic mutations (Table 1), as a result of disease of a natural C1 esterase inhibitor and a recombinant protein C1 esterase
processes, or because of the development of autoantibodies to comple­ inhibitor has been approved for the inherited disorder hereditary
ment components [12,13]. A classic example is paroxysmal nocturnal angioedema in patients deficient in C1 esterase inhibitor (Table 3) [31].
haemoglobinuria (PNH), caused by a somatic mutation in the phos­ The deficiency of the C1 esterase inhibitor results in dysregulation of the
phatidylinositol glycan complementation class A gene in haematopoietic plasma kallikrein-kinin system and excessive bradykinin production,
stem cells [19]. This mutation affects the biosynthesis of glyco­ resulting in episodes of soft tissue swelling.
sylphosphatidylinositol (GPI), where a chronic reduction or absence of
GPI leads to a deficiency in membrane-anchored complement regulators 4. Complement activation by nanoparticles
resulting in inadvertent complement attack on erythrocytes [19].
There are many therapeutic opportunities for inhibiting or modu­ Numerous studies have demonstrated complement activation by
lating the complement system at preclinical and clinical levels different organic and inorganic nanoparticles and regulatory-approved
[13,20,21]. For instance, United States FDA-approved complement nanopharmaceuticals in human serum, plasma, and full blood
therapeutic interventions include a humanised recombinant monoclonal (reviewed in [6,7]). Generally, complement activation could compro­
antibody against C1s, designer biologics such as pegcetacoplan, which mise nanocarrier stability (e.g., liposome and lipid nanoparticles),
inhibits the central conversion of C3 to C3b and C3a, recombinant causing drug leakage, promoting premature clearance of nanoparticles
humanised anti-C5 monoclonal antibodies, and a C5a receptor antago­ by the blood and tissue phagocytic cells and compromising their ther­
nist for suppressing anaphylatoxin signalling (Table 2) [21–24]. Inhi­ apeutic efficacy for intended non-phagocytic cell targets, and when
bition of C3 and its convertases (e.g., using compstatin and its related uncontrolled, might induce adverse reactions and promote disease
family members Cp40 and pegcetacoplan) has the broadest effect, since progression [6,7].
this blocks the amplification loop and, in principle down regulates the Regarding the mechanistic issues, recent studies have shown that in a
terminal activity of the complement pathway. However, strong activa­ typical nanoparticle suspension, only a small fraction of nanoparticles
tion of calcium-sensitive pathways could still lead to C5 activation in­ (or their aggregates in plasma) actively trigger complement activation
dependent of C3 (the so-called C3 bypass pathway) [25,26]. While C3 [32,33]. This is a reflection of nanoparticle population heterogeneity.
(and C5) inactivation in humans may increase susceptibility to Menin­ While the precise mechanisms are still under investigation, there are
gococcal and other upper respiratory tract infections (e.g., Pneumo­ indications that complement activation could proceed on binding of
coccal and Haemophilus influenza type B infections) (Table 2) [27], nascent C3b or C3(H2O) to some nanoparticles, resulting in surface as­
prolonged treatment of non-human primates with compstatin neither sembly of the alternative pathway C3 convertase (Fig. 3) [32,33]. While
weakened immune system nor susceptibility to infection [28]. However, direct binding of C3b or C3(H2O) to pristine surfaces is possible, recent
patients undergoing C5, C3 and C1s complement inhibitor therapies are studies have shown predominant binding to surface-adsorbed proteins
recommended for vaccination against encapsulated bacteria (Table 2) [32,33]. For example, C3b binding to a few surface-deposited antibodies
(e.g., Bexsero, Prevnar® and Menitorix vaccines). Other examples of has been suggested [32], and some antibodies can potentially trigger
complement modulators at preclinical or undergoing clinical evaluation complement activation through the alternative pathway [34]. This
includes soluble form of natural complement regulators (e.g., sCR1, “seed” population of nanoparticles with C3 convertases on their surfaces
sDAF, sMCP and sCD59), small molecule factor B and factor D inhibitors generate large quantities of C3b molecules from native C3 in plasma (C3
(iptacopan and danicopan, respectively) for oral administration and is a highly abundant serum protein with a normal range of 0.8–1.6 mg/
anti-factor D antibody fragments, peptide and antibody based inhibitors mL) that potentially opsonises bystander nanoparticles in their vicinity.
of C1s and activated C1s (apart from Enjaymo®, Table 2), anti-MASP3, This, in turn, could accelerate and amplify complement activation
anti-C2, anti-C5a, anti-C6 and Factor Bb antibodies, anti-C5 aptamer through the assembly of C3 convertases on surfaces of more nano­
and fusion constructs such as CR2–factor H (TT30) and CR2–CR1 (TT32) particles [33]. However, C3 opsonisation of nanoparticles was shown to
[13,20,21,29,30]. For instance, with the latter, the CR2 moiety recog­ be slow and inefficient, but reversible [33,35]. The latter was ascribed to
nises surface-deposited C3, whereas CR1 blocks C3 convertases [30]. the dynamic process of protein adsorption–desorption at the

230
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

(caption on next page)

231
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

Fig. 2. Schematic representation of C3 structure and its sequential fragmentation by complement regulators. C3 is a member of α-2 macroglobulin family of proteins,
which has a 13-domain structure and two polypeptide chains (α and β chains) held together by a single disulfide bond. C3 convertases of classical, lectin and
alternative pathways cleave the α chain of C3 resulting in generation of two fragments C3a and C3b (the α chain of C3b is usually referred to as α′). C3a (a 9 kDa
peptide) is an anaphylatoxin with a short half-life in plasma (it is rapidly processed by carboxypeptidases) exerting proinflammatory functions via its receptor (C3aR)
expressed by a number of immune cells such as monocytes/macrophages, mast cells, neutrophils, basophils, and eosinophils. C3b is an opsonic molecule when is
bound covalently to nanoparticle surfaces aids nanoparticle recognition by cells expressing complement receptors 1 (CR1 or CD35), 3 (CR3 or CD11b/CD18) and 4
(CR4 or CD11c/CD18) and complement receptor of the immunoglobulin family (CRIg). For example, human Kupffer cells express CR1, CR3 and CR4 and human
dendritic cells express CRIg. C3b can also initiate formation of the alternative pathway C3 convertase as well as classical, lectin and alternative pathway C5 con­
vertases (see Fig. 1). Surface bound C3b is rapidly inactivated (due to release of the 2 kDa fragment C3f) by complement regulators such as factor H (FH) and FH like
1 protein in tandem with factor I (FI) to generate inactive C3b (iC3b). Plasma membrane bound complement regulators such as CR1 and CD46 (membrane cofactor
protein) also inactivate C3b bound to cell surfaces, thus protecting cells from complement damage. iC3b is also an opsonic molecule recognised by cells expressing
CR3, CR4 and CRIg. Further processing of iC3b by complement regulators FI, CR1 and other proteases (e.g., elastase) release the fluid phase fragment C3c and the
surface-bound fragment C3dg. C3dg is further processed by proteases (e.g., thrombin) to C3d, which remain bound to surfaces (e.g., nanoparticles). Surface-bound
C3dg/C3d are recognised by CR2 (or CD21). B lymphocytes express CD21; however, the role of B cells in the uptake of complement opsonised nanoparticles has not
received much attention. The lower part of the scheme shows an intramolecular thioester bond in the C3d region of the α chain, which is susceptible to nucleophilic
attack. This becomes the exposed reactive moiety of the metastable binding site of proteolytically produced C3b. The thioester generates a carbonyl group that binds
the C3b irreversibly to reactive pristine groups (e.g., hydroxyl and amine moieties) expressed either on nanoparticles and/or on their surface-bound biomole­
cule corona.

Table 2
Currently approved complement inhibitors by the United States Food and Drug Administration (FDA).
Inhibitor Class Target FDA Current indication
approval
year

Soliris® Recombinant humanised C5 2007 Paroxysmal nocturnal haemoglobinuria (PNH); atypical haemolytic uremic
1
(Eculizumab) monoclonal antibody (IgG2/4 syndrome (aHUS); generalised myasthenia gravis (gMG) in adult patients who are
κ) anti-acetylcholine receptor antibody positive; neuromyelitis optical spectrum
disorder in adult patients who are anti-aquaporin-4 antibody positive
Ultomiris® Recombinant humanised C5 2018 Adult and paediatric patients one month of age and older with PNH or aHUS;
(Ravulizumab) 2 monoclonal antibody (IgG4) patients with gMG who are anti-acetylcholine receptor antibody positive
Tavneos® Small molecule C5aR1 (CD88) 2021 Adjunct treatment of adult patients with severe active anti-neutrophil cytoplasmic
(Avacopan) 3 C33H35F4N3O2 antagonist autoantibody-associated vasculitis
Empaveli® Pegcetacoplan (injection for C3 2021 Adult patients with PNH
(Aspaveli®) 4 subcutaneous use)
Enjaymo® Humanised recombinant C1s 2022 Cold agglutinin disease
(Sutimlimab-jome) monoclonal IgG4 antibody
5
6
Syfovre® Pegcetacoplan (for intravitreal C3 2023 Geographic atrophy secondary to age-related macular degeneration
injection)

Soliris®, Ultomiris®, Empaveli® and Enjaymo®, labels warn on serious infections caused by encapsulated bacteria and recommend appropriate vaccination strategies
prior to the inhibitor administration.
1
https://www.accessdata.fda.gov/drugsatfda_docs/label/2019/125166s431lbl.pdf (accessed 30/10/2023).
2
https://www.accessdata.fda.gov/drugsatfda_docs/label/2022/761108s023lbl.pdf (accessed 30/10/2023).
3
https://www.accessdata.fda.gov/drugsatfda_docs/label/2021/214487s000lbl.pdf (accessed 30/10/2023).
4
https://www.accessdata.fda.gov/drugsatfda_docs/label/2021/215014s000lbl.pdf (accessed 30/10/2023).
5
https://www.accessdata.fda.gov/drugsatfda_docs/label/2022/761164s000lbl.pdf (accessed 30/10/2023).
6
https://www.accessdata.fda.gov/drugsatfda_docs/label/2023/217171s000lbl.pdf (accessed 30/10/2023).

or both [6,7,36]. Here, two interrelated parameters that modulate


Table 3
complement activation are particle size (curvature) and surface archi­
FDA approved C1 esterase inhibitor preparations as inhibitor of the kinin,
tecture (Fig. 4) [6]. The latter includes chemical composition, the
fibrinolytic and contact activation pathway.
spacing arrangement/periodicity of surface functional groups/ligands,
Inhibitor Class FDA Current indication and the structural/conformational state of surface projected macro­
approval
year
molecules (e.g., synthetic polymers, polysaccharides) [6,7,36]. Collec­
tively, these parameters modulate complement activation through
Cinryze Human plasma-derived 2008 Hereditary angioedema
multivalent engagement with and conformational regulation of surface-
concentrate of C1 (HAE) in adults, adolescents
esterase inhibitor and paediatric patients (6 bound antibodies and complement pattern-recognition molecules
years of age and older (Fig. 4). For instance, unbound IgM assumes a planar structure and does
Berinert Human plasma-derived 2009 Acute abdominal, facial, or not activate complement, while bound IgM must be in strained confor­
concentrate of C1 laryngeal HAE in adults and mation to trigger complement activation through the C1 complex
esterase inhibitor paediatric patients
[37,38]. Optimal straining of IgM requires multiple Fab interactions
Ruconest Recombinant protein 2014 Acute attacks in adult and with surface-exposed epitopes. This is best achieved on the surface of
C1 esterase inhibitor adolescent patients with spherical particles in the 100–250 nm size range [38]. A dramatic
HAE decrease or increase in particle size, however, minimises multivalent
engagement of all Fab regions leading to insufficient IgM straining;
nanoparticle-plasma interface [35]. otherwise, high-affinity interactions are needed to sufficiently strain the
Depending on their physicochemical properties, nanoparticles could Cµ2 domain of the immunoglobulin [38]. Thus, curvature- and epitope
also trigger complement activation through classical or lectin pathways, density-dependent variables on nanoparticles could explain inter-

232
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

Fig. 3. A proposed model for complement activation and C3 opsonisation through non-specific plasma protein adsorption on nanoparticles. The scheme shows a
dextran (yellow and yellow arrows) coated superparamagnetic iron oxide nanoworm (the brown region and the white star represent part of the iron core) interacting
with plasma proteins (grey). Experiments have shown that some surface adsorbed plasma proteins and/or plasma proteins intercalated into dextran chains inad­
vertently cause activation of the alternative pathway of the complement system. The hypothesis state that adsorbed and/or intercalated plasma proteins may undergo
conformational changes or form clusters that expose reactive moieties that interact with nascent C3b or C3(H2O) (green). This leads to the formation of the
alternative pathway convertase (C3bBbP, white circle) only on a small fraction of nanoparticles (Bb = red and P = dark blue). These convertases generate additional
C3b molecules that opsonise nearby nanoparticles. Protein-C3b complexes are continuously formed and released (light blue circle) from the surfaces of nanoparticles.
Thus, this turnover might minimise surface deposition of fluid-phase regulators such as factors H and I. Accordingly, nanoparticles in the blood may continuously
activate complement through the alternative pathway turnover. Considering the above, the hypothesis does not disregard surface deposition of IgG–C3b complexes
that are already formed in the fluid phase and with the ability to assemble C3bBbP convertases. P = properdin.

individual differences in the extent of complement activation on IgM alternative pathways [reviewed in 6,7,36]. Many clinically approved
binding. Curvature also plays a role in IgG binding for a given Fc density, nanopharmaceuticals also contain poly(ethylene glycol) (PEG)-lipid
where activation of the classical pathway depends on antigen-driven conjugates [41,42]. However, the emergence of antibodies against poly
multivalent engagement with IgG clusters, particularly IgG hexamers (ethylene glycol) (PEG) [43–46], where the majority exhibit low affinity
[39]. The composition of N-glycans in the CH2 domain of IgG also plays for PEG, has raised the question of whether low-affinity anti-PEG anti­
an important role in hexamer assembly through IgG Fc:Fc interactions bodies can reach sufficient straining (as in IgM) or form clusters (as in
[40] and, hence, modulation of complement activation through C1q IgG) on PEGylated surfaces to trigger classical pathway activation [6,7].
binding. While, surface PEG density, conformation and mobility could modulate
Liposomes and lipid nanoparticles are among the most popular ve­ the extent of antibody-dependent classical pathway activation, com­
hicles in drug delivery, and there are many examples of regulatory- plement activation might prevail on the binding of anti-PEG antibody
approved lipid-based nanopharmaceuticals and vaccines [41,42]. sub-classes that trigger the alternative pathway [47].
Many studies have shown that binding of naturally occurring anti- Ultrastructurally, pattern-recognition molecules such as C1q, fico­
phospholipid and anti-cholesterol antibodies to liposomes and lipid lins, and MBL oligomers sense surface patterns/target motifs spaced at
nanoparticles could trigger complement activation through classical and broad intervals of 2–20 nm (Fig. 4) [9,48–50]. This ensures recognition

233
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

Fig. 4. Schematic summary showing the role of nanoparticle curvature and surface pattern periodicity on complement activation. Antibody-dependent complement
activation depends on the conformation and strain of the binding antibody (e.g., as a single IgM or clusters of IgG), which, in turn, is modulated by nanoparticle
curvature and spacing/surface density of exposed epitopes (left panel). Structural and geometrical restrictions in pattern recognition molecules such as C1q and MBL
allow these molecules to sense surface patterns with dimensions in the range of 2–20 nm, thereby triggering complement activation through classical and lectin
pathways, respectively (right panel). Surfaces that display pattern periodicity below the nanometer-scale spacing arrangement are expected to escape complement
surveillance by C1q and MBL. This is known as the “angstrom-scale spacing arrangement (ASSA) phenomenon” for complement evasion. However, complement
activation through non-specific protein adsorption could still proceed. Reproduced with permission [6].

of target surfaces with wider and denser display of ligands. For instance, serine proteases or mannosylated proteins (e.g., apolipoprotein B100)
C1q recognises surface clusters of anionic charge or hydrophobic motifs and glycosylated variants of IgG and IgM that bind MBL [56].
[51], whereas MBL oligomers bind to sugars such as D-mannose and N-
acetyl-D-glucosamine [9]. Thus, nanoparticle surface pattern periodicity 5. Overcoming nanoparticle-mediated complement activation
and spacing are important in accommodating complement pattern-
recognition molecules. Cardiolipin liposomes [51] and poly(2-methyl- To date, different strategies have emerged for designing nano­
2-oxazoline)-coated nanoparticles [52] are examples of entities that particles that show little or no complement activation (Fig. 5). The
trigger human complement activation through C1q sensing. Variable simplest strategy is nanoparticle surface enrichment and/or function­
results, however, have been reported on C1q binding to carbon-based alisation with materials that recruit fluid phase complement regulators
nanomaterials. While one study reported C1q-dependent complement such as factor H. Examples include sialic acid, factor H-specific phage
activation by carbon nanotubes [53], another study found no comple­ peptides and heparin [57,58]. There are also serendipitous examples of
ment activation on C1q binding [54]. These inconsistencies are pre­ nanoparticles that do not trigger complement activation in human
sumably related to differences in surface functionality and chemistry of serum, such as hexosomes assembled from citrem and glyceryl mono­
carbon nanomaterials in modulating C1q conformation. There are also oleate [59]. Although the exact mechanisms with this class of hexosomes
interesting examples of nanoparticles that lack pristine ligands for were not investigated, it is plausible that the citric acid moiety of citrem
recognition by lectin pathway pattern-recognition molecules, yet they in combination with the glycerol component of glyceryl monooleate
activate the lectin pathway [55]. One example is poloxamine 908-coated form dynamic platforms resembling factor H binding domains on sialic
polystyrene nanoparticles, but surface-projected poloxamine 908 only in acid (factor H bind to glycerol side chain and carboxyl group of sialic
a mushroom-brush intermediate/brush conformation triggers comple­ acid [58,60]). In contrast to the above, another strategy is direct grafting
ment activation through the lectin pathway [55]. Poloxamine 908 is a of complement regulators to nanoparticles [61].
star-shaped block copolymer of ethylene oxide and propylene oxide With PEGylated liposomes, the anionic phosphate-oxygen of the
blocks, generating repetitive patterns of relative polarity and hydro­ PEG-phospholipid conjugate was shown to play a key role in comple­
phobicity on surface adsorption, which might transiently resemble ment activation and methylation of phosphate-oxygen dramatically
structural motifs of natural substrates for MBL and ficolins. Additionally, reduced complement activation [62]. However, non-specific protein
selected plasma proteins might intercalate into the loops and trails of binding and protein trapping (including antibodies) in PEG clouds might
surface-projected copolymer blocks. Thus, an alternative possibility for still contribute to complement activation through the alternative
complement activation arises if intercalated proteins constitute pattern- pathway [35]. Thus, better strategies are needed to minimise protein
recognition molecules (e.g., MBL and ficolins) with their associated binding/trapping. One such approach is surface functionalization of

234
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

Fig. 5. Schematic representation of main approaches in overcoming complement activation by therapeutic nanoparticles. These approaches include nanoparticle
engineering with components that deter or repel complement pattern-recognition molecules as well as C3 opsonisation such as through polymer-pairing and surface
patterning with angstrom-scale periodicity, or strategies that attract fluid-phase complement regulators to nanoparticle surfaces, or direct functionalization of
nanoparticle surfaces with complement regulators, or co-administration of nanoparticles with designer complement inhibitors for precision surface complement
inhibition (e.g., CR2–CR1 fusion protein).

nanoparticles with combinations of short- and long-chain polymers. For alternative pathway turnover [64]. Generation 3 PAMAM dendrimers
instance, surface functionalization of nanoparticles with combinations are open ellipsoids, whereas generations 4 and 5 are “hard spheres”
of PEG-phospholipid conjugates with long and short PEG chains reduced [67]. Presumably, the presentation of end-terminal amine functional­
complement activation, since more rigid short PEG chains caused ities in ASSA, together with dendrimer shape and the very small size (<5
stretching of longer PEG chains contributing to more effective protein nm), accounts for the lack of alternative pathway activation. It would be
exclusion [63]. Alternatively, considering surface sensing by comple­ intriguing if the display of chemically reactive functional groups in ASSA
ment pattern-recognition molecules is restricted to spacing intervals of on engineered larger nanoparticles could minimise C3b binding and
2–20 nm [9,48,50], surfaces that display patterns in angstrom-scale surface assembly of the alternative pathway convertases. However,
intervals might also escape complement activation. This is referred to particle curvature and non-specific protein adsorption could further play
as “angstrom-scale spacing arrangement (ASSA) phenomenon” for a role in modulating complement responses. Also, whether larger par­
complement evasion and demonstrated with the low generation poly ticles and macro-scale platforms with surface patterns in the ASSA can
(amidoamine) (PAMAM) dendrimers of different end-terminal func­ escape sensing by complement pattern-recognition molecules and avoid
tional moieties (generations 3–5 with dendrimer sizes of 3.2–4.8 nm, complement activation remains to be tested.
respectively, that display 32–128 periodic surface functional moieties The aforementioned complement escape strategies are broadly
such as amine, carboxyl and pyrrolidone in a confined space with applicable to new nanoparticle design. However, universal approaches
angstrom-scale separation from each other) [64]. are needed to suppress complement activation by existing nano­
Typically, the internal thioester bond in the α-chain of nascent C3b pharmaceuticals and follow-on products. One approach is to co-
undergoes nucleophilic attack by structures rich in amine and hydroxyl administer soluble complement inhibitors with nanopharmaceuticals
functionalities; this increases alternative pathway turnover [65,66]. (Figs. 5 & 6). However, on the one hand, high doses of soluble inhibitors
Contrary to this, amine-terminated dendrimers do not directly affect may be needed due to suboptimal potency. On the other hand, co-

235
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

administration with long-lasting inhibitors (e.g., pegcetacoplan) could nanomolar efficacy, blocked complement opsonisation and overcome
lead to possible adverse effects and immunosuppression. To address nanoparticle uptake by monocytes/granulocytes in the blood [33]. It is
these shortfalls, recently complement activation was successfully avoi­ hypothesised that inhibiting the function of C3 convertases on the “seed”
ded when nanoparticles were co-administered with the short-circulating population of nanoparticles overcomes the subsequent cleavage of
fusion construct CR2–CR1 (Fig. 6) [33]. Since, only a small percentage plasma C3 and the opsonisation of bystander nanoparticles [33]. In rats,
of nanoparticles randomly trigger complement and C3 opsonised (i.e., CR2–CR1 also prevented lethargy caused by bolus injection of nano­
the “seed” population), CR2–CR1 exclusively targeted C3 fragments and particles, without inducing long-lasting complement suppression [33].
C3 convertases on nanoparticle surfaces with picomolar to low

Fig. 6. Inhibition of complement activation by selected synthetic complement inhibitors. The scheme depicts complement inhibition by two different groups of
designer complement inhibitors: compstatin family of peptides and a fusion peptide assembled from two complement regulators (CR2–CR1). Compstatin is a cyclic
tridecapeptide. Other family members include Cp40 (a 14 amino-acid cyclic peptide) and pegcetacoplan (consisting of two 15 amino-acid cyclic peptides conjugated
via a linear 40 kDa poly(ethylene glycol) molecule). Pegcetacoplan has a half-life of 8–10 days in humans compared with a half-life of 12 h for Cp40 in cynomolgus
monkeys. Compstatin binds native C3 and C3b and sterically interfere with the assembly of the C3 convertase complexes. Thus, unlike other C3-regulatory proteins,
the mechanism of action of compstatin does not involve the destabilisation of the C3 convertase or the accelerated degradation of the C3b. Compstatin neither
interfere with the generation of C3(H2O) nor blocks C3 cleavage by non-C3 convertase proteases. Similar to compstatin, Cp40 also binds to fluid phase and surface-
bound C3 targets. In contrast, pegcetacoplan, has weaker affinities but preferentially targets surface complement inhibition, presumably through an avidity effect
binding two C3b molecules. This mode of action prevents alternative pathway convertase binding to C3 and subsequent cleavage. Fusion constructs such as CR2–CR1
inhibit complement activation with picomolar to low nanomolar efficacy on many nanoparticles in human and rat serum. The CR2 moiety targets surface C3 deposits
(C3b, iC3b, C3c and C3d), but not the fluid phase C3. The CR1 moiety targets C3 convertases. Although being a short-circulating complement regulator (a fast phase
of 2.5 min and a slow phase of 64 min in rats), co-injection of CR2–CR1 with nanoparticles targets activated C3 on nanoparticle surfaces and inhibits complement
activation as well as acute responses in rats.

236
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

6. Conclusions and future directions complement proteins in addition to the secretion of complement pro­
teins from the pancreas into the duodenum, but oedema in the mucosa
The interactions between complement proteins and nanoparticles or bleeding might also provide local complement [81]. Another example
have long been realized; however, the importance of complement in is human tear, where C1q, C3, C4, C5, factor B, C9, and DAF have been
nanomedicine research is still underappreciated [6,7,36]. By virtue of detected in closed-eye tears, whereas only C3, C4, and factor B were
their structural complexity and pattern display, nanoparticles trigger or found in open-eye and reflex tears [82]. On the other hand, lactoferrin is
evade complement in different ways; however, there are species differ­ present in all tear types [82], and lactoferrin-derived peptides poten­
ences in complement responses to nanoparticles and the results cannot tially inhibit the classical pathway [83]. Thus, considering the afore­
be generalised and extrapolated from one species to another mentioned discussion, nanoparticle-mediated complement activation
[7,8,36,52,68]. For example, nanoparticle/antigen adjuvanticity is processes and response magnitude are expected to differ in different
improved on C3d/C3dg tagging as this ensures long-term B cell memory body compartments.
and optimal antibody production in mice [69–71] and, again, this Growing evidence suggests perturbations in complosome activities
concept was recently revisited and applied for enhancing the immuno­ could contribute to a wide range of human diseases [5,84]. However, the
genicity of lipid-nanoparticle mRNA vaccines in mice [72]. Importantly, impact of nanoparticle-based therapies on cell type-specific complosome
C3d/C3dg adjuvanticity has only been demonstrated in the murine functionality has been overlooked and rarely touched upon. Not only
model and, apparently, does not apply to human B cells [73]. This is many types of immune cells such as CD4+ and CD8+ T cells, B cells,
presumably related to differences between the function of CR1 and CR2 monocytes, macrophages and neutrophils have intracellular pools of C3
between the two species [73]. Another example is nanoparticles coated and C5 (and some also express intracellular C3a receptor and C5a re­
with poly(2-methyl-2-oxazoline), which triggers complement activation ceptor 1 that engage with ribosomes, inflammasomes and other intra­
in human but not in murine serum [52]. Here, C3 opsonisation was cellular machineries) [5,15,84–88], but also many types of healthy
essential for nanoparticle recognition by human blood leucocytes and nonimmune cells such as fibroblasts, hepatocytes, a range of endothelial
monocyte-derived macrophages [52]. Notwithstanding, our current and epithelial cells, neurons and pancreatic β-cells have active intra­
mechanistic understanding of nanoparticle-mediated complement acti­ cellular complement components [5,84,87]. Although the roles of
vation processes in human serum/plasma has opened exciting ap­ complosome are still debatable, nevertheless, it would be interesting to
proaches for the design and engineering of complement-evading consider to what extent nanoparticle localisation to nonimmune cells
entities. These developments are important since uncontrolled activa­ could perturb complosome. Lipid nanoparticles with cationic or ionis­
tion or over-activation of the complement system by therapeutic nano­ able cationic lipids, which are known to activate and fix complement,
particles could compromise their safety and therapeutic efficacy and serve as good examples, and the currently available cationic and ionis­
contribute to disease progression, but still, it is overlooked. For example, able cationic lipids have been shown to induce pro-inflammatory re­
intratumoral complement activation by extravasated long circulating sponses [89–91]. Considering these, cytoplasmic sensing of C3 can
drug-free nanoparticles has promoted tumour growth in syngeneic activate mitochondrial antiviral-signalling protein located in the outer
immunocompetent mice bearing TC-1 tumour [74]. By considering membrane of the mitochondria, peroxisomes, and mitochondrial-
inter-individual as well as species differences in complement responses associated endoplasmic reticulum membrane and induce pro-
to nanoparticles [7,8,36,75], we need more integrated studies that map inflammatory cytokine secretion [92].
and correlate materials’ angstrom- and nano-scale parameters to com­ Finally, continuous developments in multiparametric and plasmonic
plement responses not only extracellularly, but also intracellularly complement assays, multi-omics complement technologies, precision
(complosome) and depending on microenvironmental conditions (e.g., intracellular complement sensors, and species- and pan-specific com­
hypoxia can alter complement activity as well as regulation of comple­ plement inhibitors (including cell-permeable inhibitors) together with
ment proteins in many cell types [76]). Therefore, attention must also be artificial intelligence guidance are expected to address and help with the
paid to the local and compartmental concentration of immunoglobulins, precision engineering of complement- and complosome-tuneable ther­
complement proteins and regulators as well as non-specific proteins that apeutic nanoparticulate systems, thus paving the path towards stratified
could modulate complement responses. For instance, an early study in approaches in nanomedicine therapies. Efforts might extend to more
the leg lymph of normal men identified much lower levels of immuno­ sophisticated nanoparticle designs compatible with the contact and the
globulins and complement proteins than in serum, which was consistent kinin systems. In parallel, advances in “complotype” genetics (the
with the capillary sieving mechanism [77]. Thus lymph levels of IgG, repertoire of inherited rare or common polymorphisms in genes
IgM, C1q, C1s, C1 inhibitor, C4, and C9 were found to be 17 %, 7.4 %, encoding complement components) [93], which has an impact on sus­
12 %, 21 %, 16 %, 22 %, and 24 % of that of serum, respectively [77]. ceptibility to inflammatory diseases and conditions, might further help
The total mean complement haemolytic activity in lymph was approx­ with future stratification strategies in co-administering nanomedicines
imately 7-fold lower than in serum [77]. Another related example is with appropriate complement inhibitors.
salivary scavenger and agglutinin (SALSA), also known as gp340, and its Last, but not least, apart from their therapeutic potential, nano­
isoforms, which is found at mucosal surfaces of the lungs, oral cavity, particles could also be utilised as functional tools in clinical research and
gastrointestinal tract, and vagina as well as in amniotic fluid [78]. investigation. Thus, developments in “pseudo pathogen-like” nano­
SALSA binds to C1q, MBL, ficolins, surfactant proteins A and D, secre­ particles could substitute for highly virulent pathogens and provide
tory IgA, fibrin, and fibrinogen, which suggests a role in regulating simpler and safer ways of assessing complement responses. A functional
inflammation and immune responses [78]. Future studies are needed to example is nanoparticles bearing ~70 copies of the recombinant
assess whether the binding of SALSA to locally administered nano­ receptor-binding domain of SARS-CoV-2, which were employed to
particles plays a central role in directing local complement activation. monitor complement responses and C3 opsonisation in sera of vacci­
Nevertheless, studies with immobilised SALSA (SALSA coated onto a nated, convalescent and naïve individuals [94]. The results demon­
microtiter plate) have demonstrated complement activation, whereas strated the heterogeneity of C3 opsonisation and leukocyte uptake,
SALSA in the fluid phase caused a dose-dependent inhibition of the which might predict individual responses to clear virions and modulate
lectin pathway [79]. Also, not all complement proteins and regulators the severity of the COVID-19 [94].
are present in different body compartments, and their levels may vary in
disease processes. For example, complement components up to C5 have CRediT authorship contribution statement
been demonstrated in lumen samples obtained from bacterially infected
human intestines, but the MAC was absent from the lumen environment Hajira B. Haroon: Formal analysis, Writing – original draft, Writing
[80]. Intestinal epithelial cells have been suggested as a source of these – review & editing. Elisha Dhillon: Writing – original draft. Z. Shadi

237
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

Farhangrazi: Formal analysis, Writing – review & editing. Panagiotis [12] G. Bajic, S.E. Degn, S. Thiel, G.R. Andersen, Complement activation, regulation,
and molecular basis of complement-related disease, EMBO J. 34 (2015)
N. Trohopoulos: Formal analysis, Writing – review & editing, Funding
2735–2757.
acquisition. Dmitri Simberg: Formal analysis, Writing – review & [13] J. Barratt, I. Weitz, Complement factor D as a strategic target for regulating the
editing, Funding acquisition. S. Moein Moghimi: Conceptualization, alternative complement pathway, Front. Immunol. 12 (2021), 712572.
Formal analysis, Writing – original draft, Writing – review & editing, [14] M.G. Kiss, C.J. Binder, The multifaceted impact of complement on atherosclerosis,
Atherosclerosis 351 (2022) 29–40.
Funding acquisition. [15] M.G. Kiss, N. Papac-Milicevic, F. Porsch, D. Tsiantoulas, T. Hendrix, M. Takaoka, H.
Q. Dinh, M.-S. Narzt, L. Göderle, M. Ozsvár-Kozma, M. Schuster, N. Fortelny,
Declaration of Competing Interest A. Hladik, S. Knapp, F. Gruber, M.C. Pickering, C. Bock, F.K. Swirski, K. Ley,
A. Zernecke, C. Cochain, C. Kemper, Z. Mallat, C.J. Binder, Cell-autonomous
regulation of complement C3 by factor H limits macrophage efferocytosis and
The authors declare the following financial interests/personal re­ exacerbates atherosclerosis, Immunity 56 (2023) 1809–1824.
lationships which may be considered as potential competing interests: [16] M.M. Markiewski, R.A. DeAngelis, F. Benecia, S.K. Ricklin-Lichtsteiner,
A. Koutoulaki, C. Gerard, G. Coukos, J.D. Lambris, Modulation of antitumor
[P.N.T. declares financial interests in CosmoPHOS Ltd (Greece). Z.S.F. & immune responses by complement, Nat. Immunol. 9 (2008) 1225–1235.
S.M.M. declare financial interests in S. M. Discovery Inc. and S. M. [17] E. Magrini, L. Minute, M. Dambra, C. Garlanda, Complement activation in cancer:
Discovery Ltd. H.B.H., E.D. & D.S. declare that they have no competing Effects on tumor-associated myeloid cells and immunosuppression, Semin.
Immunol. 60 (2021), 101642.
financial interests or personal relationships that could have appeared to [18] R. Bulla, C. Tripodo, D. Rami, G.S. Ling, C. Agostinis, C. Guamotta, S. Zorzet,
influence the work reported in this paper.]. P. Durigutto, M. Botto, F. Tedesco, C1q acts in the tumour microenvironment as a
cancer-promoting factor independently of complement activation, Nat. Commun. 7
(2016) 10346.
Data availability
[19] M. Bessler, P.J. Hillmen, T. Miyata, N. Yamada, J. Takeda, L. Luzzatto,
T. Kinoshita, Paroxysmal nocturnal haemoglobinuria (PNH) is caused by somatic
No data was used for the research described in the article. mutations in the OIG-A gene, EMBO J. 13 (1994) 110–117.
[20] D. Ricklin, D.C. Mastellos, J.D. Lambris, Therapeutic targeting of the complement
system, Nat. Rev. Drug Discov. (2019), https://doi.org/10.1038/s41573-019-
Acknowledgements 0055-y.
[21] W.M. Zelek, L. Xie, B.P. Morgan, C.L. Harris, Compendium of current complement
S.M.M. acknowledges support by the European Union’s Horizon therapeutics, Mol. Immunol. 114 (2019) 341–352.
[22] C. de Castro, F. Grossi, I.C. Weitz, J. Maciejewski, V. Sharma, E. Roman, R.
2020 programme funded under H2020-EU.1.3. – Excellent Science – A. Brodsky, L. Tan, C. Di Casoli, D. El Mehdi, P. Deschatelets, C. Francois, C3
Marie Skłodowska-Curie Actions, grant agreement ID. 956544 (DIR­ inhibition with pegcetacoplan in subjects with paroxysmal nocturnal
NANO: Directing the Immune Response through Designed Nano­ hemoglobinuria treated with eculizumab, Am. J. Hematol. 95 (2020) 1334–1343.
[23] M. Kolev, T. Barbour, S. Baver, C. Francois, P. Deschatelets, With complements: C3
materials). Z.S.F. (S. M. Discovery Group Ltd.) is a partner organisation inhibition in the clinic, Immunol. Rev. 313 (2023) 358–375.
in DIRNANO consortium. H.B.H. is an Early Stage Researcher financially [24] C. Lamers, D.C. Mastellos, D. Ricklin, J.D. Lambris, Compstatins: the dawn of
supported by the DIRNANO programme. E.D. is a recipient of a research clinical C3-targeted complement inhibition, Trend. Pharmacol. 43 (2022)
629–640.
scholarship from Newcastle University funded by George Brown [25] M. Mannes, A. Dopler, O. Zolk, S.J. Lang, R. Halbgebauer, B. Höchsmann,
Endowment, George Henderson Endowment and Florence Kirkby A. Skerra, C.K. Braun, M. Huber-Lang, H. Schrezenmeier, C.Q. Schmidt,
Endowment. P.N.T. & S.M.M. acknowledge support by the European Complement inhibition at the level of C3 or C5: mechanistic reasons for ongoing
terminal pathway, Blood 137 (2021) 443–455.
Union’s Seventh Framework Programme (FP7-NMP-2012-Large-6)
[26] L. Zhang, Y. Dai, P. Huang, T.L. Saunders, D.A. Fox, J. Xu, F. Lin, Absence of
under the grant agreement no. 310337 (CosmoPHOS-nano Large-Scale complement component 3 does not prevent classical pathway-mediated hemolysis,
Project). D.S. acknowledges support by the National Institute of Blood Adv. 3 (2019) 1808–1814.
Health grants R01AI154959 and R01 CA257958. [27] E. Ispasanie, L. Muri, A. Schubart, C. Thorburn, N. Zamurovic, T. Holbro,
M. Kammüller, G. Pluschke, Alternative complement pathway inhibition does not
abrogate Meningococcal killing by serum of vaccinated individuals, Front.
References Immunol. 12 (2021), 747594.
[28] E.S. Reis, N. Berger, X. Wang, S. Koutsogiannaki, R.K. Doot, J.T. Gumas, P.
[1] N.S. Merle, S.E. Church, V. Fremeaux-Bacchi, L.T. Roumenina, Complement system G. Foukas, R.R.G. Resuello, J.V. Tuplano, D. Kukis, A.F. Tarantal, A.J. Young,
part I – molecular mechanisms of activation and regulation, Front. Immunol. 6 T. Kajikawa, A.M. Soulika, D.C. Mastellos, D. Yancopoulou, A.-R. Biglarnia,
(2015) 262. M. Huber-Lang, G. Hajihengallis, B. Nilsson, J.D. Lambris, Safety profile after
[2] P. Garred, A.J. Tenner, T.E. Mollnes, Therapeutic targeting of the complement prolonged C3 inhibition, Clin. Immunol. 197 (2018) 96–106.
system: from rare diseases to pandemics, Pharmacol. Rev. 73 (2021) 792–827. [29] A.M. Risitano, R. Notaro, C. Pascariello, M. Sica, L. del Vecchio, C.J. Horvath,
[3] H.I. Kenawy, I. Boral, A. Bevington, Complement-coagulation cross-talk: a M. Fridkis-Hareli, C. Selleri, M.A. Lindorfer, R.P. Taylor, L. Luzzatto, V.M. Holers,
potential mediator of the physiological activation of complement by low pH, Front. The complement receptor 2/factor H fusion protein TT30 protects paroxysmal
Immunol. 6 (2015) 215. nocturnal hemoglobinuria erythrocytes from complement-mediated hemolysis and
[4] C. Gaboriaud, M. Loevellec, V. Rossi, C. Dumestre-Pérard, N.M. Thielens, C3 fragment, Blood 119 (2012) 6307–6316.
Complement system and alrmin HMGB1 crosstalk: For better or worse, Front. [30] M. Fridkis-Hareli, M. Storek, E. Or, R. Altman, S. Katti, F. Sun, T. Peng, J. Hunter,
Immunol. 13 (2022), 869720. K. Johnson, Y. Wang, A.S. Lundberg, G. Mehta, N.K. Banda, V.M. Holers, The
[5] B.C. King, A.M. Blom, Intracellular complement: Evidence, definitions, human complement receptor 2 (CR2)/CR1 fusion protein TT32, a novel targeted
controversies, and solutions, Immunol. Rev. 313 (2023) 104–119. inhibitor of the classical and alternative pathway C3 convertases, prevents arthritis
[6] S.M. Moghimi, H.B. Haroon, A. Yaghmur, D. Simberg, P.N. Trohopoulos, in active immunization and passive transfer mouse models, Mol. Immunol. 105
Nanometer- and angstrom-scale characteristics that modulate complement (2019) 150–164.
responses to nanoparticles, J. Control. Release 351 (2022) 432–443. [31] E. Karnaukhova, C1-inhibitor: Structure, functional diversity and therapeutic
[7] S.M. Moghimi, H.B. Haroon, A. Yaghmur, A.C. Hunter, E. Papini, Z.S. Farhangrazi, development, Curr. Med. Chem. 29 (2022) 467–488.
D. Simberg, P.N. Trohopoulos, Perspectives on complement and phagocytic cell [32] V.P. Vu, G.B. Gifford, F. Chen, H. Benasutti, G. Wang, E.V. Groman, R. Scheinman,
responses to nanoparticles: From fundamentals to adverse reactions, J. Control. L. Saba, S.M. Moghimi, D. Simberg, Immunoglobulin deposition on biomolecule
Release 356 (2023) 115–129. corona determines complement opsonization efficiency of preclinical and clinical
[8] S.M. Moghimi, D. Simberg, Critical issues and pitfalls in serum and plasma nanoparticles, Nat. Nanotechnol. 14 (2019) 260–268.
handling for complement analysis in nanomedicine and bionanotechnology, Nano [33] Y. Li, S. Jacques, H. Gaikwad, G. Wang, N.K. Banda, V.M. Holers, R. Scheinman, S.
Today 44 (2022), 101479. Tomlinson, S.M. Moghimi, D. Simberg, Inhibition of acute complement responses
[9] L.C. Gjelstrup, J.D. Kaspersen, M.A. Behrens, J.S. Pedersen, S. Thiel, P. Kingshott, toward bolus injected nanoparticles by targeted short-circulating regulatory
C.L.P. Oliveira, N.M. Thielens, T. Vorup-Jensen, The role of nanometer-scaled proteins, Nat. Nanotechnol. (2023) https://www.nature.com/articles/s41565-02
ligand patterns in polyvalent binding by large mannan-binding lectin oligomers, 3-01514-z.
J. Immunol. 188 (2012) 1292–1306. [34] W.D. Ratnoff, D.T. Fearon, K.F. Austen, The role of antibody in the activation of the
[10] T.R. Kjaer, L. Jensen, A. Hansen, R. Dani, J.C. Jensenius, J. Dobó, P. Gál, S. Thiel, alternative pathway, in: H.J. Müller-Eberhard, P.A. Miescher, (eds.), Complement,
Oligomerization of mannan-binding lectin dictates binding properties and Springer, Berlin, Heidelberg, 1985, pp. 215–225, doi: 10.1007/978-3-642-82416-
complement activation, Exp. Immunol. 84 (2016) 12–19. 6_11.
[11] K. Peters, Physiology and pathology of the C3 amplification cycle: A retrospective, [35] F. Chen, G. Wang, J.I. Griffin, B. Brenneman, N.K. Banda, V.M. Holers, D.S. Backos,
Immunol. Rev. 313 (2023) 217–224. L. Wu, S.M. Moghimi, D. Simberg, Complement proteins bind to nanoparticle
protein corona and undergo dynamic exchange in vivo, Nat. Nanotechnol. 12
(2017) 387–393.

238
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

[36] S.M. Moghimi, D. Simberg, E. Papini, Z.S. Farhangrazi, Complement activation by [62] S.M. Moghimi, I. Hamad, T.L. Andresen, K. Jørgensen, J. Szebeni, Methylation of
drug carriers and particulate pharmaceuticals: principles, challenges and the phosphate oxygen moiety of phospholipid-methoxy(polyethylene glycol)
opportunities, Adv. Drug Deliv. Rev. 157 (2020) 83–95. conjugate prevents PEGylated liposome-mediated complement activation and
[37] A. Feinstein, E.A. Munn, Conformation of the free and antigen-bound IgM antibody anaphylatoxin production, FASEB J. 20 (2006) 2591–2593.
molecules, Nature 224 (1969) 1307–1309. [63] M. Pannuzzo, S. Esposito, L.-P. Wu, J. Key, S. Aryal, C. Celia, L. di Marzio, S.
[38] M.B. Pedersen, X. Zhou, E.K. Larsen, U.S. Sorensen, J. Kjems, J.V. Nygaard, J. M. Moghimi, P. Decuzzi, Overcoming nanoparticle-mediated complement
R. Nyengaard, R.L. Meyer, T. Boesen, T. Vorup-Jensen, Curvature of synthetic and activation by surface PEG pairing, Nano Lett. 20 (2020) 4312–4321.
natural surfaces is an important target feature in classical pathway complement [64] L.-P. Wu, M. Ficker, J.B. Christensen, D. Simberg, P.N. Trohopoulos, S.M. Moghimi,
activation, J. Immunol. 184 (2010) 1931–1945. Dendrimer end-terminal motif-dependent evasion of human complement and
[39] J. Strasser, R.N. de Jong, F.J. Beurskens, G. Wang, A.J.R. Heck, J. Schuurman, P.W. complement activation through IgM hitchhiking, Nat. Commun. 12 (2021) 4858.
H.I. Parren, P. Hinterdorfer, J. Preiner, Unravelling the macromolecular pathways [65] Y. Arima, M. Kawagoe, M. Toda, H. Iwata, Complement activation by polymers
of IgG oligomerization and complement activation on antigenic surfaces, Nano carrying hydroxyl groups, ACS Appl. Mater. Interfaces 1 (2009) 2400–2407.
Lett. 19 (2019) 4787–4796. [66] M. Toda, T. Kitazawa, I. Hirata, Y. Hirano, H. Iwata, Complement activation on
[40] B. Peschke, C.W. Keller, P. Weber, I. Quast, J.D. Lünemann, Fc-galactosylation of surfaces carrying amino groups, Biomaterials 29 (2008) 407–417.
human immunoglobulin gamma isotypes improves C1q binding and enhances [67] P.K. Maiti, T. Çagm, G. Wang, W.A. Goddard III, Structure of PAMAM dendrimers:
complement-dependent cytotoxicity, Front. Immunol. 8 (2017) 646. generations 1 through 11, Macromolecules 37 (2004) 6236–6254.
[41] A.C. Anselmo, S. Mitragotri, Nanoparticles in the clinic: An update, Bioeng. Transl. [68] G. Wang, J.I. Griffin, S. Inturi, B. Brenneman, N.K. Banda, V.M. Holers, S.
Med. 4 (2019) e10143. M. Moghimi, D. Simberg, In vitro and in vivo differences in murine third
[42] R.K. Thapa, J.O. Kim, Nanomedicine-based commercial formulations: current complement component (C3) opsonisation and macrophage/leukocyte responses to
developments and future prospects, J. Pharmaceutical Invest. 53 (2023) 19–33. antibody-functionalized iron oxide nanoworms, Front. Immunol. 8 (2017) 151.
[43] B.-M. Chen, T.-L. Cheng, S.R. Roffler, Polyethylene glycol immunogenicity: [69] P.W. Dempsey, M.E. Allison, S. Akkaraju, C.C. Goodnow, D.T. Fearon, C3d of
theoretical, clinical, and practical aspects of anti-polyethylene glycol antibodies, complement as a molecular adjuvant: bridging innate and acquired immunity,
ACS Nano 15 (2021) 14022–14048. Science 271 (1996) 348–350.
[44] Y. Bavli, B.-M. Chen, G. Gross, A. Hershko, K. Turjeman, S. Roffler, Y. Barenholz, [70] R.C. Rickert, Regulation of B lymphocyte activation by complement C3 and the B
Anti-PEG antibodies before and after a first dose of Comirnaty® (mRNA-LNP-based cell coreceptor complex, Curr. Opin. Immunol. 17 (2005) 237–243.
SARS-CoV-2 vaccine), J. Control. Release 354 (2023) 316–322. [71] A. Yalcindag, R. He, D. Laouini, H. Alnius, M. Carroll, H.C. Oettgen, R.S. Geha, The
[45] G. Guerrini, S. Gioria, A.V. Sauer, S. Lucchesi, F. Montagnani, G. Pastore, complement component C3 plays a critical role in both TH1 and TH2 responses to
A. Cibattini, D. Medaglini, L. Calzolai, Monitoring anti-PEG antibodies level upon antigen, J. Allergy Clin. Immunol. 117 (2006) 1455–1462.
repeated lipid nanoparticle-based COVID-19 vaccine administration, Int. J. Mol. [72] B. Li, A.Y. Jiang, I. Raji, C. Atyeo, T.M. Raimondo, A.G.R. Gordon, L.H. Rhym,
Sci. 23 (2022) 8838. T. Samad, C. Maclsaac, J. Witten, H. Mughal, T.M. Chicz, Y. Xu, R.P. McNamara,
[46] J.M. Carreño, G. Singh, J. Tcheou, K. Srivastava, C. Gleason, H. Murmatsu, S. Bhatia, G. Alter, R. Langer, D.G. Anderson, Enhancing the immunogenicity of
P. Desai, J.A. Aberg, R.L. Miller, N. Pardi, V. Simon, F. Krammer, PARIS study lipid-nanoparticles mRNA vaccines by adjuvanting the ionizable lipid and the
group, mRNA-1273 but not BNT162b2 induces antibodies against polyethylene mRNA, Nat. Biomed. Eng. (2023), https://doi.org/10.1038/s41551-023-01082-6.
glycol (PEG) contained in mRNA-based vaccine formulations, Vaccine 40 (2022) [73] K. Kovács, B. Mácsik-Valent, J. Matkó, Z. Bajtay, A. Erdei, Complement receptor
6114–6124. type 2 (CR2, CD21); coengagement with the B-cell receptor inhibits the activation,
[47] E. Chen, B.-M. Chen, Y.-C. Su, Y.-C. Chang, Y. Barenholz, S.R. Roffler, Premature proliferation, and antibody production of human B cells, Front. Immunol. 12
drug release from polyethylene glycol (PEG)-coated liposomal doxorubicin via (2021), 620427.
formation of the membrane attack complex, ACS Nano 14 (2020) 7808–7822. [74] S.M. Moghimi, Cancer nanomedicine and the complement system activation
[48] R.B. Sim, R. Wallis, Immune attack on nanoparticles, Nat. Nanotechnol. 6 (2011) paradigm: anaphylaxis and tumour growth, J. Control. Release 190 (2014)
80–81. 556–562.
[49] A. Miller, A. Phillips, J. Gor, R. Wallis, S.J. Perkins, Near-planar solution structures [75] H. Benasutti, G. Wang, V.P. Vu, R. Scheinman, E. Groman, L. Saba, D. Simberg,
of mannose-binding lectin oligomers provide insight on activation of lectin Variability of complement response toward preclinical and clinical nanocarriers in
pathway of complement, J. Biol. Chem. 287 (2012) 3930–3945. the general population, Bioconjug. Chem. 28 (2017) 2747–2755.
[50] C. Gaboriaud, P. Frachet, N.M. Thielens, G.J. Arlaud, The human C1q globular [76] M.M. Olcina, R. Kim, S. Melemenidis, E.E. Graves, A. Giaccia, The tumour
domain: structure and recognition of non-immune self ligands, Front. Immunol. 2 microenvironment links complement system dysregulation and hypoxic signaling,
(2012) 92. Br. J. Radiol. 92 (2019) 20180069.
[51] A.J. Bradley, D.E. Brooks, R. Norrie-Jones, D.V. Devine, C1q binding to liposomes [77] W.L. Olszewski, A. Engeset, H. Lukasiewicz, Immunoglobulins, complement and
is surface charge dependent and is inhibited by peptides consisting of residues lysozyme in leg lymph of normal men, Scand. J. Clin. Lab. Invest. 37 (1977)
14–26 of the human C1qA chain in a sequence independent manner, Biochim. 669–674.
Biophys. Acta-Biomemb. 1418 (1999) 19–30. [78] M.P. Reichhardt, S. Meri, SALSA: a regulator of the early steps of complement
[52] R. Tavano, L. Gabrielli, E. Lubian, C. Fedeli, S. Visentin, P.P. de Laureto, activation on mucosal surfaces, Front. Immunol. 7 (2016) 85.
G. Arrigoni, A. Geffner-Smith, F. Chen, D. Simberg, G. Morgese, E.M. Benetti, [79] M.P. Reichhardt, V. Loimaranta, S. Thiel, J. Finne, S. Meri, H. Java, The salivary
L. Wu, S.M. Moghimi, F. Mancin, E. Papini, C1q-mediated complement activation scavenger and agglutinin binds MBL and regulates the lectin pathway of
and C3 opsonization trigger recognition of stealth poly(2-methyl-2-oxazoline)- complement in solution and on surfaces, Front. Immunol. 3 (2012) 205.
coated silica nanoparticles by human phagocytes, ACS Nano 12 (2018) 5834–5847. [80] S.M. Riordan, C.J. McIver, D. Wakefield, P.C. Andreopoulos, V.M. Duncombe, T.
[53] C. Salvador-Morales, E. Flahaut, E. Sim, J. Sloan, M. Green, R.B. Sim, Complement D. Bolin, M.C. Thomas, Local and systemic complement activity in small intestinal
activation and protein adsorption by carbon nanotubes, Mol. Immunol. 43 (2006) bacterial overgrowth, Dig. Dis. Sci. 42 (1997) 1128–1136.
193–201. [81] U. Jain, A.R. Otley, J. Van Limbergen, A.W. Stadnyk, The complement system in
[54] W.L. Ling, A. Biro, I. Bally, P. Tacnet, A. Deniaud, E. Doris, P. Frachet, G. Schoehn, inflammatory bowel disease, Inflamm. Bowel Dis. 20 (2014) 1628–1637.
E. Pebay-Peyroula, G.J. Arlaud, Proteins of the innate immune system crystallize [82] M.D. Willcox, C.A. Morris, A. Thakur, R.A. Sack, J. Wickson, W. Boey, Complement
on carbon nanotubes but are not activated, ACS Nano 5 (2011) 730–737. and complement regulatory proteins in human tears, Invest. Ophthalmol. Vis. Sci.
[55] I. Hamad, O. Al-Hanbali, A.C. Hunter, K.J. Rutt, T.L. Andresen, S.M. Moghimi, 38 (1997) 1–8.
Distinct polymer architecture mediates switching of complement activation [83] Ø. Samuelsen, H.H. Haukland, H. Ulvatne, L.H. Vorland, Anti-complement effects
pathways at the nanosphere—serum interface: implications for stealth nanoparticle of lactoferrin-derived peptides, FEMS Immunol. Med. Microbiol. 41 (2004)
engineering, ACS Nano 4 (2010) 6629–6638. 141–148.
[56] J.N. Arnold, M.R. Wormald, D.M. Suter, C.M. Radcliffe, D.J. Harvey, R.A. Dwek, P. [84] E.E. West, C. Kemper, Complosome – the intracellular complement system, Nat.
M. Rudd, R.B. Sim, Human serum IgM glycosylation. Identification of glycoforms Rev. Nephrol. 19 (2023) 416–439.
that can bind to mannan-binding lectin, J. Biol. Chem. 280 (2005) 29080–29087. [85] M.K. Liszewski, M. Kolev, G. Le Friec, M. Leung, P.G. Bertram, A.F. Fara, M. Subias,
[57] Y.-Q. Wu, H. Qu, G. Sfyroera, A. Tzekou, B.K. Kay, B. Nilsson, K. Nilsson Ekdahl, M.C. Pickering, C. Drouet, S. Meri, T.P. Arstila, P.T. Pekkarinen, M. Ma, A. Cope,
D. Ricklin, J.d. Lambris, Protection of nonself surfaces from complement attack by T. Reinheckel, S. Rodriguez de Cordoba, B. Afzali, J.P. Atkinson, C. Kemper,
factor H-binding peptides: implications for therapeutic medicine, J. Immunol. 186 Intracellular complement activation sustains T cell homeostasis and mediates
(2011) 4269–4277. effector differentiation, Immunity 39 (2013) 1143–1157.
[58] S.R. Moore, S.S. Menon, C. Cortes, V.P. Ferreira, Hijacking factor H for complement [86] M. Elvington, M.K. Liszewski, P. Bertram, H.S. Kulkarni, J.P. Atkinson, A C3(H2O)
immune evasion, Front. Immunol. 12 (2021), 602277. recycling pathway is a component of the intracellular complement system, J. Clin.
[59] P.P. Wibroe, I.D.M. Azmi, C. Nilsson, A. Yaghmur, S.M. Moghimi, Citrem Invest. 127 (2017) 970–981.
modulates internal nanostructure of glyceryl monooleate dispersions and bypasses [87] P. Singh, C. Kemper, Complement, complosome, and complotype: A perspective,
complement activation: towards development of safe tunable intravenous lipid Eur. J. Immunol. (2023), https://doi.org/10.1002/eji.202250042.
nanocarriers, Nanomed. Nanotechnol. Biol. Med. 11 (2015) 1909–1914. [88] E.E. West, M. Kolev, C. Kemper, Complement and the Regulation of T cell
[60] B.S. Blaum, J.P. Hannan, A. Herbert, D. Kavanagh, D. Uhrin, T. Stehle, Structural responses, Annu. Rev. Immunol. 36 (2018) 309–338.
basis for sialic acid-mediated self-recognition by complement factor H, Nat. Chem. [89] S. Ndeupen, Z. Qin, S. Jacobsen, A. Bouteau, H. Estanbouli, B.Z. Igártó, The mRNA-
Biol. 11 (2015) 77–82. LNP platform’s lipid nanoparticle component used in preclinical vaccine studies is
[61] Z. Wang, E.D. Hood, J. Nong, J. Ding, O.A. Marcos-Contreras, P.M. Glassman, K. highly inflammatory, iScience 24 (2021), 103479.
M. Rubey, M. Zaleski, C.L. Espy, D. Gullipali, T. Miwa, V.R. Muzykantov, W.- [90] S. Tahtinen, A.-J. Tong, P. Himmels, J. Oh, A. Paler-Martinez, L. Kim, S. Wichner,
C. Song, J.W. Myerson, J.S. Brenner, Combating complement’s deleterious effects Y. Oei, M.J. McCarron, E.C. Freund, Z.A. Amir, C.C. de la Cruz, B. Haley,
on nanomedicine by conjugating complement regulatory proteins to nanoparticles, C. Blanchette, J.M. Schartner, W. Ye, M. Yadav, U. Shain, L. Delamarre, I. Mellman,
Adv. Mat. 34 (2022) e2107070.

239
H.B. Haroon et al. European Journal of Pharmaceutics and Biopharmaceutics 193 (2023) 227–240

IL-1 and IL-Ira are key regulators of the inflammatory response to RNA vaccines, [94] H. Gaikward, Y. Li, G. Wang, R. Li, S. Dai, C. Rester, R. Kedl, L. Saba, N.K. Banda, R.
Nat. Immunol. 23 (2022) 532–542. I. Scheinman, C. Patrick, K.M.G. Mallela, S.M. Moghimi, D. Simberg, Antibody-
[91] S.M. Moghimi, Pro-inflammatory concerns with lipid nanoparticles, Mol. Ther. 30 dependent complement responses toward SARS-CoV-2 receptor-binding domain
(2022) 2109–2110. immobilized on “pseudovirus-like” nanoparticles, ACS Nano 16 (2022) 8704–8715.
[92] J.C.H. Tam, S.R. Bidgood, W.A. McEwan, L.C. James, Intracellular sensing of [95] S.R. Barnum, C4a: An anaphylatoxin in name only, J. Innate Immun. 7 (2015)
complement C3 activates cell autonomous immunity, Science 345 (2014) 1256070. 333–339.
[93] C.L. Harris, M. Heurich, S.R. de Cordoba, B.P. Morgan, The complotype: dictating
risk for inflammation and infection, Trends Immunol. 33 (2012) 513–521.

240

You might also like