Abstract
Polyvalent interactions mediate the formation of higher-order macromolecular assemblies to improve the sensitivity, specificity, and temporal response of biological signals. In host defense, innate immune pathways recognize danger signals to alert host of insult or foreign invasion, while limiting aberrant activation from auto-immunity and cellular senescence. Of recent attention are the unique higher-order assemblies in the cGAS-STING pathway. Natural stimulation of cGAS enzymes by dsDNA induces phase separation and enzymatic activation for switchlike production of cGAMP. Subsequent binding of cGAMP to STING induces oligomerization of STING molecules, offering a scaffold for kinase assembly and signaling transduction. Additionally, the discovery of PC7A, a synthetic polymer which activates STING through a non-canonical biomolecular condensation, illustrates the engineering design of agonists by polyvalency principles. Herein, we discuss a mechanistic and functional comparison of natural and synthetic agonists to advance our understanding in STING signaling and highlight the principles of polyvalency in innate immune activation. The combination of exogenous cGAMP along with synthetic PC7A stimulation of STING offers a synergistic strategy in spatiotemporal orchestration of the immune milieu for a safe and effective immunotherapy against cancer.
Keywords: Polyvalency, Biomolecular Condensate, Phase Separation, Innate Immunity, cGAS, STING, Polymer
Graphical Abstract
Polyvalent interactions lead to the formation of macromolecular condensates. These polyvalent condensations are utilized by both natural and engineered systems to stimulate the cGAS-STING pathway for innate immune activation.
1. Introduction
Innate immunity is an ancient form of host defense primarily characterized by recognition of evolutionarily conserved pathogen structures [1]. Molecular pattern recognition receptors (PRRs) are invariant, germline-encoded protein receptors of the host, which recognize both non-self pathogen-associated molecular patterns (PAMPs) and danger-associated molecular patterns (DAMPs) from self, amplifying the signal of PAMP or DAMP recognition through signaling cascades [2–4]. The resulting amplification generally converges in secretion of cytokines by host cells, particularly leukocytes and antigen-presenting cells, activating adaptive immunity to generate specific effector responses towards PAMPs or DAMPs [5]. The signaling amplification and antigen presentation of non-self is highly contextual, depending on the anatomical and subcellular localization of antigen within the host [6, 7].
The presence of double-stranded deoxyribonucleic acid (dsDNA) in the cytosol is generally an aberrant phenomena in eukaryotic cells, indicative of infection or host disease [8]. The recognition of dsDNA by cyclic guanosine monophosphate-adenosine monophosphate (cyclic GMP-AMP, cGAMP) synthase (cGAS) amplifies an innate response with production of cGAMP as a secondary messenger [9, 10]. Evolutionary adaptations within this pathway utilize polyvalent interactions to limit recognition of self dsDNA while preserving sensitivity towards abnormal-self or non-self dsDNA [11–14]. Downstream, the stimulator of interferon genes (STING) is the adaptor protein of cGAMP. STING undergoes a substantial conformational change with subsequent oligomeric assembly upon binding with cGAMP [15, 16]. This facilitates the phosphorylation and recruitment of TANK-binding kinase 1 (TBK1) and interferon regulatory factor 3 (IRF3) molecules, leading to enhanced innate activation [17, 18]. Herein, we discuss how nature harnesses polyvalency in the regulation of cGAS-STING signaling and highlight the development of synthetic polymers, which induce biomolecular condensation, as an emerging therapeutic principle for cancer immunotherapy.
2. Higher-Order Assemblies in Cell Biology
Polyvalency is a characteristic of macromolecules which have the capacity to form multiple simultaneous interactions. Representative polyvalent interactions include receptor-ligand pairs each with multiple binding-units (Figure 1A). The quantity of interactions is the molecular valence state, defined as i, and influences system binding behavior. The binding strength of a typical monovalent interaction (i = 1) is quantifiable by the kinetic rate constants (kon and koff) and binding affinity (KB). For example, an immunoglobulin antigen-binding fragment (F(ab)) is a minimal structural component displaying affinity for a specific antigen [19]. The F(ab) molecule binds to antigen in a monovalent stoichiometry and is subjected to equilibrium association and dissociation. However, the incorporation of two F(ab) components into an immunoglobulin G (IgG) antibody shows bivalent binding (i = 2) to antigen (Figure 1A). Effectively, bivalency causes stronger total binding avidity between IgG and antigen through a decrease in the kinetic koff rate as well as an increased probabilistic binding geometry [20, 21].
Figure 1.
Polyvalency enables formation of higher-order structures with a switchlike response to input stimulus signal. (A) The affinity between a monovalent ligand and receptor pair (KBmono) is determined by the ratio of the association and dissociation rate constants (konmono and koffmono, respectively). When the same ligand receptor pair binds in a polyvalent fashion, the dissociation constant is decreased significantly (koffpoly ≪ koffmono). This change in molecular association is quantified by molecular avidity (KBpoly) betweem polyvalent pairs. The ratio between binding avidity and affinity constants is known as enhancement factor (β). The F(ab) portion of an antibody shows monovalent binding to antigen (Ag). IgG displays bivalent binding to antigen, which reduces kinetic dissociation and enhances avidity. IgM is a polyvalent molecule which has high affinity for antigen. (B) Scaffold proteins are multivalent molecules which physcially assemble and recruit monovalent client molecules. These complexes have many diverse functions in signal transduction pathways, including complex signal processing function. The MAPK-KSR signaling scaffold recruits several kinases after EGF binding and EGFR phosphorylation, enabling specific phosphorylation and transcription of growth programs. (C) Interactions between polyvalent molecules causes phase separation known as membraneless assemblies. The spontaneous nature of liquid droplets reduces the time and energy required to form higher-order structures as compared to static structures. A functional consideration of LLPS assemblies is the switchlike amplfication of output signals in response to an input stimulus. Interactions between Nephrin, Nck, and N-WASP on the cell membrane increase condensate dwell time through LLPS for rapid actin assembly and chemotaxis.
In a polyvalent system (i ≥ 3) avid binding describes the holistic kinetics and energetics through summation of individual affinities [22], each contributing energetically towards a higher-order assembly [23]. For example, immunoglobulin M (IgM) is a polymeric immunoglobulin molecule with decavalent binding to and high avidity for antigen due to its high valency and structural configuration [24]. Quantitatively, the enhancement factor, β, is the ratio between two binding affinities: KBpoly and KBmono, empirically comparing the change in system binding [22]. For a canonical IgM molecule which has specificity towards particular 2,4-dinitrophenyl ligands on the bacteriophage ϕX174, the association affinity is in the 1010−12 M−1 range as compared to 104−5 M−1 range for its respective monovalent component, defining a β factor of 106−7-fold [25]. Functionally, IgM shows utility in early activation of immune complement during infection, displaying the influence of high avidity binding on biological response [26]. In addition, avid IgM macromolecules may form a supramolecular assembly with its complementary antigen, inducing subsequent aggregation or assembly [27]. This higher-order assembly of polyvalent molecules is a general organizing principle, endowing additional function beyond stoichiometric assembly, including modulating enzymatic kinetics [28–30], enabling cytoskeletal arrangement and assembly [31–33], and regulation of gene transcription [34–36].
2.1. Protein Assembly in Signal Transduction Pathways
Signal transduction describes a sequential process of protein activation in response to ligand binding of a protein receptor, converging in a specific biological response. Intermediate proteins in the signaling cascade influences threshold behavior, signal amplification, and spatial regulation of the signaling response [37]. Particularly, high valency scaffold proteins improve specificity in signal transduction through higher-order assembly (Figure 1B). Scaffolds interact with other modular proteins in a polyvalent configuration or induce stoichiometric recruitment and binding of low valency client molecules [38]. Specifically, the mitogen activated protein kinase (MAPK) signaling pathway canonically utilizes scaffold assemblies to link multiple kinases in a single cascade [39]. In mammalian cells, the kinase suppressor of Ras (KSR) is a scaffold adaptor which localizes Ras-GTPase with other signaling kinases including Raf, MAPK, and extracellular signal-regulated kinase (ERK) [40, 41]. Along this pathway, diverse regulatory elements and behavior facilitate emergent behavior such as pathway branching, allostery, and feedback for specific convergence upon cellular growth and proliferation programs [42, 43]. These amplification events can display ultrasensitive or switchlike behavior, a hallmark of positive cooperativity [44].
2.2. Liquid-Liquid Phase Separation of Proteins
Molecular self-organization and assembly occurs through out-of-equilibrium thermodynamic processes, which require significant free energy investment for formation [45, 46]. To minimize the cost of self-assembly and optimize self-organization, biological systems spatiotemporally enrich molecules within transient, dynamic structures, through a process known as liquid-liquid phase separation (LLPS) [47]. These multi-component assemblies form by polyvalent weak, non-covalent interactions [48, 49] and undergo liquid-liquid demixing through spontaneous immiscibility events characteristic of polymer-polymer separation [50, 51]. In addition to conserving free energy, intracellular LLPS entities conserve essential and limited cellular components, such as lipid bilayers, by separating in a membraneless capacity [52].
Functionally, condensates optimize enzymatic reaction conditions by influencing kinetics, specificity, or inhibition [53]. Specifically, the process of actin nucleation and polymerization in podocyte cells of the kidney utilizes high-avidity, low-affinity polyvalent interactions between several components: phosphorylated tyrosine (pTyr) residues in the transmembrane receptor Nephrin, the Src homology 2 (SH2) domains on the adaptor protein Nck, and the proline-rich motifs (PRMs) on its ligand N-WASP (Figure 1C) [54]. The system undergoes membrane-tethered LLPS after chemoattractant molecules bind to Nephrin, increasing the dwell time of the Nephrin-Nck-N-WASP protein assembly and rapidly polymerizing actin via ARP-2/3 complex recruitment for a sensitive chemotactic response [54]. Significantly, additional biological systems exploit LLPS to enhance the sensitivity of a particular input signal to achieve a switchlike or positively cooperative response [48, 55, 56]. Comparatively, while both LLPS and scaffolding serve as organizing assemblies for amplifying molecular signals within the cell, a key distinction is the interaction between polyvalent scaffolds and monovalent clients as compared to the heterotypic interaction between two or more scaffold proteins. While polyvalent-monovalent assemblies regulate sensitivity through mass transport of client proteins and availability of free scaffold binding sites [38], polyvalent-polyvalent assemblies enhance sensitivity dependent on the local concentrations of its constitutive polyvalent components [57]. Thus, both assemblies depend on the relative concentration of polyvalent components, with heterotypic polyvalent pairs having increasing propensity to phase separate, enhancing transduction sensitivity.
3. Natural Polyvalency in the cGAS-STING Pathway
The canonical cGAS-STING signaling cascade is an innate immune pathway for recognition of cytosolic dsDNA as a danger signal. Upon sequence-independent binding of dsDNA, cGAS enzymatically synthesizes the secondary messenger, cGAMP, for downstream binding and activation of STING [9, 10, 58]. STING is a endoplasmic reticulum (ER) transmembrane receptor protein which undergoes a conformational change and translocates from the ER to the Golgi apparatus after activation [59]. Activated STING recruits the adaptor protein TBK1 [17] for phosphorylation of the transcription factor interferon regulatory factor 3 (IRF3), resulting in the expression and secretion of type I interferon (IFN) [60]. In general, the cGAS-STING pathway is an essential immune defense mechanism against DNA-containing pathogens as well as an alert system for cancer or environmental insult. The fundamental nature of the cGAS-STING pathway necessitates tight regulation to prevent stimulation by host genomic and mitochondrial DNA, while allowing rapid signal amplification in response to aberrant cytosolic dsDNA [12, 61, 62]. Evolutionary adaptions in the cGAS-STING pathway utilize polyvalent interactions for formation of higher-order assemblies, achieving a specific and rapid response.
3.1. dsDNA-Induced Phase Condensation of cGAS Enzyme
The recognition step in the cGAS-STING pathway involves cytosolic dsDNA binding to cGAS in a sequence-independent manner. Due to the lack of sequence-specificity for dsDNA and the abundant presence of dsDNA within cells, cGAS activation obtains specificity, in part, through dsDNA-length dependent activation [63, 64]. Structurally, cGAS contains an intrinsically disordered and positively charged N-terminal domain as well as an ordered C-terminal nucleotidyltransferase (NTD) domain [65, 66]. The dsDNA molecules bind to both domains of cGAS in a 2:2 dsDNA to cGAS stoichiometry. Subsequent conformational changes to cGAS enable enzymatic production of the secondary messenger cGAMP [67, 68]. Critically, dsDNA binding to cGAS induces foci formation, which facilitate cGAMP production [9, 69]. Recent studies reveal these foci are liquid-like droplets formed through LLPS and are mediated by polyvalent ionic interactions between the negatively-charged phosphate backbone of dsDNA and the positively-charged N- or C-terminus of cGAS (Figure 2A) [11]. While dsDNA activates cGAS independently of genetic sequence, the length of dsDNA segments impacts the size and frequency of LLPS as well as cGAMP production [11]. This understanding suggests the valence state of the dsDNA polymer is critical for activation of the cGAS-STING pathway, wherein a minimal threshold requires 45 base pairs in the dsDNA for substantial enzymatic production of cGAMP by cGAS [11, 63, 64]. Functionally, the cGAS-DNA biomolecular condensate concentrates cGAS enzyme and AMP/GMP substrates in droplets for enhanced enzymatic production of cGAMP [11, 70]. Furthermore, cellular sequestering of cGAS in a monomeric configuration within the nucleosome prevents aberrant cGAS activation by inhibiting polyvalent condensation, suggesting the higher order assembly is regulated by the polyvalent interactions between dsDNA and cGAMP [61, 71–73]. Overall, this phenomenon enables a switchlike response in cGAMP production, amplifying the initial dsDNA stimuli for further activation of downstream STING.
Figure 2.
Higher-order assemblies in the natural cGAS-STING pathway. (A) Cytosolic dsDNA enters the cytoplasm upon infection by DNA-containing pathogen or by damage to host. This dsDNA binds with cGAS to form a biomolecular condensate through LLPS. The function of this assembly is to concentrate cGAS with the substrates (GMP, AMP) to catalytically synthesize cGAMP. (B) cGAMP binds to STING on the ER membrane, which causes lid-closing of the STING dimer and oligomerization to form a higher-order scaffold. The whole complex translocates to the Golgi apparatus through ERGIC where TBK1 binds and phoshophrylates STING residues as well as other TBK1 molecules. This platform recruits IRF3 which is also phosphorylated. Activated IRF3 dimerizes and translocates to the nucleus, initiating transcription of type I IFNs.
3.2. Oligomerization of STING Dimers for Downstream Immune Activation
STING is an ER membrane-bound protein consisting of four transmembrane helices, a dimeric cytosolic ligand-binding domain (LBD), and a C-terminal tail (CTT) [74]. Functionally, STING (also known as TMEM173, MITA, ERIS, and MPYS) is essential in production of IFN-β in response to pathogen infection [74–77]. The IFN response is mediated by cyclic dinucleotide (CDN) ligand binding to STING [78] wherein the eukaryotic natural ligand of STING is 2’3’-cGAMP [58, 79]. CDN binding in a V-shaped pocket at the dimer interface induces STING activation by promoting ‘lid-closing’ conformational changes to the cytosolic LBD [15, 80, 81]. The conformational change causes release of the CTT, exposing an oligomerization interface on the LBD [15]. STING molecules begin side-by-side packing on the ER membrane and assemble into a higher-order oligomeric structure [16], which migrates to the ER-Golgi intermediate compartment (ERGIC) and is observable as puncta in the perinuclear region (Figure 2B) [82]. The oligomeric cGAMP-STING assembly further recruits TBK1 to a conserved motif in the CTT of STING [18]. Although there is constitutive interaction between STING and TBK1 in the absence of cGAMP, STING-TBK1 binding is enhanced by cGAMP-activation through STING oligomerization [17]. TBK1 molecules cluster on oligomeric STING and phosphorylate the activation loop of an adjacent TBK1 molecule in a trans-autophosphorylation mechanism, which is normally inaccessible in a cis configuration [83, 84]. Further, phosphorylation of STING by TBK1 must occur on a neighboring STING dimer in the assembly due to geometric constraints [17]. TBK1 phosphorylation of adjacent TBK1 molecules in trans as well as neighboring STING dimers can only occur in the context of an oligomeric cGAMP-STING-TBK1 higher order assembly, endowing additional functional specificity on the pathway.
The STING-TBK1 complex recruits the transcription factor IRF3, acting as a signaling platform through mediating STING-TBK1-IRF3 interactions concurrently [85]. TBK1 phosphorylates STING at an evolutionarily conserved, IRF3-binding motif, pLxIS (p is a hydrophilic residue, x is any residue, and S is the phosphorylation residue), which is also present on other adaptor molecules: mitochondrial antiviral-signaling protein (MAVS), TIR-domain-containing adapter-inducing interferon-β (TRIF), and TLR adaptor interacting with SLC15A4 on the lysosome (TASL) [60, 86]. IRF3 traffics and binds to the phosphorylated STING-TBK1 assembly and is subsequently phosphorylated by TBK1 at an identical pLxIS motif on IRF3 itself [87]. Phosphorylated IRF3 homodimerizes, translocates to the nucleus, and engages with an interferon-stimulated response element (ISRE) on DNA for transcription of IFN-β and CXCL10 [88, 89]. Overall, the higher-order assembly of STING oligomers, induced by cGAMP, enables a sensitive and specific mechanism for activation of TBK1 and IRF3 through polyvalent structural avidity. Further, regulation and inhibitory structures of cGAMP-STING-TBK1 occur through similar, but excessive, polyvalent interactions including an LLPS transition to a gel-like state [90] or complete inhibition of the oligomerization interface [16].
4. Synthetic Polyvalent STING Agonists with Prolonged Innate Activation
4.1. An Amphiphilic Block Copolymer, PC7A, Directly Activates STING
Therapeutic stimulation of STING by natural or synthetic agonists represents an attractive target in cancer immunotherapy, due to the link between innate immunity stimulation and a cell-mediated adaptive response [91]. Nanoparticles represent alluring therapeutic or vaccination platforms because the ability to encapsulate small molecular cargo as well as size-dependent accumulation within myeloid cells, which have high capacity to secrete soluble cytokines [92]. Therefore, we synthesized a library of ultra-pH-sensitive (UPS) polymers with characteristic linear or cyclic tertiary amine sidechains to screen for a candidate STING agonist delivery platform (Figure 3A) [93]. UPS polymers self-assemble into nanoparticles in aqueous environments and respond to tunable pH-thresholds with remarkably strong positive cooperativity [94]. The nanoparticles have a poly(ethylene glycol) (PEG) shell and a uniform (diameter ≈ 30 nm) micelle structure, showing suitable biocompatibility properties. Upon exposure to pH below a particular threshold, the micelles spontaneously disassemble into a positively-charged unimer state. Previously work demonstrates the applicability of this platform in delineating acidotic tumor margins for image-guided surgical resection [95] or for PET imaging [96]. Furthermore, UPS responses to endo-lysosomal pH enables digitization [97] and perturbation [98] of the endosomal maturation process. Screening of the UPS library revealed polymers with specific cyclic sidechains disrupt endosomal membranes and allows cytosolic delivery of cargo such as adjuvants and antigens [99].
Figure 3.
Non-canonical activation of STING by a synethic PC7A polymer. (A) A library of UPS polymers with distinct sidechains and pKa show positive cooperativity in response to pH as an input signal. PC7A shows polyvalency-dependent binding of and condensation with STING dimers. Adapted from [100]. (B) cGAMP is formulated within the core of PC7A nanoparticles as a dual STING agonist. Treatment of monocytes shows distinct spatiotemporal IFNβ in response to cGAMP or PC7A NP alone. (C) A hydrolyzable polycarbonate backbone is substituted for the PMMA backbone to establish a biodegradable, polyvalent STING agonist. After subcutaneous injection, a nodule forms as a result of local inflammation and wound healing response. This nodule remains persistent in PC7A treated mice but vanishes in PSC7A-treated mice over time. Adapted from [106].
Encapsulation of tumor-associated antigens or tumor neoantigens in UPS nanoparticles enabled formation of a micellar nanoparticle vaccine [100]. Screening of the UPS library by a T cell killing assay identified direct immune activation by a polymer with a cyclic, seven-membered amine ring (PC7A). The immune response was characterized by cytotoxic T lymphocyte (CTL) antigen-specific killing and a high titer of antigen-specific IgG1 and IgG2c. Interestingly, other linear and cyclic amine polymers did not initiate strong immune activity or T cell killing. Mechanistic investigation revealed strong STING-IFNα/β-dependence but independence of myeloid differentiation primary response protein (MyD88)/TRIF and MAVS pathways. Subcutaneous administration of the nanoparticle vaccine in mice targeted draining lymph nodes and initiated an immune response against multiple tumor models. The resulting tumor growth suppression outperformed treatment by established adjuvants such as cytosine and guanine with phosphodiester backbone oligonucleotides (CpG ODNs) and polyinosinic: polycytidylic acid (poly(I:C)). These results show PC7A has intrinsic immune adjuvant effects, a unique property of a polymeric nanoparticle.
STING activation by PC7A was compared to CDNs such as 2’3’-cGAMP [101]. While both PC7A and cGAMP demonstrated the intracellular re-distribution of STING molecules and phosphorylation of downstream proteins (TBK1 and IRF3) in mouse embryonic fibroblasts, the temporal activation of STING by PC7A was slower than cGAMP (Figure 3B). The kinetic differences in STING activation suggested a differing mechanism in STING activation by PC7A. Mechanistically, PC7A binds to a non-competitive surface site on the α5 helix of STING that is distinct from the cGAMP pocket. Mutation of two acidic residues (E296A/D297A) sufficiently abolishes polymer binding, while cGAMP-resistant STING variants (e.g., R232H) retains PC7A activation. As a result, PC7A polymer serves as a supramolecular scaffold and directly engages polyvalent interactions to multimerize STING molecules for immune activation (Figure 3A) [101].
The formation and size of STING-PC7A puncta was dependent on polymer repeating unit, suggesting a critical role in the valency of PC7A sidechains [101]. Downstream transcription of interferon stimulated genes (ISGs) such as CXCL10, was also dependent on the number of repeating units. STING activity reached a maximum when the repeating unit was at 70 (Figure 3A). Interestingly, longer polymers (e.g., >70 repeating units) induced higher phase condensation levels, but STING activity was reduced. We attribute the attenuated STING activity to the weaker signaling capacity of oversized condensates with excessive cross-linking and poor molecular dynamics. This hypothesis was supported by the lower reversibility and slower fluorescence recovery rate of STING-GFP after photobleaching. A recent study also observed a cGAMP dose-dependent STING condensation, wherein a high dose of agonist treatment induced a gel-like transition and suppressed innate immune signaling [90]. Together, these results emphasize the critical role of polymeric binding valence and suggests precise control of phase separation achieves optimal immunological outcome.
The seven-member tertiary-amine ring of PC7A (azepanyl moiety) activates STING even in a monomeric configuration. Lipids containing a heterocyclic amino head group, formulated within mRNA-containing lipid nanoparticles, demonstrate dose-dependent STING-activation [102]. The resulting mRNA vaccination platform shows potent DC maturation and activation for presentation of translated tumor associated antigen to CTLs, enabling an effective antitumor immune response against established sold tumors in mice. While the cyclic amino lipids activate STING, it is unclear if the lipids share the same binding site as PC7A or bind within the CDN binding pocket. Molecular dynamics of the heterocyclic lipid and the STNG ligand binding domain suggest association within the CDN binding pocket [102], which is not indicated in the mutagenic study of STING binding to PC7A as discussed earlier [101].
4.2. cGAMP-loaded PC7A Nanoparticles Synergize for Enhanced STING Activation
Since PC7A and cGAMP activate STING through distinct binding moieties and spatiotemporal profiles, encapsulation of cGAMP within PC7A nanoparticles (polySTING) synergizes dual activity in STING-mediated immune activation. Intratumoral treatment of tumor-bearing mice with polySTING showed robust tumor growth inhibition and prolonged survival as compared to cGAMP and PC7A alone [101]. In polySTING treatment, cGAMP initiates an initial burst in IFNβ secretion whereas PC7A achieves a sustained activity (Figure 3B). This dual-activation strategy prolongs the innate immune activation. Moreover, incubation of fresh surgically-resected human tissue with polySTING shows elevated STING-activation and type I IFN response in sentinel lymph node specimens as well as in several types of resected human tumor tissues. In particular, CD45+ myeloid cells are primarily responsible for the type I IFN response after polySTING treatment in patient samples, indicating the STING-mediated immunomodulation of leukocytes in the tumor microenvironment by nanoparticles.
4.3. A Biodegradable Block Copolymer Improves the Therapeutic Window of STING Activity
Therapeutic administration of STING agonists for cancer immunotherapy generally requires repeat dosing [103]. A concern with PC7A is increased dose accumulation after repeated administration. While the poly(methyl methacrylate) (PMMA) backbone of the PC7A polymer is relatively inert and non-toxic, accumulation of polymeric PC7A sidechain agonists could cause adverse effects and dose-limiting toxicity [104, 105]. Therefore, we rationally designed and synthesized a biodegradable polymer with similar sidechains as PC7A, known as PSC7A (Figure 3C) [106]. This degradable polymer is composed of a polycarbonate backbone, which spontaneously hydrolyzes over time, causing degradation into small molecule monomer sidechains or oligomeric material. An in vitro hydrolysis assay showed a degradation half-time of the polymer to be 12 or 27 days at pH 6.5 or pH 7.4, respectively. Similar to PC7A, PSC7A also directly binds to the STING CTD. Moreover, encapsulation of tumor antigen in the PSC7A nanoparticle formed a stable biodegradable nanoparticle vaccine. The acute response of PSC7A vaccine towards STING activation and CTL response in mice is of equal magnitude compared to PC7A. However, PSC7A shows significant reduction in systemic inflammatory cytokines after acute treatment. Over time, PC7A induces a persistent inflammatory nodule on the skin at the injection site. This nodule was initially observed with PSC7A, but the nodule reduced in size and was completely diminished after 40 days, demonstrating a favorable safety profile and increasing the therapeutic dosing window (Figure 3C).
5. Comparing STING Agonists
5.1. Small Molecule STING Agonists
Traditional STING agonists are small molecule CDN analogs [107]. In addition to 2’3’-cGAMP, several naturally occurring CDNs include cyclic di-adenosine monophosphate (c-di-AMP) [108], cyclic di-guanine monophosphate (c-di-GMP) [78], and 3’3’-cGAMP [109]. However, degradation by phosphodiesterases, anionic hydrophilicity prohibiting passive membrane transport, and poor pharmacokinetics, limit the therapeutic application of natural CDNs [104, 110–112]. Moreover, agonists initially considered as promising therapeutic candidates, such as DMXAA [113] and CMA [114], only activate murine but not human STING. Synthetic CDN analogs, such as thiophosphate c-di-GMP analogs [115], ADU-S100 [104], IAC-8779 and IACS-8803 [116], CL656 [117], and E7766 [118] were developed to overcome these limitations (Table 1). Several clinical trials are ongoing with intra-tumoral administration of CDNs for cancer immunotherapy, which are the focus of other reviews [119, 120]. Generally, CDNs show low biological activity and limited therapeutic efficacy, necessitating the further development of STING-activating therapies.
Table 1.
Small Molecule STING Agonists
STING Agonist | Type | Mechanism | Taxonomy | Reference |
---|---|---|---|---|
2’3’-cGAMP | CDN | Cell Intrinsic | Metezoan | [15, 58] |
3’3’-cGAMP | CDN | Cell Intrinsic | Bacterial | [109] |
c-di-AMP | CDN | Cell Intrinsic | Bacterial | [108] |
c-di-GMP | CDN | Cell Intrinsic | Bacterial | [78] |
DMXAA | Non-CDN | Lid-Closed | Mouse | [113] |
CMA | Non-CDN | Lid-Closed | Mouse | [114] |
Dithio c-di-GMP | CDN | Lid-Closed | Human | [115] |
ADU-S100 | CDN | Lid-Closed | Human | [104] |
IACS-8779 | CDN | Lid-Closed | Human | [116] |
IACS-8803 | CDN | Lid-Closed | Human | [116] |
CL656 | CDN | Lid-Closed | Human | [117] |
E7766 | CDN | Lid-Closed | Human | [118] |
SR-717 | Non-CDN | Lid-Closed | Human | [121] |
MSA-2 | Non-CDN | Lid-Closed | Human | [122] |
di-ABZI | Non-CDN | Lid-Open | Human | [124, 125] |
A list of synthetic small molecules which direcetly bind and activate STING. ‘Type’ indicates the class of drugs: cyclic dinucleotides (CDN) or CDN mimetics (non-CDN). ‘Mechanism’ details the inherent occurance of CDN’s (cell intrinsic) or the structural configuration of STING (lid-closed/lid-open) after synthetic agonist binding. ‘Taxonomy’ specifies either the biological kingdom for cell-intrinsic agonists or the biological species which are activated by synthetic agonists.
Non-nucleotide CDNs exist as an alternative to CDNs (Table 1). Representative examples include the small molecules SR-717 [121] and MSA-2 [122], which activate STING with a similar ‘lid-closing’ conformation as cGAMP [123]. In contrast, the class of amidobenzimidazole (ABZI) non-CDN agonists activates STING in a ‘lid-open’ manner but within the CDN-binding pocket [124, 125]. Further molecular and structural biology investigations are necessary to delineate the biological consequences of activating STING in its open or closed conformations. Critically, the administration route of these compounds is not limited to intratumoral injection, enabling systemic administration either intravenously or orally. However, concerns still exist regarding the precise delivery of these compounds for potent activation of STING in relevant dendritic cell populations within tumor or secondary immune organs.
5.2. Engineered STING Agonist Delivery Platforms
Nano-engineering offers unique advantages by simultaneously enabling potent and safe on-demand delivery of small molecule STING agonists (Table 2). As discussed previously, several polymeric micelles with heterocyclic amines show inherent STING activity through direct binding to a non-competitive surface on the ligand binding domain of STING [100–102, 106]. The polymer agonists display polyvalent condensation of STING molecules, inducing prolonged innate immune responses [101], but it is still unknown if this mechanism is specific to PC7A polymer or to all heterocyclic amine molecules with the azepanyl moiety. Further, several classes of nanoparticles with hydrophilic cores such as polymersomes [126–129], liposomes [130–134], extracellular vesicles [135, 136], and mesoporous silica [137, 138] show stable loading of anionic CDNs. Both polymersomes and liposomes show efficient cytosolic delivery of CDN agonists, typically through stimuli-responsive moieties, and have potential for high loading capacity of anionic CDN payload. In particular, unique synergy and convergence on IFNβ secretion between Mn2+ ions and CDN-induced STING activation shows potential for dual-delivery of non-redundant payloads within an engineered liposome formulation [134]. Extracellular vesicles, such as exosomes or viral vectors, show intrigue in regards to feasibility for development but uniformity and yield of production remain concerns.
Table 2.
Engineered STING Agonists for Immunotherapy
Class | Unique Component | Agonist Cargo | Route | Key Finding | Reference |
---|---|---|---|---|---|
Polymeric Micelle | PEG-b-PC7A | N/A (Intrinsic) | s.c. | Polyvalent Condensation | [100, 170] |
PEG-b-PC7A | cGAMP | i.t. | Dual Kinetic STING Activity | [101] | |
PEG-b-PSC7A | N/A (Intrinsic) | s.c. | Biodegradable Polymeric Agonist | [106] | |
Polymersome | PEG-b-(DEAEMA-co-BMA) | cGAMP | i.t./i.v. | Endosome Disruption | [126–128] |
PEG-b-P(FcMA-co-PEMA) | diABZI | i.v. | Hydroxyl Radical Sensitization | [129] | |
Liposome | Heterocyclic Amine Lipid | N/A (Intrinsic) | s.c. | Direct STING Binding | [102] |
Cationic Lipid | cGAMP/c-di-GMP | i.t./s.c. | Traditional Delivery Systems | [130, 131] | |
Pulmonary Surfactants | cGAMP | i.h. | Synergy with Influenza Vaccine | [133] | |
Mn2+ Chelated Lipid | c-di-AMP | s.c./i.v. | Metal-Ion Modulation | [134] | |
Extracellular Vesicle | Virus-like Particles | cGAMP | N/A | Endogenous cGAMP Packaging | [135] |
HEK293T Exosome | cGAMP | i.t. | Passive Diffusion into Cytosol | [136] | |
Mesoporous Silica | PEGylated Silica | di-GMP | i.t. | In situ Vaccination | [137] |
Labile Silica Matrix | c-di-AMP | i.t. | Biodegradable Low Density Silica | [138] | |
Microparticle | Acetylated-Dextran | cGAMP | i.t./i.m. | Traditional Delivery Systems | [139, 140] |
PLGA | cGAMP | i.t. | Pulsatile cGAMP Release | [141] | |
Hydrogel | Hyaluronic Acid | c-di-AM(PS)2 | Local | Gradual Release Post-Surgery | [142] |
Peptide-Drug Nanotube Complex | c-di-AMP | i.t. | Gradual Release of Agonist | [143] | |
STING Protein | Exogenous STINGΔTM Protein | cGAMP-STING ΔTM | s.c. | Self-Assembly of cGAMP-STING ΔTM Dimer into Tetramers | [144] |
Omomyc Protein | cGAMP-STING ΔTM | i.t. | Cell-Penetrating Fusion Protein for Cytosolic Delivery | [145] | |
Antibody Drug Conjugate | Immunosynthen | Unknown | i.v. | Targeted Delivery of Agonist via Antibody | [146] |
A list of engineered systems which activate STING either through intrinsic mechanisms or delivery of STING-activating CDN. ‘Class’ refers to the overarching type of STING agonist delivery system. ‘Unique Component’ refers to the distinguishing feature of the system. ‘Cargo’ refers to the specific STING-activating molecule loaded within the system. ‘Route’ refers to the administration route for therapeutic utility. s.c. is subcutaneous; i.t. is intratumoral; i.v. is intravenous; i.h. is inhalation; i.m. is intramuscular. ‘Key Finding’ defines how the system activates STING or offers utility as a delivery platform.
The sustained delivery of STING agonists involving microparticles [139–141] or hydrogels [142, 143] are developed to avoid repeated administration of the agonists, while enabling an optimal time window for immune activities to overcome tumor recurrence or metastasis. The inherent size of these biomaterials is less appealing in regards to acute biological clearance and clinical utility but remain selections for investigation because of favorable long-term biodegradable properties as well as tractability and feasibility in formulation. Additionally, the delivery of STING protein lacking the transmembrane domain and in complexation with cGAMP (cGAMP-STINGΔTM) is a novel approach for STING activation by triggering the signaling pathway independently from endogenous STING [144, 145]. It is unclear if these therapeutics induce the natural polyvalent assembly of cGAMP-STING-TBK1 on the ER membrane or, conversely, assemble in the cytosol for TBK1/IRF3 recruitment. Moreover, antibody-drug conjugates represent a strategy for systemic targeting of STING agonists to cell types of interest and prolonging the drug half-life as compared to small molecule circulation [146]. A major challenge herein is the development of a covalent, yet labile, linkage between antibody and agonist for cytosolic delivery and STING activation.
A key distinction between the classes of STING activators is the polyvalent condensation induced by PC7A and PSC7A, which enables sustained STING activity over 6–48 h [101]. In comparison, cGAMP activates STING with downstream functions peaking at 6 h after treatment. The distinctions in temporal kinetics of STING activation need to be investigated further in the contexts of immunological activation and priming of adaptive responses. However, the synergy between lid-closing activation and polymer-induced condensation of STING enables dual ‘burst’ and ‘sustain” temporal profiles, which proves effective in the immunotherapy against solid tumors [101].
6. Conclusion and Future Outlook
Polyvalent interactions drive many of biology’s organizing principles and functions. Polyvalency strengthens higher-order assemblies by reducing the dissociation rate and enhancing the overall association avidity of molecular recognition groups. The functional requirements imparted on the innate immune system utilize polyvalent condensation for enhanced signal processing and amplification. The interactions in the cGAS-STING pathway amplify presence of cytosolic dsDNA through cGAMP secondary messenger signaling, terminating in the transcription and secretion of type I IFN. How this pathway prevents aberrant activation is appreciated by the understanding of its regulatory polyvalent structural assemblies. Further, a synthetic polymer, PC7A, activates STING in a novel manner, inducing polyvalent condensation and prolonging type I IFN response. This therapeutic strategy synergizes with cGAMP for a dual-activating STING agonist platform wherein cGAMP provides a ‘burst’ and PC7A enables a ‘sustain’ in type I IFN secretion. Additionally, a biodegradable backbone of a similar polymer, PSC7A, directly activates STING but improves the therapeutic window of repeat dosing as it reduces vaccine-induced nodule formation over time.
Discrimination between self and non-self is highly dependent on the contextual presence of molecular patterns [147]. Generally, the balance in recognition of microbial non-self, missing self, and altered self against normal self requires continual exposure to self-molecules and robust inhibitory pathways to prevent aberrant auto-immunity [148]. Immune surveillance mechanisms must have robust recognition of microbial non-self and altered self-patterns (Figure 4A). A proposed model for non-self antigenic discrimination in the presence of abundant self-antigen is the ‘discontinuity theory’, which proposes the temporal aspect of antigen exposure determines the resulting immune response [149, 150]. An enduring effector response against a structurally abnormal stimulus (such as in the case of cancer immunotherapy) requires intermittent exposure with clear removal of stimulus to prevent overexposure [151]. This is contrasted to the persistent presence of innate stimulus, which causes immune tolerance [152], hypersensitivity [153], or exhaustion [154]. Moreover, the ‘trained innate immunity’ model may induce amplified adaptive response upon repeat innate stimulus triggers [155]. Through both immunogenic discontinuity as well as trained innate immunity stimulation, we theorize a novel agonist design principle herein.
Figure 4.
A design principle for exploiting polyvalency in innate immune activation. (A) Immune receptors are constantly engaged in many interactions, particularly from low-affinitiy self molecules. However, PAMPs or DAMPs bind specific PRRs with high affinity, amplifying signals downstream. With a specific time-dependent pattern of innate activation (intermittent, sustained, or persistent; solid line, light green background), an adaptive response (dotted line, salmon color background) initiates a durable effector response or behaves in a hyporesponsive or tolerogenic manner. (B) A polyvalent macromolecule displays binding avidity (KBavidity) to innate immune receptors (light blue). We propose a biodegradable design which disintegrates into monomeric components over time. The affinity between monomeric ligand and the immune receptor is greatly reduced, allowing for ligand-receptor dissociation. This is reflected by an ON/OFF stimulus response curve as a function of time, which fits with an ‘intermittment dosing’ regimen for optimal effector immune response.
We propose a polyvalent agonist design to introduce intermittent immune activation in the context of cancer immunotherapy (Figure 4B). This principle relies on immunostimulatory molecules which have relatively low monovalent affinity for intended immune receptors [156]. Methods to measure binding affinity include isothermal titration calorimetry [157], fluorescence anisotropy [158], and surface plasmon resonance [159]. We suggest these immunostimulatory molecules serve as monomeric building blocks with functional groups for polymerization. Examples of polymerization include reversible addition–fragmentation chain transfer (RAFT) [160], atom transfer radical polymerization (ATRP) [161], or ring opening polymerization (ROP) [162]. Considerations for additional functions include amphiphilic block copolymer segments for stability and controlled self-assembly [163], pH-responsive or functional moieties for controlled release of cargo [94, 164], and membrane destabilizing groups for cytosolic delivery of nanomaterial [165].
The binding avidity of polymeric material to the immune receptors shows enhancement as compared to monomeric ligands. To prevent hyperactivation or tolerance, we propose biodegradable polymeric backbones, such as polycarbonate (degradation rate constant ≈ 10−10 s−1) [106], polyester (10−8 –10−9 s−1) [141] [166], or polyphosphoester (10−6 s−1) [167] as well as many other backbones [168, 169]. Intermittent dosing of biodegradable material is possible without persistent hyperactivation and resulting immune-related toxicity due to a reduction in biological activity upon degradation [106]. Ultimately, combining the natural and engineering means of activating the cGAS-STING pathway presents a potent design strategy for rational development of novel innate immunostimulatory agonists. Further work will refine this model to understand the specific molecular interactions between polyvalent ligand and innate immune receptors for safe and efficacious cancer immunotherapy.
Highlights.
Polyvalency enriches biomacromolecules in higher-order assemblies with amplified signaling responses to small changes in stimulus.
Cytosolic dsDNA and cGAS enzyme undergo polyvalent liquid-liquid phase separation, promoting cGAMP production.
cGAMP binding to STING causes oligomerization for recruitment and activation of downstream signaling molecules in a cGAMP-STING-TBK1 complex.
PC7A, a synthetic polymer, induces polyvalent STING condensation and prolongs a type I interferon response as compared to cGAMP activation.
Polyvalent condensation is a design principle in natural and engineered systems for immune activation of the cGAS-STING pathway.
7. Acknowledgements
This work was supported by grants from the National Institutes of Health (U54 CA244719, U01 CA218422, and R01 CA216839) as well as the Mendelson-Young Endowment in Cancer Therapeutics. Z.T.B. is supported by the Ruth L. Kirschstein Predoctoral National Research Service Award from the National Institutes of Health (F31 CA247152). J.G. dedicates this article to George M. Whitesides, who made seminal contributions to chemical polyvalency in molecular recognition and therapeutic development.
Abbreviations:
- AMP
adenosine monophosphate
- CDN
cyclic dinucleotide
- cGAMP
cyclic guanosine monophosphate-adenosine monophosphate
- cGAS
cyclic guanosine monophosphate-adenosine monophosphate synthase
- CTD
C-terminal domain
- CTL
cytotoxic T lymphocyte
- CTT
C-terminal tail
- DAMP
danger-associated molecular pattern
- dsDNA
double-stranded deoxyribonucleic acid
- ER
endoplasmic reticulum
- ERGIC
ER-Golgi intermediate compartment
- F(ab)
antigen binding fragment
- GMP
guanosine monophosphate
- IFN
interferon
- IgG
immunoglobulin G
- IgM
immunoglobulin M
- IRF3
interferon regulator factor 3
- ISG
interferon stimulated gene
- KSR
kinase suppressor of Ras
- LBD
ligand binding domain
- LLPS
liquid-liquid phase separation
- MAPK
mitogen activated protein kinase
- NTD
nucleotidyltransferase
- PAMP
pattern-associated molecular pattern
- PC7A
polymer with a cyclic, seven-membered amine ring
- PEG
polyethylene glycol
- PMMA
poly(methyl methacrylate)
- PRR
pattern recognition receptor
- STING
stimulator of interferon genes
- TBK1
TANK-binding kinase 1
- UPS
ultra-pH-sensitive
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- [1].Medzhitov R, Janeway C, Innate Immunity, New England Journal of Medicine 343(5) (2000) 338–344. [DOI] [PubMed] [Google Scholar]
- [2].Kawai T, Akira S, TLR signaling, Seminars in Immunology 19(1) (2007) 24–32. [DOI] [PubMed] [Google Scholar]
- [3].Loo Y-M, Gale M, Immune Signaling by RIG-I-like Receptors, Immunity 34(5) (2011) 680–692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [4].Heim VJ, Stafford CA, Nachbur U, NOD Signaling and Cell Death, Frontiers in Cell and Developmental Biology 7(208) (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [5].Medzhitov R, Janeway CA Jr, Innate immune recognition and control of adaptive immune responses, Seminars in Immunology 10(5) (1998) 351–353. [DOI] [PubMed] [Google Scholar]
- [6].Raz E, Organ-specific regulation of innate immunity, Nature Immunology 8(1) (2007) 3–4. [DOI] [PubMed] [Google Scholar]
- [7].Kagan JC, Defining the subcellular sites of innate immune signal transduction, Trends in Immunology 33(9) (2012) 442–448. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [8].Vance RE, Cytosolic DNA Sensing: The Field Narrows, Immunity 45(2) (2016) 227–228. [DOI] [PubMed] [Google Scholar]
- [9].Sun L, Wu J, Du F, Chen X, Chen ZJ, Cyclic GMP-AMP Synthase Is a Cytosolic DNA Sensor That Activates the Type I Interferon Pathway, Science 339(6121) (2013) 786–791. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [10].Wu J, Sun L, Chen X, Du F, Shi H, Chen C, Chen ZJ, Cyclic GMP-AMP Is an Endogenous Second Messenger in Innate Immune Signaling by Cytosolic DNA, Science 339(6121) (2013) 826–830. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [11].Du M, Chen ZJ, DNA-induced liquid phase condensation of cGAS activates innate immune signaling, Science 361(6403) (2018) 704–709. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [12].Li T, Huang T, Du M, Chen X, Du F, Ren J, Chen ZJ, Phosphorylation and chromatin tethering prevent cGAS activation during mitosis, Science (2021) eabc5386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [13].Zhou W, Mohr L, Maciejowski J, Kranzusch PJ, cGAS phase separation inhibits TREX1-mediated DNA degradation and enhances cytosolic DNA sensing, Molecular Cell 81(4) (2021) 739–755.e7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [14].Zhong L, Hu M-M, Bian L-J, Liu Y, Chen Q, Shu H-B, Phosphorylation of cGAS by CDK1 impairs self-DNA sensing in mitosis, Cell Discovery 6(1) (2020) 26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [15].Shang G, Zhang C, Chen ZJ, Bai X.-c., Zhang X, Cryo-EM structures of STING reveal its mechanism of activation by cyclic GMP–AMP, Nature 567(7748) (2019) 389–393. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [16].Ergun SL, Fernandez D, Weiss TM, Li L, STING Polymer Structure Reveals Mechanisms for Activation, Hyperactivation, and Inhibition, Cell 178(2) (2019) 290–301.e10. [DOI] [PubMed] [Google Scholar]
- [17].Zhang C, Shang G, Gui X, Zhang X, Bai X.-c., Chen ZJ, Structural basis of STING binding with and phosphorylation by TBK1, Nature 567(7748) (2019) 394–398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [18].Zhao B, Du F, Xu P, Shu C, Sankaran B, Bell SL, Liu M, Lei Y, Gao X, Fu X, Zhu F, Liu Y, Laganowsky A, Zheng X, Ji J-Y, West AP, Watson RO, Li P, A conserved PLPLRT/SD motif of STING mediates the recruitment and activation of TBK1, Nature 569(7758) (2019) 718–722. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [19].Holliger P, Hudson PJ, Engineered antibody fragments and the rise of single domains, Nature Biotechnology 23(9) (2005) 1126–1136. [DOI] [PubMed] [Google Scholar]
- [20].Crothers DM, Metzger H, The influence of polyvalency on the binding properties of antibodies, Immunochemistry 9(3) (1972) 341–57. [DOI] [PubMed] [Google Scholar]
- [21].Kaufman EN, Jain RK, Effect of Bivalent Interaction upon Apparent Antibody Affinity: Experimental Confirmation of Theory Using Fluorescence Photobleaching and Implications for Antibody Binding Assays, Cancer Research 52(15) (1992) 4157–4167. [PubMed] [Google Scholar]
- [22].Mammen M, Choi S-K, Whitesides GM, Polyvalent Interactions in Biological Systems: Implications for Design and Use of Multivalent Ligands and Inhibitors, Angewandte Chemie International Edition 37(20) (1998) 2754–2794. [DOI] [PubMed] [Google Scholar]
- [23].Fasting C, Schalley CA, Weber M, Seitz O, Hecht S, Koksch B, Dernedde J, Graf C, Knapp E-W, Haag R, Multivalency as a Chemical Organization and Action Principle, Angewandte Chemie International Edition 51(42) (2012) 10472–10498. [DOI] [PubMed] [Google Scholar]
- [24].Chesebro B, Bloth B, Svehag SE, The ultrastructure of normal and pathological IgM immunoglobulins, J Exp Med 127(3) (1968) 399–410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [25].Hornick CL, Karush F, Antibody affinity—III the role of multivalence, Immunochemistry 9(3) (1972) 325–340. [DOI] [PubMed] [Google Scholar]
- [26].Boes M, Role of natural and immune IgM antibodies in immune responses, Molecular Immunology 37(18) (2000) 1141–1149. [DOI] [PubMed] [Google Scholar]
- [27].Pauling L, Campbell DH, Pressman D, THE NATURE OF THE FORCES BETWEEN ANTIGEN AND ANTIBODY AND OF THE PRECIPITATION REACTION, Physiological Reviews 23(3) (1943) 203–219. [Google Scholar]
- [28].Wheeldon I, Minteer SD, Banta S, Barton SC, Atanassov P, Sigman M, Substrate channelling as an approach to cascade reactions, Nature Chemistry 8(4) (2016) 299–309. [DOI] [PubMed] [Google Scholar]
- [29].Castellana M, Wilson MZ, Xu Y, Joshi P, Cristea IM, Rabinowitz JD, Gitai Z, Wingreen NS, Enzyme clustering accelerates processing of intermediates through metabolic channeling, Nature Biotechnology 32(10) (2014) 1011–1018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [30].Kleiman Laura B., Maiwald T, Conzelmann H, Lauffenburger Douglas A., Sorger Peter K., Rapid Phospho-Turnover by Receptor Tyrosine Kinases Impacts Downstream Signaling and Drug Binding, Molecular Cell 43(5) (2011) 723–737. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [31].Schek III HT, Gardner MK, Cheng J, Odde DJ, Hunt AJ, Microtubule assembly dynamics at the nanoscale, Current Biology 17(17) (2007) 1445–1455. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [32].Herrmann H, Aebi U, Intermediate filaments: structure and assembly, Cold Spring Harbor Perspectives in Biology 8(11) (2016) a018242. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [33].Rottner K, Faix J, Bogdan S, Linder S, Kerkhoff E, Actin assembly mechanisms at a glance, Journal of Cell Science 130(20) (2017) 3427–3435. [DOI] [PubMed] [Google Scholar]
- [34].Van Dyke MW, Roeder RG, Sawadogo M, Physical analysis of transcription preinitiation complex assembly on a class II gene promoter, Science 241(4871) (1988) 1335–1338. [DOI] [PubMed] [Google Scholar]
- [35].Thanos D, Maniatis T, Virus induction of human IFNβ gene expression requires the assembly of an enhanceosome, Cell 83(7) (1995) 1091–1100. [DOI] [PubMed] [Google Scholar]
- [36].Yang L, Gal J, Chen J, Zhu H, Self-assembled FUS binds active chromatin and regulates gene transcription, Proceedings of the National Academy of Sciences 111(50) (2014) 17809–17814. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [37].Wu H, Higher-Order Assemblies in a New Paradigm of Signal Transduction, Cell 153(2) (2013) 287–292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [38].Banani SF, Rice AM, Peeples WB, Lin Y, Jain S, Parker R, Rosen MK, Compositional Control of Phase-Separated Cellular Bodies, Cell 166(3) (2016) 651–663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [39].Chol K-Y, Satterberg B, Lyons DM, Elion EA, Ste5 tethers multiple protein kinases in the MAP kinase cascade required for mating in S. cerevisiae, Cell 78(3) (1994) 499–512. [DOI] [PubMed] [Google Scholar]
- [40].Nguyen A, Burack WR, Stock JL, Kortum R, Chaika OV, Afkarian M, Muller WJ, Murphy KM, Morrison DK, Lewis RE, McNeish J, Shaw AS, Kinase suppressor of Ras (KSR) is a scaffold which facilitates mitogen-activated protein kinase activation in vivo, Mol Cell Biol 22(9) (2002) 3035–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [41].Witzel F, Maddison L, Blüthgen N, How scaffolds shape MAPK signaling: what we know and opportunities for systems approaches, Frontiers in Physiology 3(475) (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [42].Good MC, Zalatan JG, Lim WA, Scaffold Proteins: Hubs for Controlling the Flow of Cellular Information, Science 332(6030) (2011) 680–686. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [43].Bhattacharyya RP, Reményi A, Good MC, Bashor CJ, Falick AM, Lim WA, The Ste5 Scaffold Allosterically Modulates Signaling Output of the Yeast Mating Pathway, Science 311(5762) (2006) 822–826. [DOI] [PubMed] [Google Scholar]
- [44].Malleshaiah MK, Shahrezaei V, Swain PS, Michnick SW, The scaffold protein Ste5 directly controls a switch-like mating decision in yeast, Nature 465(7294) (2010) 101–105. [DOI] [PubMed] [Google Scholar]
- [45].Schrödinger E, What is Life? The Physical Aspect of the Living Cell, Cambridge University Press; 1944. [Google Scholar]
- [46].Karsenti E, Self-organization in cell biology: a brief history, Nature Reviews Molecular Cell Biology 9(3) (2008) 255–262. [DOI] [PubMed] [Google Scholar]
- [47].Alberti S, Phase separation in biology, Current Biology 27(20) (2017) R1097–R1102. [DOI] [PubMed] [Google Scholar]
- [48].Li P, Banjade S, Cheng H-C, Kim S, Chen B, Guo L, Llaguno M, Hollingsworth JV, King DS, Banani SF, Russo PS, Jiang Q-X, Nixon BT, Rosen MK, Phase transitions in the assembly of multivalent signalling proteins, Nature 483(7389) (2012) 336–340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [49].Riback JA, Zhu L, Ferrolino MC, Tolbert M, Mitrea DM, Sanders DW, Wei M-T, Kriwacki RW, Brangwynne CP, Composition-dependent thermodynamics of intracellular phase separation, Nature 581(7807) (2020) 209–214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [50].Hyman AA, Weber CA, Jülicher F, Liquid-Liquid Phase Separation in Biology, Annual Review of Cell and Developmental Biology 30(1) (2014) 39–58. [DOI] [PubMed] [Google Scholar]
- [51].De Gennes P-G, Gennes P-G, Scaling concepts in polymer physics, Cornell university press; 1979. [Google Scholar]
- [52].Brangwynne CP, Eckmann CR, Courson DS, Rybarska A, Hoege C, Gharakhani J, Jülicher F, Hyman AA, Germline P granules are liquid droplets that localize by controlled dissolution/condensation, Science 324(5935) (2009) 1729–1732. [DOI] [PubMed] [Google Scholar]
- [53].Banani SF, Lee HO, Hyman AA, Rosen MK, Biomolecular condensates: organizers of cellular biochemistry, Nature Reviews Molecular Cell Biology 18(5) (2017) 285–298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [54].Case LB, Zhang X, Ditlev JA, Rosen MK, Stoichiometry controls activity of phase-separated clusters of actin signaling proteins, Science 363(6431) (2019) 1093–1097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [55].Yoo H, Triandafillou C, Drummond DA, Cellular sensing by phase separation: Using the process, not just the products, Journal of Biological Chemistry 294(18) (2019) 7151–7159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [56].Su X, Ditlev JA, Hui E, Xing W, Banjade S, Okrut J, King DS, Taunton J, Rosen MK, Vale RD, Phase separation of signaling molecules promotes T cell receptor signal transduction, Science 352(6285) (2016) 595–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [57].Farahi N, Lazar T, Wodak SJ, Tompa P, Pancsa R, Integration of Data from Liquid–Liquid Phase Separation Databases Highlights Concentration and Dosage Sensitivity of LLPS Drivers, International Journal of Molecular Sciences 22(6) (2021) 3017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [58].Gao P, Ascano M, Wu Y, Barchet W, Gaffney Barbara L., Zillinger T, Serganov Artem A., Liu Y, Jones Roger A., Hartmann G, Tuschl T, Patel Dinshaw J., Cyclic [G(2′,5′)pA(3′,5′)p] Is the Metazoan Second Messenger Produced by DNA-Activated Cyclic GMP-AMP Synthase, Cell 153(5) (2013) 1094–1107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [59].Ishikawa H, Ma Z, Barber GN, STING regulates intracellular DNA-mediated, type I interferon-dependent innate immunity, Nature 461(7265) (2009) 788–792. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [60].Liu S, Cai X, Wu J, Cong Q, Chen X, Li T, Du F, Ren J, Wu Y-T, Grishin NV, Chen ZJ, Phosphorylation of innate immune adaptor proteins MAVS, STING, and TRIF induces IRF3 activation, Science 347(6227) (2015) aaa2630. [DOI] [PubMed] [Google Scholar]
- [61].Pathare GR, Decout A, Glück S, Cavadini S, Makasheva K, Hovius R, Kempf G, Weiss J, Kozicka Z, Guey B, Melenec P, Fierz B, Thomä NH, Ablasser A, Structural mechanism of cGAS inhibition by the nucleosome, Nature 587(7835) (2020) 668–672. [DOI] [PubMed] [Google Scholar]
- [62].Zhao B, Xu P, Rowlett CM, Jing T, Shinde O, Lei Y, West AP, Liu WR, Li P, The molecular basis of tight nuclear tethering and inactivation of cGAS, Nature 587(7835) (2020) 673–677. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [63].Zhou W, Whiteley AT, de Oliveira Mann CC, Morehouse BR, Nowak RP, Fischer ES, Gray NS, Mekalanos JJ, Kranzusch PJ, Structure of the Human cGAS–DNA Complex Reveals Enhanced Control of Immune Surveillance, Cell 174(2) (2018) 300–311.e11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [64].Andreeva L, Hiller B, Kostrewa D, Lässig C, de Oliveira Mann CC, Jan Drexler D, Maiser A, Gaidt M, Leonhardt H, Hornung V, Hopfner K-P, cGAS senses long and HMGB/TFAM-bound U-turn DNA by forming protein–DNA ladders, Nature 549(7672) (2017) 394–398. [DOI] [PubMed] [Google Scholar]
- [65].Kranzusch Philip J., Lee Amy S.-Y., Berger James M., Doudna Jennifer A, Structure of Human cGAS Reveals a Conserved Family of Second-Messenger Enzymes in Innate Immunity, Cell Reports 3(5) (2013) 1362–1368. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [66].Kato K, Ishii R, Goto E, Ishitani R, Tokunaga F, Nureki O, Structural and Functional Analyses of DNA-Sensing and Immune Activation by Human cGAS, PLOS ONE 8(10) (2013) e76983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [67].Civril F, Deimling T, de Oliveira Mann CC, Ablasser A, Moldt M, Witte G, Hornung V, Hopfner K-P, Structural mechanism of cytosolic DNA sensing by cGAS, Nature 498(7454) (2013) 332–337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [68].Zhang X, Wu J, Du F, Xu H, Sun L, Chen Z, Brautigam Chad A., Zhang X, Chen Zhijian J., The Cytosolic DNA Sensor cGAS Forms an Oligomeric Complex with DNA and Undergoes Switch-like Conformational Changes in the Activation Loop, Cell Reports 6(3) (2014) 421–430. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [69].Li X, Shu C, Yi G, Chaton Catherine T., Shelton Catherine L., Diao J, Zuo X, Kao CC, Herr Andrew B., Li P, Cyclic GMP-AMP Synthase Is Activated by Double-Stranded DNA-Induced Oligomerization, Immunity 39(6) (2013) 1019–1031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [70].Zhao M, Xia T, Xing J-Q, Yin L-H, Li X-W, Pan J, Liu J-Y, Sun L-M, Wang M, Li T, Mao J, Han Q-Y, Xue W, Cai H, Wang K, Xu X, Li T, He K, Wang N, Li A-L, Zhou T, Zhang X-M, Li W-H, Li T, The stress granule protein G3BP1 promotes pre-condensation of cGAS to allow rapid responses to DNA, EMBO reports n/a(n/a) e53166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [71].Boyer JA, Spangler CJ, Strauss JD, Cesmat AP, Liu P, McGinty RK, Zhang Q, Structural basis of nucleosome-dependent cGAS inhibition, Science 370(6515) (2020) 450–454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [72].Kujirai T, Zierhut C, Takizawa Y, Kim R, Negishi L, Uruma N, Hirai S, Funabiki H, Kurumizaka H, Structural basis for the inhibition of cGAS by nucleosomes, Science 370(6515) (2020) 455–458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [73].Cao D, Han X, Fan X, Xu R-M, Zhang X, Structural basis for nucleosome-mediated inhibition of cGAS activity, Cell Research 30(12) (2020) 1088–1097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [74].Ishikawa H, Barber GN, STING is an endoplasmic reticulum adaptor that facilitates innate immune signalling, Nature 455(7213) (2008) 674–678. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [75].Zhong B, Yang Y, Li S, Wang Y-Y, Li Y, Diao F, Lei C, He X, Zhang L, Tien P, Shu H-B, The Adaptor Protein MITA Links Virus-Sensing Receptors to IRF3 Transcription Factor Activation, Immunity 29(4) (2008) 538–550. [DOI] [PubMed] [Google Scholar]
- [76].Jin L, Waterman PM, Jonscher KR, Short CM, Reisdorph NA, Cambier JC, MPYS, a Novel Membrane Tetraspanner, Is Associated with Major Histocompatibility Complex Class II and Mediates Transduction of Apoptotic Signals, Mol Cell Biol 28(16) (2008) 5014–5026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [77].Sun W, Li Y, Chen L, Chen H, You F, Zhou X, Zhou Y, Zhai Z, Chen D, Jiang Z, ERIS, an endoplasmic reticulum IFN stimulator, activates innate immune signaling through dimerization, Proceedings of the National Academy of Sciences 106(21) (2009) 8653–8658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [78].Burdette DL, Monroe KM, Sotelo-Troha K, Iwig JS, Eckert B, Hyodo M, Hayakawa Y, Vance RE, STING is a direct innate immune sensor of cyclic di-GMP, Nature 478(7370) (2011) 515–518. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [79].Shi H, Wu J, Chen ZJ, Chen C, Molecular basis for the specific recognition of the metazoan cyclic GMP-AMP by the innate immune adaptor protein STING, Proceedings of the National Academy of Sciences 112(29) (2015) 8947–8952. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [80].Huang Y-H, Liu X-Y, Du X-X, Jiang Z-F, Su X-D, The structural basis for the sensing and binding of cyclic di-GMP by STING, Nature Structural & Molecular Biology 19(7) (2012) 728–730. [DOI] [PubMed] [Google Scholar]
- [81].Shang G, Zhu D, Li N, Zhang J, Zhu C, Lu D, Liu C, Yu Q, Zhao Y, Xu S, Gu L, Crystal structures of STING protein reveal basis for recognition of cyclic di-GMP, Nature Structural & Molecular Biology 19(7) (2012) 725–727. [DOI] [PubMed] [Google Scholar]
- [82].Saitoh T, Fujita N, Hayashi T, Takahara K, Satoh T, Lee H, Matsunaga K, Kageyama S, Omori H, Noda T, Yamamoto N, Kawai T, Ishii K, Takeuchi O, Yoshimori T, Akira S, Atg9a controls dsDNA-driven dynamic translocation of STING and the innate immune response, Proceedings of the National Academy of Sciences 106(49) (2009) 20842–20846. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [83].Larabi A, Devos Juliette M., Ng S-L, Nanao Max H., Round A, Maniatis T, Panne D, Crystal Structure and Mechanism of Activation of TANK-Binding Kinase 1, Cell Reports 3(3) (2013) 734–746. [DOI] [PubMed] [Google Scholar]
- [84].Tu D, Zhu Z, Zhou Alicia Y., Yun C.-h., Lee K-E, Toms Angela V., Li Y, Dunn Gavin P., Chan E, Thai T, Yang S, Ficarro Scott B., Marto Jarrod A., Jeon H, Hahn William C., Barbie David A., Eck Michael J., Structure and Ubiquitination-Dependent Activation of TANK-Binding Kinase 1, Cell Reports 3(3) (2013) 747–758. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [85].Tanaka Y, Chen ZJ, STING Specifies IRF3 Phosphorylation by TBK1 in the Cytosolic DNA Signaling Pathway, Science Signaling 5(214) (2012) ra20–ra20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [86].Heinz LX, Lee J, Kapoor U, Kartnig F, Sedlyarov V, Papakostas K, César-Razquin A, Essletzbichler P, Goldmann U, Stefanovic A, Bigenzahn JW, Scorzoni S, Pizzagalli MD, Bensimon A, Müller AC, King FJ, Li J, Girardi E, Mbow ML, Whitehurst CE, Rebsamen M, Superti-Furga G, TASL is the SLC15A4-associated adaptor for IRF5 activation by TLR7–9, Nature 581(7808) (2020) 316–322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [87].Zhao B, Shu C, Gao X, Sankaran B, Du F, Shelton CL, Herr AB, Ji J-Y, Li P, Structural basis for concerted recruitment and activation of IRF-3 by innate immune adaptor proteins, Proceedings of the National Academy of Sciences 113(24) (2016) E3403–E3412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [88].Kawai T, Takeuchi O, Fujita T, Inoue J.-i., Mühlradt PF, Sato S, Hoshino K, Akira S, Lipopolysaccharide Stimulates the MyD88-Independent Pathway and Results in Activation of IFN-Regulatory Factor 3 and the Expression of a Subset of Lipopolysaccharide-Inducible Genes, The Journal of Immunology 167(10) (2001) 5887–5894. [DOI] [PubMed] [Google Scholar]
- [89].Doyle SE, Vaidya SA, O’Connell R, Dadgostar H, Dempsey PW, Wu T-T, Rao G, Sun R, Haberland ME, Modlin RL, Cheng G, IRF3 Mediates a TLR3/TLR4-Specific Antiviral Gene Program, Immunity 17(3) (2002) 251–263. [DOI] [PubMed] [Google Scholar]
- [90].Yu X, Zhang L, Shen J, Zhai Y, Jiang Q, Yi M, Deng X, Ruan Z, Fang R, Chen Z, Ning X, Jiang Z, The STING phase-separator suppresses innate immune signalling, Nature Cell Biology 23(4) (2021) 330–340. [DOI] [PubMed] [Google Scholar]
- [91].Barber GN, STING: infection, inflammation and cancer, Nature Reviews Immunology 15(12) (2015) 760–770. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [92].Irvine DJ, Dane EL, Enhancing cancer immunotherapy with nanomedicine, Nature Reviews Immunology 20(5) (2020) 321–334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [93].Ma X, Wang Y, Zhao T, Li Y, Su L-C, Wang Z, Huang G, Sumer BD, Gao J, Ultra-pH-Sensitive Nanoprobe Library with Broad pH Tunability and Fluorescence Emissions, Journal of the American Chemical Society 136(31) (2014) 11085–11092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [94].Li Y, Zhao T, Wang C, Lin Z, Huang G, Sumer BD, Gao J, Molecular basis of cooperativity in pH-triggered supramolecular self-assembly, Nature Communications 7(1) (2016) 13214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [95].Zhao T, Huang G, Li Y, Yang S, Ramezani S, Lin Z, Wang Y, Ma X, Zeng Z, Luo M, de Boer E, Xie XJ, Thibodeaux J, Brekken RA, Sun X, Sumer BD, Gao J, A Transistor-like pH Nanoprobe for Tumour Detection and Image-guided Surgery, Nat Biomed Eng 1 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [96].Huang G, Zhao T, Wang C, Nham K, Xiong Y, Gao X, Wang Y, Hao G, Ge WP, Sun X, Sumer BD, Gao J, PET imaging of occult tumours by temporal integration of tumour-acidosis signals from pH-sensitive (64)Cu-labelled polymers, Nat Biomed Eng 4(3) (2020) 314–324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [97].Wang Y, Wang C, Li Y, Huang G, Zhao T, Ma X, Wang Z, Sumer BD, White MA, Gao J, Digitization of Endocytic pH by Hybrid Ultra-pH-Sensitive Nanoprobes at Single-Organelle Resolution, Advanced Materials 29(1) (2017) 1603794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [98].Wang C, Wang Y, Li Y, Bodemann B, Zhao T, Ma X, Huang G, Hu Z, DeBerardinis RJ, White MA, Gao J, A nanobuffer reporter library for fine-scale imaging and perturbation of endocytic organelles, Nature Communications 6(1) (2015) 8524. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [99].Wang Z, Luo M, Mao C, Wei Q, Zhao T, Li Y, Huang G, Gao J, A Redox-Activatable Fluorescent Sensor for the High-Throughput Quantification of Cytosolic Delivery of Macromolecules, Angewandte Chemie International Edition 56(5) (2017) 1319–1323. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [100].Luo M, Wang H, Wang Z, Cai H, Lu Z, Li Y, Du M, Huang G, Wang C, Chen X, Porembka MR, Lea J, Frankel AE, Fu Y-X, Chen ZJ, Gao J, A STING-activating nanovaccine for cancer immunotherapy, Nature Nanotechnology 12(7) (2017) 648–654. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [101].Li S, Luo M, Wang Z, Feng Q, Wilhelm J, Wang X, Li W, Wang J, Cholka A, Fu Y.-x., Sumer BD, Yu H, Gao J, Prolonged activation of innate immune pathways by a polyvalent STING agonist, Nature Biomedical Engineering (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [102].Miao L, Li L, Huang Y, Delcassian D, Chahal J, Han J, Shi Y, Sadtler K, Gao W, Lin J, Doloff JC, Langer R, Anderson DG, Delivery of mRNA vaccines with heterocyclic lipids increases anti-tumor efficacy by STING-mediated immune cell activation, Nature Biotechnology 37(10) (2019) 1174–1185. [DOI] [PubMed] [Google Scholar]
- [103].Corrales L, Glickman Laura H., McWhirter Sarah M., Kanne David B., Sivick Kelsey E., Katibah George E., Woo S-R, Lemmens E, Banda T, Leong Justin J., Metchette K, Dubensky Thomas W., Gajewski Thomas F., Direct Activation of STING in the Tumor Microenvironment Leads to Potent and Systemic Tumor Regression and Immunity, Cell Reports 11(7) (2015) 1018–1030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [104].Meric-Bernstam F, Sandhu SK, Hamid O, Spreafico A, Kasper S, Dummer R, Shimizu T, Steeghs N, Lewis N, Talluto CC, Dolan S, Bean A, Brown R, Trujillo D, Nair N, Luke JJ, Phase Ib study of MIW815 (ADU-S100) in combination with spartalizumab (PDR001) in patients (pts) with advanced/metastatic solid tumors or lymphomas, Journal of Clinical Oncology 37(15_suppl) (2019) 2507–2507. [Google Scholar]
- [105].Harrington KJ, Brody J, Ingham M, Strauss J, Cemerski S, Wang M, Tse A, Khilnani A, Marabelle A, Golan T, LBA15 - Preliminary results of the first-in-human (FIH) study of MK-1454, an agonist of stimulator of interferon genes (STING), as monotherapy or in combination with pembrolizumab (pembro) in patients with advanced solid tumors or lymphomas, Annals of Oncology 29 (2018) viii712. [Google Scholar]
- [106].Wang X, Wilhelm J, Li W, Li S, Wang Z, Huang G, Wang J, Tang H, Khorsandi S, Sun Z, Evers B, Gao J, Polycarbonate-based ultra-pH sensitive nanoparticles improve therapeutic window, Nature Communications 11(1) (2020) 5828. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [107].Su T, Zhang Y, Valerie K, Wang XY, Lin S, Zhu G, STING activation in cancer immunotherapy, Theranostics 9(25) (2019) 7759–7771. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [108].Barker JR, Koestler BJ, Carpenter VK, Burdette DL, Waters CM, Vance RE, Valdivia RH, STING-dependent recognition of cyclic di-AMP mediates type I interferon responses during Chlamydia trachomatis infection, mBio 4(3) (2013) e00018–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [109].Kranzusch Philip J., Wilson Stephen C., Lee Amy S.Y., Berger James M., Doudna Jennifer A., Vance Russell E., Ancient Origin of cGAS-STING Reveals Mechanism of Universal 2′,3′ cGAMP Signaling, Molecular Cell 59(6) (2015) 891–903. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [110].Li L, Yin Q, Kuss P, Maliga Z, Millán JL, Wu H, Mitchison TJ, Hydrolysis of 2′3′-cGAMP by ENPP1 and design of nonhydrolyzable analogs, Nature Chemical Biology 10(12) (2014) 1043–1048. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [111].Kato K, Nishimasu H, Oikawa D, Hirano S, Hirano H, Kasuya G, Ishitani R, Tokunaga F, Nureki O, Structural insights into cGAMP degradation by Ecto-nucleotide pyrophosphatase phosphodiesterase 1, Nature Communications 9(1) (2018) 4424. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [112].Ritchie C, Cordova AF, Hess GT, Bassik MC, Li L, SLC19A1 Is an Importer of the Immunotransmitter cGAMP, Molecular Cell 75(2) (2019) 372–381.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [113].Conlon J, Burdette DL, Sharma S, Bhat N, Thompson M, Jiang Z, Rathinam VAK, Monks B, Jin T, Xiao TS, Vogel SN, Vance RE, Fitzgerald KA, Mouse, but not Human STING, Binds and Signals in Response to the Vascular Disrupting Agent 5,6-Dimethylxanthenone-4-Acetic Acid, The Journal of Immunology 190(10) (2013) 5216–5225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [114].Cavlar T, Deimling T, Ablasser A, Hopfner KP, Hornung V, Species-specific detection of the antiviral small-molecule compound CMA by STING, Embo j 32(10) (2013) 1440–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [115].Shanahan CA, Gaffney BL, Jones RA, Strobel SA, Identification of c-di-GMP derivatives resistant to an EAL domain phosphodiesterase, Biochemistry 52(2) (2013) 365–77. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [116].Ager CR, Zhang H, Wei Z, Jones P, Curran MA, Di Francesco ME, Discovery of IACS-8803 and IACS-8779, potent agonists of stimulator of interferon genes (STING) with robust systemic antitumor efficacy, Bioorganic & Medicinal Chemistry Letters 29(20) (2019) 126640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [117].Lioux T, Mauny MA, Lamoureux A, Bascoul N, Hays M, Vernejoul F, Baudru AS, Boularan C, Lopes-Vicente J, Qushair G, Tiraby G, Design, Synthesis, and Biological Evaluation of Novel Cyclic Adenosine-Inosine Monophosphate (cAIMP) Analogs That Activate Stimulator of Interferon Genes (STING), J Med Chem 59(22) (2016) 10253–10267. [DOI] [PubMed] [Google Scholar]
- [118].Kim D-S, Endo A, Fang FG, Huang K-C, Bao X, Choi H.-w., Majumder U, Shen YY, Mathieu S, Zhu X, Sanders K, Noland T, Hao M-H, Chen Y, Wang JY, Yasui S, TenDyke K, Wu J, Ingersoll C, Loiacono KA, Hutz JE, Sarwar N, E7766, a Macrocycle-Bridged Stimulator of Interferon Genes (STING) Agonist with Potent Pan-Genotypic Activity, ChemMedChem n/a(n/a). [DOI] [PubMed] [Google Scholar]
- [119].Le Naour J, Zitvogel L, Galluzzi L, Vacchelli E, Kroemer G, Trial watch: STING agonists in cancer therapy, Oncoimmunology 9(1) (2020) 1777624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [120].Amouzegar A, Chelvanambi M, Filderman JN, Storkus WJ, Luke JJ, STING Agonists as Cancer Therapeutics, Cancers 13(11) (2021) 2695. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [121].Chin EN, Yu C, Vartabedian VF, Jia Y, Kumar M, Gamo AM, Vernier W, Ali SH, Kissai M, Lazar DC, Nguyen N, Pereira LE, Benish B, Woods AK, Joseph SB, Chu A, Johnson KA, Sander PN, Martínez-Peña F, Hampton EN, Young TS, Wolan DW, Chatterjee AK, Schultz PG, Petrassi HM, Teijaro JR, Lairson LL, Antitumor activity of a systemic STING-activating non-nucleotide cGAMP mimetic, Science 369(6506) (2020) 993–999. [DOI] [PubMed] [Google Scholar]
- [122].Pan B-S, Perera SA, Piesvaux JA, Presland JP, Schroeder GK, Cumming JN, Trotter BW, Altman MD, Buevich AV, Cash B, Cemerski S, Chang W, Chen Y, Dandliker PJ, Feng G, Haidle A, Henderson T, Jewell J, Kariv I, Knemeyer I, Kopinja J, Lacey BM, Laskey J, Lesburg CA, Liang R, Long BJ, Lu M, Ma Y, Minnihan EC, O’Donnell G, Otte R, Price L, Rakhilina L, Sauvagnat B, Sharma S, Tyagarajan S, Woo H, Wyss DF, Xu S, Bennett DJ, Addona GH, An orally available non-nucleotide STING agonist with antitumor activity, Science 369(6506) (2020) eaba6098. [DOI] [PubMed] [Google Scholar]
- [123].Zhang X, Shi H, Wu J, Zhang X, Sun L, Chen C, Chen ZJ, Cyclic GMP-AMP containing mixed phosphodiester linkages is an endogenous high-affinity ligand for STING, Mol Cell 51(2) (2013) 226–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [124].Ramanjulu JM, Pesiridis GS, Yang J, Concha N, Singhaus R, Zhang S-Y, Tran J-L, Moore P, Lehmann S, Eberl HC, Muelbaier M, Schneck JL, Clemens J, Adam M, Mehlmann J, Romano J, Morales A, Kang J, Leister L, Graybill TL, Charnley AK, Ye G, Nevins N, Behnia K, Wolf AI, Kasparcova V, Nurse K, Wang L, Puhl AC, Li Y, Klein M, Hopson CB, Guss J, Bantscheff M, Bergamini G, Reilly MA, Lian Y, Duffy KJ, Adams J, Foley KP, Gough PJ, Marquis RW, Smothers J, Hoos A, Bertin J, Design of amidobenzimidazole STING receptor agonists with systemic activity, Nature 564(7736) (2018) 439–443. [DOI] [PubMed] [Google Scholar]
- [125].A Developed STING Agonist Has Systemic Antitumor Activity, Cancer Discovery 9(1) (2019) 12–12. [Google Scholar]
- [126].Shae D, Becker KW, Christov P, Yun DS, Lytton-Jean AKR, Sevimli S, Ascano M, Kelley M, Johnson DB, Balko JM, Wilson JT, Endosomolytic polymersomes increase the activity of cyclic dinucleotide STING agonists to enhance cancer immunotherapy, Nature Nanotechnology 14(3) (2019) 269–278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [127].Shae D, Baljon JJ, Wehbe M, Christov PP, Becker KW, Kumar A, Suryadevara N, Carson CS, Palmer CR, Knight FC, Joyce S, Wilson JT, Co-delivery of Peptide Neoantigens and Stimulator of Interferon Genes Agonists Enhances Response to Cancer Vaccines, ACS Nano 14(8) (2020) 9904–9916. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [128].Wehbe M, Wang-Bishop L, Becker KW, Shae D, Baljon JJ, He X, Christov P, Boyd KL, Balko JM, Wilson JT, Nanoparticle delivery improves the pharmacokinetic properties of cyclic dinucleotide STING agonists to open a therapeutic window for intravenous administration, Journal of Controlled Release 330 (2021) 1118–1129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [129].Zhou Q, Wang Y, Li X, Lu N, Ge Z, Polymersome Nanoreactor-Mediated Combination Chemodynamic-Immunotherapy via ROS Production and Enhanced STING Activation, Advanced Therapeutics 4(11) (2021) 2100130. [Google Scholar]
- [130].Hanson MC, Crespo MP, Abraham W, Moynihan KD, Szeto GL, Chen SH, Melo MB, Mueller S, Irvine DJ, Nanoparticulate STING agonists are potent lymph node–targeted vaccine adjuvants, The Journal of Clinical Investigation 125(6) (2015) 2532–2546. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [131].Koshy ST, Cheung AS, Gu L, Graveline AR, Mooney DJ, Liposomal Delivery Enhances Immune Activation by STING Agonists for Cancer Immunotherapy, Advanced Biosystems 1(1–2) (2017) 1600013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [132].Liu Y, Crowe WN, Wang L, Lu Y, Petty WJ, Habib AA, Zhao D, An inhalable nanoparticulate STING agonist synergizes with radiotherapy to confer long-term control of lung metastases, Nature Communications 10(1) (2019) 5108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [133].Wang J, Li P, Yu Y, Fu Y, Jiang H, Lu M, Sun Z, Jiang S, Lu L, Wu MX, Pulmonary surfactant–biomimetic nanoparticles potentiate heterosubtypic influenza immunity, Science 367(6480) (2020) eaau0810. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [134].Sun X, Zhang Y, Li J, Park KS, Han K, Zhou X, Xu Y, Nam J, Xu J, Shi X, Wei L, Lei YL, Moon JJ, Amplifying STING activation by cyclic dinucleotide–manganese particles for local and systemic cancer metalloimmunotherapy, Nature Nanotechnology 16(11) (2021) 1260–1270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [135].Gentili M, Kowal J, Tkach M, Satoh T, Lahaye X, Conrad C, Boyron M, Lombard B, Durand S, Kroemer G, Loew D, Dalod M, Théry C, Manel N, Transmission of innate immune signaling by packaging of cGAMP in viral particles, Science 349(6253) (2015) 1232–1236. [DOI] [PubMed] [Google Scholar]
- [136].McAndrews KM, Che SPY, LeBleu VS, Kalluri R, Effective delivery of STING agonist using exosomes suppresses tumor growth and enhances antitumor immunity, Journal of Biological Chemistry 296 (2021) 100523. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [137].Chen Y-P, Xu L, Tang T-W, Chen C-H, Zheng Q-H, Liu T-P, Mou C-Y, Wu C-H, Wu S-H, STING Activator c-di-GMP-Loaded Mesoporous Silica Nanoparticles Enhance Immunotherapy Against Breast Cancer, ACS Applied Materials & Interfaces 12(51) (2020) 56741–56752. [DOI] [PubMed] [Google Scholar]
- [138].Park KS, Xu C, Sun X, Louttit C, Moon JJ, Improving STING Agonist Delivery for Cancer Immunotherapy Using Biodegradable Mesoporous Silica Nanoparticles, Advanced Therapeutics 3(10) (2020) 2000130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [139].Junkins RD, Gallovic MD, Johnson BM, Collier MA, Watkins-Schulz R, Cheng N, David CN, McGee CE, Sempowski GD, Shterev I, McKinnon K, Bachelder EM, Ainslie KM, Ting JPY, A robust microparticle platform for a STING-targeted adjuvant that enhances both humoral and cellular immunity during vaccination, Journal of Controlled Release 270 (2018) 1–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [140].Watkins-Schulz R, Tiet P, Gallovic MD, Junkins RD, Batty C, Bachelder EM, Ainslie KM, Ting JPY, A microparticle platform for STING-targeted immunotherapy enhances natural killer cell- and CD8+ T cell-mediated anti-tumor immunity, Biomaterials 205 (2019) 94–105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [141].Lu X, Miao L, Gao W, Chen Z, McHugh KJ, Sun Y, Tochka Z, Tomasic S, Sadtler K, Hyacinthe A, Huang Y, Graf T, Hu Q, Sarmadi M, Langer R, Anderson DG, Jaklenec A, Engineered PLGA microparticles for long-term, pulsatile release of STING agonist for cancer immunotherapy, Science Translational Medicine 12(556) (2020) eaaz6606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [142].Park CG, Hartl CA, Schmid D, Carmona EM, Kim H-J, Goldberg MS, Extended release of perioperative immunotherapy prevents tumor recurrence and eliminates metastases, Science Translational Medicine 10(433) (2018) eaar1916. [DOI] [PubMed] [Google Scholar]
- [143].Wang F, Su H, Xu D, Dai W, Zhang W, Wang Z, Anderson CF, Zheng M, Oh R, Wan F, Cui H, Tumour sensitization via the extended intratumoural release of a STING agonist and camptothecin from a self-assembled hydrogel, Nature Biomedical Engineering 4(11) (2020) 1090–1101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [144].He Y, Hong C, Yan EZ, Fletcher SJ, Zhu G, Yang M, Li Y, Sun X, Irvine DJ, Li J, Hammond PT, Self-assembled cGAMP-STINGΔTM signaling complex as a bioinspired platform for cGAMP delivery, Science Advances 6(24) (2020) eaba7589. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [145].Sun X, Ni Y, He Y, Yang M, Tani T, Kitajima S, Barbie DA, Li J, Engineering the Immune Adaptor Protein STING as a Functional Carrier, Advanced Therapeutics 4(8) (2021) 2100066. [Google Scholar]
- [146].Bukhalid RA, Duvall JR, Cetinbas NM, Catcott KC, Avocetien K, Bentley KW, Bradley S, Carter T, Chin C-N, Clardy S, Collins SD, Eitas T, Jones BD, Kelleher EW, Mosher R, Nazzaro M, Protopopova M, Shaw P, Slocum K, Ter-Ovanesyan E, Qin L, Thomas JD, Xu L, Yang L, Zurita J, Toader D, Damelin M, Lowinger TB, Abstract 6706: Systemic administration of STING agonist antibody-drug conjugates elicit potent anti-tumor immune responses with minimal induction of circulating cytokines, Cancer Research 80(16 Supplement) (2020) 6706–6706. [Google Scholar]
- [147].Medzhitov R, Janeway CA, Decoding the Patterns of Self and Nonself by the Innate Immune System, Science 296(5566) (2002) 298–300. [DOI] [PubMed] [Google Scholar]
- [148].Schlee M, Hartmann G, Discriminating self from non-self in nucleic acid sensing, Nature Reviews Immunology 16(9) (2016) 566–580. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [149].Pradeu T, Jaeger S, Vivier E, The speed of change: towards a discontinuity theory of immunity?, Nature Reviews Immunology 13(10) (2013) 764–769. [DOI] [PubMed] [Google Scholar]
- [150].Pradeu T, Vivier E, The discontinuity theory of immunity, Sci Immunol 1(1) (2016) AAG0479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [151].Wherry EJ, Kurachi M, Molecular and cellular insights into T cell exhaustion, Nature Reviews Immunology 15(8) (2015) 486–499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [152].Tripathy SK, Keyel PA, Yang L, Pingel JT, Cheng TP, Schneeberger A, Yokoyama WM, Continuous engagement of a self-specific activation receptor induces NK cell tolerance, Journal of Experimental Medicine 205(8) (2008) 1829–1841. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [153].Ercolini AM, Miller SD, The role of infections in autoimmune disease, Clinical & Experimental Immunology 155(1) (2009) 1–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [154].Wherry EJ, Ha S-J, Kaech SM, Haining WN, Sarkar S, Kalia V, Subramaniam S, Blattman JN, Barber DL, Ahmed R, Molecular Signature of CD8+ T Cell Exhaustion during Chronic Viral Infection, Immunity 27(4) (2007) 670–684. [DOI] [PubMed] [Google Scholar]
- [155].Netea MG, Domínguez-Andrés J, Barreiro LB, Chavakis T, Divangahi M, Fuchs E, Joosten LAB, van der Meer JWM, Mhlanga MM, Mulder WJM, Riksen NP, Schlitzer A, Schultze JL, Stabell Benn C, Sun JC, Xavier RJ, Latz E, Defining trained immunity and its role in health and disease, Nature Reviews Immunology 20(6) (2020) 375–388. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [156].Tassa C, Duffner JL, Lewis TA, Weissleder R, Schreiber SL, Koehler AN, Shaw SY, Binding Affinity and Kinetic Analysis of Targeted Small Molecule-Modified Nanoparticles, Bioconjugate Chemistry 21(1) (2010) 14–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [157].Velazquez-Campoy A, Freire E, Isothermal titration calorimetry to determine association constants for high-affinity ligands, Nature Protocols 1(1) (2006) 186–191. [DOI] [PubMed] [Google Scholar]
- [158].Heyduk T, Ma Y, Tang H, Ebright RH, Fluorescence anisotropy: Rapid, quantitative assay for protein-DNA and protein-protein interaction, Methods in Enzymology, Academic; Press1996, pp. 492–503. [DOI] [PubMed] [Google Scholar]
- [159].Homola J, Piliarik M, Surface Plasmon Resonance (SPR) Sensors, in: Homola J (Ed.), Surface Plasmon Resonance Based Sensors, Springer Berlin Heidelberg, Berlin, Heidelberg, 2006, pp. 45–67. [Google Scholar]
- [160].Boyer C, Bulmus V, Davis TP, Ladmiral V, Liu J, Perrier S, Bioapplications of RAFT Polymerization, Chemical Reviews 109(11) (2009) 5402–5436. [DOI] [PubMed] [Google Scholar]
- [161].Matyjaszewski K, Atom Transfer Radical Polymerization (ATRP): Current Status and Future Perspectives, Macromolecules 45(10) (2012) 4015–4039. [Google Scholar]
- [162].Feng J, Zhuo R-X, Zhang X-Z, Construction of functional aliphatic polycarbonates for biomedical applications, Progress in Polymer Science 37(2) (2012) 211–236. [Google Scholar]
- [163].Samaddar P, Deep A, Kim K-H, An engineering insight into block copolymer self-assembly: Contemporary application from biomedical research to nanotechnology, Chemical Engineering Journal 342 (2018) 71–89. [Google Scholar]
- [164].Zhou K, Wang Y, Huang X, Luby-Phelps K, Sumer BD, Gao J, Tunable, Ultrasensitive pH-Responsive Nanoparticles Targeting Specific Endocytic Organelles in Living Cells, Angewandte Chemie International Edition 50(27) (2011) 6109–6114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [165].Liu S, Cheng Q, Wei T, Yu X, Johnson LT, Farbiak L, Siegwart DJ, Membrane-destabilizing ionizable phospholipids for organ-selective mRNA delivery and CRISPR–Cas gene editing, Nature Materials 20(5) (2021) 701–710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [166].Zhou S, Deng X, Yang H, Biodegradable poly(ε-caprolactone)-poly(ethylene glycol) block copolymers: characterization and their use as drug carriers for a controlled delivery system, Biomaterials 24(20) (2003) 3563–3570. [DOI] [PubMed] [Google Scholar]
- [167].Huang S-W, Wang J, Zhang P-C, Mao H-Q, Zhuo R-X, Leong KW, Water-Soluble and Nonionic Polyphosphoester: Synthesis, Degradation, Biocompatibility and Enhancement of Gene Expression in Mouse Muscle, Biomacromolecules 5(2) (2004) 306–311. [DOI] [PubMed] [Google Scholar]
- [168].Ulery BD, Nair LS, Laurencin CT, Biomedical Applications of Biodegradable Polymers, J Polym Sci B Polym Phys 49(12) (2011) 832–864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [169].Kamaly N, Yameen B, Wu J, Farokhzad OC, Degradable Controlled-Release Polymers and Polymeric Nanoparticles: Mechanisms of Controlling Drug Release, Chemical Reviews 116(4) (2016) 2602–2663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [170].Miller J, Luo M, Wang H, Wang Z, Ding X, Campbell A, Almazan J, Gutowski S, Chen Z, Gao J, Zhao T, Abstract 4577: ONM-500: A STING-activating therapeutic nanovaccine platform for cancer immunotherapy, Cancer Research 80(16 Supplement) (2020) 4577–4577. [Google Scholar]