Polymers for siRNA Delivery: A Critical Assessment of Current

Aug 26, 2016 - Polymers for siRNA Delivery: A Critical Assessment of Current Technology Prospects for Clinical Application. Bailey M. Cooper† and Da...
0 downloads 0 Views 3MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Review

Polymers for siRNA delivery: A Critical Assessment of Current Technology Prospects for Clinical Application Bailey Mae Cooper, and David Putnam ACS Biomater. Sci. Eng., Just Accepted Manuscript • DOI: 10.1021/acsbiomaterials.6b00363 • Publication Date (Web): 26 Aug 2016 Downloaded from http://pubs.acs.org on August 29, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Biomaterials Science & Engineering is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

Polymers for siRNA delivery: A Critical Assessment of Current Technology Prospects for Clinical Application

Bailey M. Cooper1 and David Putnam1,2*

1

Meinig School of Biomedical Engineering, 2Smith School of Chemical and Biomolecular Engineering,

Cornell University, Ithaca, NY 14853

*

To whom correspondence should be addressed: 147 Weill Hall, Cornell University, Ithaca, NY 14853.

Email: [email protected]

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract The number of polymer-based vectors for siRNA delivery in clinical trials lags behind other delivery strategies, however, the molecular architectures and chemical compositions available to polymers make them attractive candidates for further exploration. Polymer vectors are extensively investigated in academic labs worldwide with fundamental progress having recently been made in the areas of high-throughput screening, synthetic methods, cellular internalization, endosomal escape and computational prediction and analysis. This review assesses recent advances within the field and highlights relevant developments from within the complementary fields of nanotechnology and protein chemistry with the intent to propose future work that addresses key gaps within the current body of knowledge – potentially advancing the development of the next generation of polymeric vectors.

Keywords: siRNA, DNA, polymer chemistry, structure-activity relationships

ACS Paragon Plus Environment

Page 2 of 50

Page 3 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

Introduction Eighteen years following the first thorough explanation of the RNAi mechanism1 and ten years following the first presentation of clinical data for an siRNA-based therapeutic,2 only seven of the more than sixty national clinical trials (NCTs) studying siRNA delivery systems employed synthetic polymeric vectors - either in the form of surgically-implanted local drug eluters (LODERs) or systemicallyadministered cyclodextrin-based systems and dynamic poly-conjugates (which employ multi-functional, environmentally responsive polymers designed to facilitate separate aspects of delivery). Other in vivo strategies under consideration include self-delivering siRNAs, in which chemical modifications are used to incorporate drug-like properties into the nucleic acid compound, and GalNAc conjugates, which employ the small molecule triantennary N-acetylgalactosamine as a targeting ligand. In a field dominated by liposomal formulations and chemically modified-naked siRNA, the prospects of nonlipidoid synthetic polymers remain unclear. Customizability, stability, low cost and ease of manufacture of polymeric vectors make them attractive candidates for research; however, they have suffered from low efficiency, cytotoxicity, poorly controlled synthesis, and a general lack of mechanistic understanding of their cellular uptake, intracellular trafficking and payload release. An assessment of the potential for polymer vector viability in the clinic is needed. The use of high throughput library-based approaches aided by predictive computational models, innovative polymer structures, new chemistries and a growing understanding of key cellular mechanisms, have propelled the field forward substantially in recent years. But even as the field rapidly matures in response to cutting-edge experimental techniques and improved mechanistic insights into siRNA delivery, there remain non-trivial obstacles to greater adoption of polymer vectors in clinical trials. This review critically explores and summarizes the integration of the new methods, materials, chemistries and computational methods brought to bear in an effort to better understand and navigate the intracellular delivery of siRNA sequences.

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Polymeric libraries - rapid screening in vitro plus considerations for in vivo studies

Figure 1. High Throughput (HT) methods of analysis allow for the rapid assessment of multiple parameters, thereby enabling researchers to identify structure-activity relationships within large libraries of polymers. Reprinted with permission from ref 4. Copyright 2013 American Chemical Society.

The broad molecular parameter space available to polymeric vectors enables investigators to explore widely diverse molecular architectures and compositions in the search of siRNA delivery systems with enhanced capabilities. Additionally, the influence of polymer structure and composition on siRNA delivery can help to enhance or refine siRNA delivery. One approach to this end is through the parallel screening of polymer libraries to facilitate the rapid evaluation of structure-activity relationships (Figure 1).3 One challenge to the high-throughput screening approach is the reliable correlation between in vitro and in vivo efficacy, as well as reliable preclinical animal models to accurately predict clinical outcomes. The development of in vitro – in vivo correlations (IVIVCs) is particularly difficult for systemically administered materials where multiple cell types and boundaries are expected to contact the material. Two-dimensional cell culture alone does not have the capability to predict events such as the chelation of ions that can impact homeostasis, activation of intravascular coagulation cascades or agglomeration of particles following intravenous administration, etc.4 Furthermore, the most appropriate cell line(s) for

ACS Paragon Plus Environment

Page 4 of 50

Page 5 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

in vitro studies might not be obvious. For example, one study found that the HeLa cell line (human cervical carcinoma) provided data that was more predictive of hepatocellular delivery to healthy mice than the Hepa1-6 cell line (mouse hepatoma). While freshly isolated mouse hepatocytes “retained native gene function for 1-2 days post-isolation” and provided the best correlation (R2=0.99), the investigators found that HeLa cells provided adequate selection pressure for in vitro screening.5 Threedimensional cultures of human cells and “lab-on-a-chip” microfluidic devices have not been widely utilized for developing IVIVCs in this field, but hold the potential to serve as more predictive screening systems.6 Whitehead et al. reviewed the correlation that exists between transfection and in vivo efficacy for a number of lipidoid vectors targeting hepatocellular siRNA delivery and established that variations in formulation technique significantly affect the IVIVC.5 A comparable study for synthetic polymers would be a useful and critical step to help facilitate the accurate identification of candidate polymers for in vivo efficacy while potentially providing insight that may lead to the development of more relevant in vitro models. Furthermore, standardization of screening practices across the field (e.g. a base set of cell lines for in vitro studies, uniform transfection SOPs (standard operating procedures), a uniform set of commercially available positive controls , etc.) would provide a foundation for the quantitative comparison of results from different studies. The development of in silico approaches, leading to predictive models,7 and the broader adoption of quantitative studies evaluating structure-activity relationships6 would both aid in the search for more accurate trends with increasing clinical significance. Zuckerman et al. recently reported on the correlation between preclinical, multispecies animal studies and the initial Phase I clinical trials of the first systemically administered cationic polymer-siRNA delivery system - CALAA-01 – where “allometric scaling across species best [correlated] with body weight rather than with body surface area”. The area under the curve (AUC) and maximum plasma concentration (Cmax) appeared linearly related in all species, and the animal studies were generally

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

found to predict clinical outcomes, but the authors noted that the pharmacokinetics might deviate for systems not dependent upon electrostatic interactions since the primary clearance mechanism is renal clearance for polyplexes in contrast to the monocyte phagocytic clearance expected for other nanoparticles.8 Therefore, a comparable study for non-ionically associated polymer-siRNA complexes is needed within the field.

Advances in chemistry To explore the rich parameter space available to polymer-based vectors though the synthesis of libraries, a number of broadly applicable synthetic strategies have been reported. The early work of Lynn, Anderson and Langer established the utility of Michael-type additions to form libraries of poly(beta-amino esters) and introduced the concept of combinatorial libraries for nucleic acid delivery. As more labs begin to span the discovery-to-development timeline, where both the fundamental understanding of transfection and the practicality of product viability are equally important, the chemistries used to investigate new polymer structures and architectures have evolved to accommodate both facile synthesis as well as feasible scale-up. Perhaps most important to the synthesis of polymer libraries that can lead to an understanding of structure-activity relationships, as well as scalability for translation into the clinic, is the fidelity and reproducibility of the reactions. Small changes to polymer structure can have significant influence on transfection, and the presence of uncharacterized side-reaction products on the polymer can lead to incorrect conclusions with respect to how polymer structure correlates to siRNA delivery. One example of how a seemingly simple polymer-analogous conjugation can have unwittingly side reactions is the use of polymers containing side chains terminated with the N-hydroxysuccinimide (NHS) leaving group (Figure 2A). The activated ester of the NHS group allows amidation when incubated with amino groups.9–11 However, two unexpected side reactions can occur. One side reaction is the formation of a

ACS Paragon Plus Environment

Page 6 of 50

Page 7 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

linear adduct generated in sterically constrained environments from attack of the amino group on one of the ring carbonyl NHS groups (Figure 2B). The other side reaction is the formation of a cyclic glutarimide through amide attack on a proximal NHS group (Figure 2C). There are ways to circumvent these specific side reactions, as well as others formed during library synthesis.12 But without first fully characterizing the conditions under which they are formed, failure to recognize their existence can lead to mischaracterization of structure-activity relationships, which can complicate robust scale-up and reproducibility for commercial or clinical applications.

A

B

C

Figure 2. (A) Polymer with NHS side chains used to make polymer libraries (B) Structure of ring addition side reaction product when amine attacks carbonyl on the NHS ring (C) Structure of glutarimide side reaction product when amine attacks NHS in close proximity

Click chemistry methods are reported for the synthesis of polymer libraries for the delivery of siRNA sequences. The Reineke group has investigated carbohydrate-based polymers and reported how polymer length and type of carbohydrate influence delivery potency using copper-catalyzed azide/alkyne cycloaddition. Trehalose and β-cyclodextrin were made part of the backbone though copolymerization of oligoethylenamine.13 The influence of carbohydrate type was nucleic aciddependent. Trehalose-containing polymers were superior for plasmid DNA whereas β-cyclodextrin was

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

superior for siRNA. UV-initiated thiol-ene click chemistry is another nice approach to the synthesis of polymer libraries, although its use has yet to be reported for siRNA delivery (Figure 3). One example of its utility was reported by the group of Nystrom wherein histamine-functionalized polymers were investigated for their ability to deliver doxorubicin to in vitro 3D breast cancer architectures.14

Figure 3. UV-activated thiol-ene click chemistry used to functionalize polymer side chains

The addition of amine to epoxide groups is clean and robust. Polymers containing side chains terminated with epoxides were used as a crosslinking strategy to make nanoparticles containing amine crosslinking sequences (Figure 4).15 Each material made in this very large library, which contained over 1500 unique compositions, was thoroughly characterized for molecular weight, hydrodynamic diameter in solution, the ability to complex with nucleic acids, cellular internalization and the ability to deliver both plasmid DNA and siRNA. The result was a comprehensive understanding of how the material composition led to delivery, both in cell culture and in animal models. In particular, polymers crosslinked with piperazine and dimethylamine worked optimally, particularly when cholesterol was included as a way to facilitate transport to hepatocytes.

ACS Paragon Plus Environment

Page 8 of 50

Page 9 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

Figure 4. Epoxide amine addition reaction (top) applied to the crosslinking of epoxideterminated polymers (bottom) toward the synthesis of polymer libraries.

Synthesis of more targeted and complex combinatorial libraries often rely on more traditional condensation reagents, like EDAC and maleimide, in order to give more flexibility and control over the final stoichiometry of components. A recent example is the work of Amiji wherein a combinatorial library, based on chitosan as a backbone, was generated varying the content of polyethylene glycol, a lipid chain, and the targeting agent (epithelial growth factor receptor binding peptide). The approach was used to investigate the delivery of both siRNA as well as cisplatin.16 A newer condensation reagent, DMTMM, has been used to create poly(acrylic acid)-based polymer libraries for siRNA delivery containing variations in side chain carbohydrate (galactose) and amine (agmatine) composition.17 Conjugation strategies like these are widely used and are simple and fairly robust for laboratory scale work to define structure-activity relationships, but are unlikely to be viable strategies for large-scale manufacturing owing to expense and purification challenges.

Complexation Strategies to package and deliver siRNA using synthetic polymers include physical entrapment within a nano- or microparticle matrix, covalent bonding, or electrostatic self-assembly. The latter is by

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

far the most common but can result in thermodynamically pseudo-stable structures18 due to the short length (~6nm) and rigid nature of siRNA.19 Stability has been improved through the use of “gene-like” RNAs (formed when sticky ends enable self-assembly into longer structures),20 the addition of crosslinking elements, and by reducing the rigidity of the polymer.21–23 Although robust binding between polymers and nucleic acids (NA) favors structural stability and protection, these benefits must be balanced against the need for timely intracellular dissociation, and recently, both cellular internalization and subcellular distribution were shown to be influenced by binding affinity in a cell line-dependent manner.21 Current research supports an optimal balance between the size and rigidity of the polymer and that of the NA being condensed, with short siRNA molecules demonstrating favorable complexation with flexible 25kDa PEI and generation 7 TEA-core PAMAM dendrimers (attributed to the relative inability of complexes formed by rigid NAs to fold, physically rearrange or result in hydrophobic aggregation of complexes in solution), while less-flexible generation 5 dendrimers required 25 to 27 mer sticky siRNAS in order to achieve sufficiently strong binding.20,22 Naked, un-modified, siRNA has a circulation half-life on the order of minutes. When using a polymer carrier, the packaging must be sufficient to protect the siRNA from enzymatic degradation prior to cytosolic delivery, yet still able to fully un-package or dissociate from the carrier before being functionally incorporated into the RISC complex. As a result, the optimum complexation strength is material-dependent, but is seldom at the maximum.24 This balance is further complicated following administration into the circulation by physiological salt conditions and competing anionic biomacromolecules. Another interesting challenge is the kidney’s glomerular basal membrane (GBM); in circulation, particles formed primarily by electrostatic interactions are prone to transient decomplexation at the GBM due to interactions with negatively charged proteoglycans – leading to rapid clearance.25 Cyclodextrin-based nanoparticles (NPs) are one example of a material prone to this phenomenon. When formulated with a hydrodynamic radius less than 100nm with a positive zeta

ACS Paragon Plus Environment

Page 10 of 50

Page 11 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

potential, Zuckerman et al. showed that the components of a cyclodextrin-based delivery system could self-assemble in circulation and potentially in urine (post excretion), but would “transiently accumulate in and be disassembled by the GBM”. This phenomenon was not observed in similarly-sized particles with anionic zeta potentials, which the authors attribute to charge repulsion at the GBM. Furthermore, they recommend heparin sulfate stability be considered when designing and validating future systems utilizing ionic complexation.26

Systemic Delivery As the field of polymeric siRNA delivery matures, an understanding of polymer-blood interactions is essential to elucidate the key structure-activity relationships needed to rationally design the next generation of vectors, as well as to identify ways to better correlate the outcome of in vitro studies to pre-clinical in vivo models. Toward this aim, strides have been made recently within the fields of drug delivery, nanotechnology and proteomics, but further work is needed to establish a meaningful understanding of the ways in which polymer architectures and polyplex parameters affect a wide range of interactions within the blood stream. Upon intravenous injection, polymeric vectors come into contact with a vast array of cells and proteins in the blood stream, which can cause the rapid formation of a complex protein corona.27 Until recently, little was known regarding the pharmacokinetics and pharmacodynamics involved with these interactions. In 2012 a systematic investigation of blood interactions at a molecular and cellular level revealed 41 plasma proteins that interact with PEI. By varying the molecular weight (from 0.6 - 25kDa) and structure (linear vs branched) of PEI, the investigators discovered this common delivery vector had a profound impact on the membrane structure of red blood cells, the conformation of albumin and blood coagulation components, and systemic oxygen delivery efficacy at higher doses (see Figure 5).28,29 A year later, proteomic analysis revealed nearly 300 different proteins that could rapidly adsorb onto the

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

surface of silica and polystyrene nanoparticles, affecting not only cellular uptake but also cytotoxicity, hemolysis and thrombocyte activation of the vectors. Tenzer et al showed that these coronas vary in quantity but not composition, and recommended further quantitative proteomic analysis to predictively model serum protein binding kinetics (see Table 1).27 Since surface proteins may effect many aspects of polymeric delivery systems (i.e. immune response, transport, coagulation, etc.), similar quantitative and qualitative proteomic studies across a range of polymeric siRNA delivery vectors could provide valuable insights into materials selection parameters for future delivery systems.

Figure 5. SEM images showing the effects of PEI on both the morphology and aggregation of red blood cells. At 0.6kDa and 1.8kDa, branched PEI was observed to induce morphological changes in a concentration-dependent manner, with RBCs progressing from biconcave to crenated to spherical with spicules along the surface. At higher molecular weights (i.e. 25kDa), both linear and branched PEI induced aggregation as well as morphological changes when administered at 0.1-1mg/mL. Reprinted with permission from ref 28. Copyright 2013 Elsevier.

ACS Paragon Plus Environment

Page 12 of 50

Page 13 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

Table 1. The 20 most prevalent proteins (measured in ppm) comprising coronas after nanoparticles had been exposed to plasma for 30 seconds. Reprinted by permission from Macmillan Publishers Ltd: Nature Nanotechnology, ref 27. Copyright 2013 No. Amorphous Silica Nanoparticles (commercial) 1 Serum albumin* 2 Apolipoprotein A-I* 3 Complement C3* 4 Ig γ-1 chain C region*

Negatively Charged Polystyrene Nanoparticles Serum albumin* Complement C3* Complement factor H β-2-glycoprotein 1

5 6

Complement factor H Kininogen-1*

7 8 9

Complement C4-A Ig κ chain C region* Serotransferrin

10

Gelsolin

Kininogen-1* Inter-α-trypsin inhibitor heavy chain H4* Ig γ-1 chain C region* Vitronectin Complement C1r subcomponent* Ig γ-3 chain C region*

11

Ceruloplasmin

12

Ig γ-3 chain C region*

Lipopolysaccharide-binding protein Gelsolin

13 14 15 16 17

Positively Charged Polystyrene Nanoparticles Serum albumin* Apolipoprotein A-I* Ig γ-1 chain C region* Inter-α-trypsin inhibitor heavy chain H4* Ig μ chain C region Ig γ-3 chain C region* Ig κ chain C region* Vitronectin Complement C3* Complement C1r subcomponent* Complement C4-A Complement C1s subcomponent α-2-macroglobulin Serotransferrin Apolipoprotein A-IV α-1-antitrypsin Keratin, type II cytoskeletal 1

α-2-macroglobulin Complement C5 Haemopexin Complement C1s subcomponent Ig γ-4 chain C region Apolipoprotein A-I* β-2-glycoprotein 1 Ig μ chain C region Complement C1r Complement C4-B subcomponent* 18 Ig α-1 chain C region Ig κ chain C region* Kininogen-1* 19 Inter-α-trypsin inhibitor heavy C4b-binding protein α chain Clusterin chain H4* 20 Haptoglobin Histidine-rich glycoprotein Ig α-1 chain C region * Denotes proteins that are among the top 20 for all indicated nanoparticles.

Polymers can increase the circulation half-life of siRNA by providing protection against endogenous ribonucleases, hindering opsonization and contributing to a particle size above the glomerular filtration limit, but a strong net cationic charge on the carrier can lead to erythrocyte aggregation and enhanced uptake by the reticuloendothelial system.28 As a result, amphiphilic polymer side chains have long been favored for their ability to increase circulation time while also improving

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

cellular uptake. Perhaps the most widely used surface-modifying polymer for this application has been poly(ethylene glycol) (PEG), yet further quantitative studies are needed pertaining to siRNA delivery. In recent years, three studies have assessed the correlation between PEG molecular weight, surface density and polymer chain architecture without using siRNA. In one study, a rapid screening technique was developed that employed highly uniform hydrogel nanoparticles with varying PEGylation densities to study protein adsorption, macrophage uptake and circulation time in the absence of particle loading. Although all particles, regardless of size, shape or composition were ultimately filtered by the organs of the mononuclear phagocyte system, circulation time increased due to delayed phagocytosis, even when the distance between PEG grafts exceeded the Flory radius (RF), resulting in a mushroom conformation (see Figure 6).30 These findings are in contrast with earlier intravenous administrations in which dense PEG brushes were required for adequate shielding and many short PEG chains provided better protection than fewer, longer PEGs.31

Figure 6. Representation of functionalized PEG on the surface of a PRINT hydrogel nanoparticle showing two possible conformations - brush (A) and mushroom (B). Reprinted with permission from ref 30. Copyright 2012 American Chemical Society.

Zheng et al. have assessed the effect of graft density of biodegradable polycaprolactone-blockpoly(ethylene glycol) (PCL-PEG) on a 25kDa hyperbranched PEI (hy-PEI) backbone and found that higher graft densities of 3 and 5 enhanced circulation time and siRNA transfection efficiency. Although

ACS Paragon Plus Environment

Page 14 of 50

Page 15 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

molecular weight was increased, micelle hydrodynamic diameter, polydispersity, and zeta potential decreased, which was attributed to the increasingly amphiphilic nature of the copolymer. At these two ratios, 95% of siRNA was completely condensed, greater stability against competing polyanions was observed, transfection rates improved (showing 72% and 84% knockdown of GAPDH, respectively), and pharmacokinetics improved (with 3-4 fold greater AUC compared to 25kDa hy-PEI complexes with free siRNA). The increased circulation time was attributed to greater polyplex stability and a decreased zeta potential theoretically resulting in a decreased rate of interaction with macrophages.32 Similarly, Zhao et al. showed that spermine densely grafted to a polyglutamate-derivative backbone yielded serumresistant complexes with weaker binding to both serum proteins and nucleotides, compared to PEI 25kDa, resulting in highly stable polyplexes capable of rapid cytosolic unpacking and transfection rates comparable to Lipofectamine 2000.33 The adsorption of serum proteins onto the surface of a delivery vehicle can be utilized as a “natural functionalization” with beneficial therapeutic effect27 or, alternatively, can mask or alter desired functionality.34 This complex phenomenon can influence both extracellular and intracellular interactions35 - further demonstrating the need for an enhanced understanding of the relationship between physiochemical properties and function in order to enable the optimization of polymer size, nanoparticle packing density, and composition leading to the rational design of clinically relevant delivery systems.

Immunological Considerations Naked siRNA can activate the innate immune system via toll-like receptors (TLRs) – especially TLRs 3 (dsRNA), 7 and 8 (ssRNA) – and cytoplasmic receptors.36 Delivery vectors are intended to shield the RNA from recognition by the immune system, and the relatively inert nature of polymers makes them attractive options. However, as previously discussed, the properties of the vector material (e.g.

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

charge, hydrophobicity, rigidity and physical size of resulting complexes) will effect both the manner by which compounds enter the cell and into which intracellular compartments they are sorted, therefore dictating which TLRs will be encountered and potentially stimulated.36 Opsonin proteins adhered to the surface of a foreign object mark the object for phagocytosis. While the immunostimulatory effects of dsRNA are understood and can be mitigated or eliminated through 2’-O-methyl or 2’ fluoro ribonucleotide modifications, polymeric vectors can still have profound immunostimulatory effects both extra- and intracellularly. These effects are important for polymeric vectors that are often cationic and enter the cell via endocytosis where they are exposed to additional pattern recognition receptors (PRRs), thereby increasing the opportunity for innate immune recognition.36–38 To ascertain a structure-activity relationship between hydrophobicity and immune response, freshly-harvested murine splenocytes (comprised of B-lymphocytes, T cells and monocytes) were exposed to gold nanoparticles coated by PEG capped with head groups of varying degrees of hydrophobicity. A nearly linear relationship between the hydrophobicity of polymer head groups and the associated immune response of splenocytes in vitro was observed (see Figures 7 and 8). While increased hydrophobicity also correlated to increased immune response for polymers with LogP values less than 2.0 in vivo, the trend could not be observed in more hydrophobic polymers – possibly due to the poor in vivo biodistribution of such hydrophobic constructs and/or a maximum immune response being reached (see Figure 8B). Quantitative studies, such as this one, exploring the relationship between key material properties and biological function are essential for improving the efficacy of polymeric vectors for siRNA delivery.39 While much work has been done in this field in recent years,37,38 and databases now facilitate the exploration of various factors effecting innate immune response,40 the field will continue to grow as new, and increasingly intricate, polymers are developed. The advantageous or deleterious effects of

ACS Paragon Plus Environment

Page 16 of 50

Page 17 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

immunostimulation by synthetic vectors must be considered when designing experiments and interpreting results.

Figure 7. Nine chemical structures with varying degrees of hydrophobicity (quantified by an estimate of Log P values via MacroModel – Maestro 8.0) were each conjugated to the ligand termini of tetra(ethylene glycol) spacers attached to gold nanoparticles via hydrocarbon linkers. The resulting library of particles was used to explore the structure-activity relationship between hydrophobicity and immune response. Reprinted with permission from ref 39. Copyright 2012 American Chemical Society.

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 8. Normalized cytokine gene expression as a function of the calculated hydrophobicity (Log P) of headgroups conjugated to gold nanoparticles. Researchers reported on a representative proinflammatory cytokine, TNFα, measured in vitro (A) as well as a representative anti-inflammatory cytokine, IL-10, measured in vivo (B). Gene expression values are reported as a proportion of the expression measured after exposure to a positive control (LPS) under the same experimental set. No correlation was observed in vivo after 6 hours. Reprinted with permission from ref 39. Copyright 2012 American Chemical Society.

Targeting Targeting siRNA to specific organ and cell populations is preferred for systemic delivery as it minimizes the required dose of both polymer and nucleic acid, and reduces potential adverse off-target effects (e.g. cytotoxicity, off-target silencing, overwhelming the RNAi machinery). In the past, polymeric carriers have been designed to actively or passively target specific organs, tissues, disease states and cell types by exploiting unique environmental parameters (e.g. pH,41–44 temperature, redox potential,45–49 leaky vasculature,44,50 over expressed surface receptors,44,51–63 enzymes,64 etc.).65,66 Solid phase

ACS Paragon Plus Environment

Page 18 of 50

Page 19 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

synthesis, click chemistry, and novel phosphoramidation reactions have enabled the synthesis of many new siRNA-peptide/carbohydrate/small molecule conjugations,67 yielding mono-disperse, targeted polymeric vectors.54 For example, the attachment of siRNA to an anti-Her2 single-chain fragmented antibody (ScFv) with C-terminal protamine peptide and His6 tag was recently heralded as a “bullseye for breast cancer” due to its remarkable targeting ability,68 but it is important to note that the conjugation site of antibodies is crucial to avoid steric interference of antigen/ligand binding and subsequent internalization by the target cell. Chemically defined, site-specific antibody-polymer conjugates (APCs) for siRNA delivery have recently been synthesized by genetically incorporating an unnatural amino acid with biorthogonal reactivity (i.e. p-acetyl L-phenylalanine (pAcF)) into an antibody then selectively coupling the ketone group of pAcF to the terminal aminooxy group of a cationic polymer via a stable oxime bond. The resulting APCs exhibited binding affinities comparable to that of their parent antibodies, selective delivery, and significant silencing at both the mRNA and protein levels. High yields, ease of formulation and mild reaction conditions make this approach an attractive prospect for pharmaceutical scale-up and clinical application.69 Transferrin (Tf) receptors are overexpressed on many types of cancer cells, and Tf has been used as a targeting ligand in several systems currently undergoing clinical trials. Excitingly, investigators recently discovered a way to improve this common targeting modality. Mathematical modeling of cellular trafficking was used to identify ligand-metal interactions as a novel design criterion for improving the cellular association of human transferrin protein (Tf). When nanoparticles were conjugated to oxalate Tf, with its slower iron release rate, in vitro and in vivo studies validated the model.70 Using a similar approach, a library of tandem peptides for dual tumor targeting and cellpenetration were designed then screened for their ability to complex and deliver siRNA in a receptor-

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

specific manner. Computational regression analysis was used to identify the structural parameters key to cell type specificity (i.e. overall charge of the peptide and valence of the targeting ligand). Through this systemic approach, myristoylation was shown to not only strengthen hydrophobic interactions and peptide attraction to the cell membrane, but to also condense siRNA into multivalent nanocomplexes and facilitate targeted delivery. This combined computational/experimental approach could help to optimize peptides for cell-type specific delivery.71 Although this study did not employ synthetic polymers, the peptides and structure-activity relationship which were identified can help to inform the delivery of future polymeric delivery systems.

Cellular Uptake Crossing the cell membrane and entering the cytosol remains a critical challenge for polymerbased delivery systems. While all systems appear to undergo some form of endocytosis, some Crossing the cell membrane and entering the cytosol remains a critical challenge for polymer-based delivery systems. While all systems appear to undergo some form of endocytosis, some paths (e.g. caveolaemediated endocytosis and clathrin-independent endocytosis) hold the potential to expedite uptake and bypass the endosome (along with the associated TLRs) and eventual lysosomal degradation.37,72,73 The uptake pathway is polymer-dependent, and earlier reviews have discussed the relationships between the vector composition, internalization and intracellular trafficking for both polymer- and lipidbased strategies.74,75 Caveolar uptake has been shown to be the dominant pathway for a library of micelleplexes76 as well as a variety of polyplexes formed from cationic polymers,77–79 while clathrinmediated endocytosis was the primary pathway for G5 PAMAM dendrimer-based nanoparticles80 and PLGA-grafted, diamine-modified poly(vinyl alcohol) nanoparticles.81 Furthermore, the preferred pathway appears to be system-dependent with some studies showing improved outcomes after clathrinmediated endocytosis77 and others after caveolae-mediated delivery.74,76 Although earlier studies have

ACS Paragon Plus Environment

Page 20 of 50

Page 21 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

suggested a relationship between most uptake mechanisms and either the size or charge of internalized particles,77,82,83 a new mechanism of macropinocytosis was recently proposed based on experiments showing that cationic polystyrene nanoparticles were endocytosed at a rate dependent upon total polymer mass rather than surface area or number of particles (Figure 9).84

Figure 9. Proposed internalization routes for cationic polystyrene particles of various sizes entering HeLa cells. An inverse correlation between particle size and number of particles per vesicle suggests an “excavator shovel like mechanism” in which the maximum volume of each vesicle is finite, and vesicles form along the cell membrane at a limited rate. Reprinted from European Journal of Pharmaceutics and Biopharmaceutics, 84 / 2, Lerch, et al. “Polymeric nanoparticles of different sizes overcome the cell membrane barrier”, (2013), with permission from Elsevier.84

Functionalization that alters the physiochemical properties of a polymer can also significantly impact the rate and mechanism of cellular uptake. For example, when facial amphipathic deoxycholic acid was conjugated to the terminal amine groups of 1.8kDa PEI, cellular uptake surpassed that of 25kDa PEI without a significant increase in toxicity, due to an apparently energy-independent, non-endocytotic pathway related to enhanced membrane permeability, which mimics the cell penetrating peptide (CPP)mediated transport of macromolecules across the cell membrane. Furthermore, the authors recommend this synthetic conjugation of molecular amphiphiles for use in future high throughput studies .85,86 Since the cell internalization pathway of a polymeric vector is dependent upon both the

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

delivery system and cell type,87 there are clear advantages to testing new vectors across a variety of cell types to better understand the potential utility of each vector. Although favorable for masking polymers and stabilizing polyplexes in circulation, PEG can hinder internalization of cationic polymer systems. Interestingly, optimized PEGylation was recently shown to increase receptor-mediated transfection efficiency of anionic nanocomplexes.88 When attached by disulfide bonds or benzoic imine linkers, PEG can shield the system in circulation then be shed in reductive or acid microenvironments to facilitate uptake at the target site.89–92 Gelatin coatings,93,94 polymerized trehalose coatings,95 crosslinking of surface amines, hydrophobic modifications, disulfide bridges, pi-pi stacking, and tyrosine modifications have been used as alternative stabilization strategies for polyplexes. 96–100 Tyrosine modifications are particularly versatile as they have been used as a hydrophobic modification (e.g. partially modifying amines in PEI), as a functionalization for pi-pi stacking (as an aromatic amino acid), and as promotor of endosomal buffering (again due to the aromatic structural component). 96,99,101,102 Hydrophobic acrylates have been used to stabilize the core of polyamine-based structures, and the cross-linking of surface amines has improved the transfection ability of polyplexes utilizing a variety of amine-based polymers. 96,97,100,103,104 These strategies promote a steady rate of transfection over time by stabilizing the polyplex and protecting the siRNA from degradation. The same fusion peptides found within glycoproteins that enable viruses to escape from the endosome can provide endosomolytic function to polymeric vectors. Naturally derived and synthetic cell penetrating peptides (CPPs) are often used to enhance internalization.96,105–107 Although these peptides have been in use for decades,108,109 their mechanism of entry is not fully understood. Endocytosis appears to be the primary route when CPP concentrations are low, but route of entry, delivery efficiency, and toxicity have been shown to depend upon polymer concentration, binding method, cell type, and disease state. Rational vector design would benefit from studies systematically comparing

ACS Paragon Plus Environment

Page 22 of 50

Page 23 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

CPP-mediated delivery systems within a common cell type to elucidate the nature of peptide-mediated delivery and toxicity patterns.105 Functionality is preserved when CPPs are non-covalently complexed to siRNA or conjugated to polymers, and the duration of action can be prolonged either by conformational stabilization or the use of unnatural amino acids. The cationic nature of CPPs typically necessitates shielding as well as targeting - either through environmentally responsive polymer design or the addition of targeting ligands - for in vivo applications.105 In one study, amphipathic CPPs provided silencing comparable to Lipofectamine 2000, vastly outperforming cationic CPPs.110 In many cases, CPPs have not elicited an innate immune response despite their frequent derivation from non-human proteins; however, there have been some notable exceptions (e.g. the innate immune response observed when a Penetratin-siRNA complex was administered intratracheally to mice).111,112 To address the risks of anti-CPP responses and immunological memory, researchers in the Langer lab recently employed a systematic proteomic study of 50 sequences and identified 3 novel human-derived CPPs that facilitated enhanced siRNA delivery and reduced immunostimulation in vivo.113

Endosomal Escape When endocytosis is the primary uptake mechanism, endosomal escape is a critical, and often rate limiting, step.114 Environmentally responsive systems,115 fusogenic peptides, membranedestabilizing polymers, and the “proton sponge” effect are frequently used to address this challenge. As previously discussed, some CPPs also exhibit a pH sensitive lytic transition (e.g. EB1, MALA, KALA) 105,116,117 when random coils transition to α-helices near pH 5, and this property has been further enhanced by cysteine conjugation resulting in disulfide bond formation,118 making them candidates for dual functionalization.119 Other cell-penetrating peptides (such as hCT) have been added to polymeric

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

carriers solely for endosomolysis, and have even been modified (e.g. by the conjugation of fatty acids and peptide sequences) to serve as delivery vehicles in the absence of polymers. 61 If researchers continue to discover novel, human derived CPPs at the current rate, the list of pH sensitive peptides is likely to increase substantially in the years ahead. Maleic acid amide (MAA) has often been used as a linker which hydrolyzes rapidly at endosomal pH levels. Recently, Kataoka’s group used Click chemistry and this anionic moiety to link siRNA to an endosome-disrupting polyaspartamide derivative with two repeating units of aminoethylene per side chain (termed PAsp(DET)), and in so doing, converted cationic cites to anions - rendering the system biologically inert. Delivery of siRNA was enhanced while the IFN-α response was significantly reduced in the murine macrophage cell line (Raw264.7) to 24.3±3.5 pg/mL vs. 60.8±12.9 pg/mL for the same polycation lacking the MAA linkage, but in vivo tests are needed to confirm the in vitro findings (Figure 10).41 Similarly, ketal linkages rapidly hydrolyze under acidic conditions, and have been used in vitro with a poly(β-amino ester) backbone to form a non-toxic, non-immunogenic carrier that delivers siRNA more efficiently than PEI.42,120

ACS Paragon Plus Environment

Page 24 of 50

Page 25 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

Figure 10. a) Illustration of releasable/enzyme-disrupting conjugate (REC) with the multifunctionality toward endosomal escape and release of mono-siRNA. b) Chemical structure of REC. c) Chemical structure of uREC. The PAsp derivative in this study has the mixed sequence of α and β isomers. Only α isomers are depicted in (b) and (c) for simplicity. Figure and caption reprinted with permission from Takemoto, et al.41 © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Although the validity and nature of the “proton sponge” effect is debated, amine groups are commonly incorporated into polymeric vectors with the goal of exploiting this method of endosomal escape.42,94,121–123 Histamines have been used, in combination with primary amines, to potentially

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

increase the buffering capacity of micelles, cSCKs, and dynamic polyconjugates, resulting in improved endosomal escape. Notably, transfection efficiency was lost when histamine conjugations replaced half of the primary amines, and histamines could play a role in the fusion of lipid bilayers under acidic conditions, resulting in lysis in the absence of a “proton sponge” effect.124,125 Recently, nanoparticlebased pH sensors with a dynamic range from pH 3.2-7.0 have been developed and used to demonstrate a lack of endosomal acidification in the presence of PEI.126 As new sensors are developed and computational modeling becomes an increasingly powerful way to predict complex interactions, it is reasonable to expect a more thorough, mechanistic understanding of endosomal escape pathways to inform polymer design in coming years.

Unpackaging Finally, once in the cytosol, the carrier must dissociate from the siRNA for RISC to become active. Hydrolyzable esters, acetal bonds and reducible disulfide bonds are perhaps the most common means for ensuring cytosolic release, and when incorporated into the polymeric backbone, they enable large polymers to be cleaved into segments of a biocompatible molecular weight and charge density.125,127,128 In recent years, investigators tuned the bioreducibility of polymeric vectors along with other physiochemical properties (e.g. hydrophobicity) to design stable systems that commonly selfassemble and provide significant rates of protein knockdown with reduced cytotoxicity over a time span ranging from minutes to weeks.47,49,94,129,130 Interestingly, recent experiments with protein have shown that the redox-responsiveness of disulfide bonds is strongly influenced by the proximity of electrostatic charge as well as steric hindrance. Researchers have demonstrated the potential to manipulate these factors to control disulfide bond stability and optimize intracellular targeting and timed release based on pH. The ability to precisely alter disulfide stability (by approx. three orders of magnitude) within neutral pH environments and complex

ACS Paragon Plus Environment

Page 26 of 50

Page 27 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

redox buffers (e.g. serum) is a significant achievement with the potential to be useful beyond proteins and into polymer vehicles.131 Other interesting advancements in cytosolic delivery include a recent study in which the 10-fold difference between intracellular and extracellular ATP levels was used to trigger the release of siRNA from polyplex micelles. Katakoa et al. used 3-fluoro-4-carboxyphenylboronic acid phenylboronic acid (FPBA) to quantitatively modify the cationic PEG-block-poly(L-lysine) copolymer. This yielded a delivery vector in which boronic acid formed a reversible covalent ester bond with the siRNA, and PEG enabled the formation of micelles upon complexation with siRNA. This system has provided a dose-dependent silencing of the proto-oncogene pol-like kinase 1 (PLK-1) in vitro.132 In recent years, quantum dot mediated Förster resonance energy transfer (QD-FRET) has been used to study the intracellular trafficking and unpacking of ionically associated constructs, in vitro. Using this approach, Lee et al. quantitatively analyzed the dissociation of siRNA from branched PEI, with and without conjugation to human transcriptional factor (Hph-1), within live cells under flow cytometry.133 Rather than using covalently labeling, Endres et al. used physical entrapment of hydrophobic QDs within the PCL core of PEG-PCL-PEI and demonstrated the feasibility of FRET-switching for the prospective study of polymer-siRNA dissociation within the cytosol.134 In the future, similar QD-FRET studies could be employed across a library of polymers and a variety of cell types in order to build a more robust structure-activity relationship between chemical structure and dissociation kinetics, thereby facilitating the design of the next generation of polymers for siRNA delivery.

Computational techniques for simulating polymer/siRNA interactions Despite an abundance of experimental work, many critical steps in siRNA delivery are not fully understood at a mechanistic level, including polymer/siRNA particle formation, cellular uptake, intracellular trafficking pathways and payload release. In recent years, innovative computational

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

strategies have provided new ways to approach these problems as researchers have looked to multiscale modeling bolstered by uncertainty quantification and molecular field theory to elucidate key physical mechanisms for the purposes of designing more efficient carriers. In silico studies have been both validated and complemented by in vitro and in vivo experiments resulting in new mechanistic insights and model systems that can enable a more quantitative evaluation of the structure-activity relationships that might lead to of the next generation of multi-component polymer vectors. Atomistic simulations, coarse grained molecular dynamics, monomer-resolved simulations, and immersed molecular electrokinetic finite element method (IMEFEM) have been successfully combined in a multiscale computational modeling framework to lower computational costs for the exploration of key parameters (e.g. a wide range of sizes, shapes, surface properties, rigidities, flow parameters, physiological conditions, etc.) while quantitatively assessing polymer-siRNA complexation and dissociation, margination dynamics, extravasation, and cellular uptake. When followed by experimental validation, such simulations have become a powerful instrument to enhance mechanistic understanding and drive the rational design of more efficient delivery vectors.135,136 Recent advancements in computational modeling have been reviewed for dendrimeric vectors136 as well as polyplexes and lipoplexes.137,138 Purely in silico studies have shown lipid substitution on PEI impairs neither charge neutralization nor the formation of polyion bridges (critical for complexation), that longer lipid chains have the potential to further stabilize and compact polyplexes as a result of lipid associations139, and that longer ligands on nanoparticles can enhance attachment to cell membranes at the cost of internalization efficiency unless careful attention is paid to side chain density, rigidity and hydrophobic/hydrophilic balance.140 When combined with isothermal titration calorimetry, the importance of minimizing the N/P ratios to achieve efficient cationic polymeric vectors with reduced cytotoxicity has been highlighted,

ACS Paragon Plus Environment

Page 28 of 50

Page 29 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

with a thermodynamic assessment of polyplex self-assembly revealing the optimum N/P ratio for cationic polymers to be lower than previously expected.22 The field has benefited from the integration of in silico studies within the traditional framework of polymeric delivery studies. When compared to prior in vitro studies, modeling revealed several energetic aspects of dendrimer-siRNA interactions with increased intermolecular hydrogen bonds affecting binding enthalpy and increased polymer flexibility decreasing the entropic penalty of complexation. Furthermore, the length, composition and flexibility of sticky siRNA overhangs greatly affected silencing ability with rigid (dA)5-7 overhangs enhancing binding capacity but flexible (dT)5-7 ends favoring the unbinding necessary for delivery.23,141 In vitro and in vivo studies supported the comparative studies of bolaamphiphiles in silico,142 and confirmed that the formation of tight complexes with proteins depended upon maintaining an arginine/phosphate charge ratio near unity rather than molar ratio or peptide length alone.143 The potential for such computational strategies remains under investigation. For example, time and length scales of atomistic simulations are currently limited by computational power; therefore, coarse grained simulations are used to model larger nanoparticles. Model scales and accuracy are expected to continue to improve as superior coarse grained simulations are developed and computational power continues to increase.137 Novel computational strategies promise to play a prominent role in resolving yet-poorly understood aspects of the delivery process in the years ahead.

Conclusion, Challenges and Future Directions In recent years, the field of polymeric siRNA delivery has made strides toward greater clinical viability largely due to the development of novel chemistries and a more thorough understanding of both cellular processes and polymer interactions at the biological interface, resulting in more nuanced understanding of structure/activity relationships. While polymers comprise only a small percentage of

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the delivery strategies currently undergoing clinical trials, the entire field of siRNA delivery is still maturing. Recent advances in polymer chemistry have enabled the design and synthesis of highly varied architectures with unprecedented consistency, and the computational capabilities available to research today make in silico screening of vast libraries possible. Additional experimental observations will be required in order to build predictive computational models, but imaging techniques are now allowing for more precise intracellular tracking, and 3D human cell culture is providing a superior means for conducting the in vitro studies needed when building and validating such models. These factors, combined with lessons learned from the fields of nanotechnology and protein chemistry - all make this an excellent time for researchers to more thoroughly evaluate the clinical viability of polymeric systems. The challenges are considerable, but so too are the possibilities to improve quality of life for many through effective and affordable siRNA delivery. As discussed previously, endosomal escape has been widely cited as a rate-limiting step, and systemic studies are still needed to elucidate the nature of peptide-mediated delivery and toxicity patterns. Further work is needed to quantitatively and qualitatively assess factors such as immunostimulatory effects, protein interactions, binding affinity, etc. A thorough assessment will require the collaborative efforts of many groups, but the use of in silico strategies and the adoption of standardized SOPs, cell lines and controls will reduce the number of hours required while facilitating the meta-analysis of resulting data. Only once polymeric vectors have been systematically studied on a large scale – with an eye toward structure-activity relationships, interactions at the biological interface, and novel design criteria – will we truly understand their potential.

ACS Paragon Plus Environment

Page 30 of 50

Page 31 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

Bibliography: (1)

Fire, A.; Xu, S.; Montgomery, M.; Kostas, S.; Driver, S.; Mello, C. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature 1998, 391 (February), 806–811.

(2)

Misra, S.; Ghatak, S.; Toole, B. P. Regulation of MDR1 expression and drug resistance by a positive feedback loop involving hyaluronan, phosphoinositide 3-kinase, and ErbB2. J. Biol. Chem. 2005, 280 (21), 20310–20315 DOI: 10.1074/jbc.M500737200.

(3)

Rinkenauer, A. C.; Vollrath, A.; Schallon, A.; Tauhardt, L.; Kempe, K.; Schubert, S.; Fischer, D.; Schubert, U. S. Parallel high-throughput screening of polymer vectors for nonviral gene delivery: evaluation of structure-property relationships of transfection. ACS Comb. Sci. 2013, 15 (9), 475– 482 DOI: 10.1021/co400025u.

(4)

Kohane, D. S.; Langer, R. Biocompatibility and drug delivery systems. Chem. Sci. 2010, 1 (4), 441 DOI: 10.1039/c0sc00203h.

(5)

Whitehead, K. A.; Matthews, J.; Chang, P. H.; Niroui, F.; Dorkin, J. R.; Severgnini, M.; Anderson, D. G. In Vitro À In Vivo Translation of Lipid Nanoparticles for Hepatocellular siRNA Delivery. 2012, No. 8, 6922–6929.

(6)

de Fougerolles, A. R. Delivery vehicles for small interfering RNA in vivo. Hum. Gene Ther. 2008, 19 (2), 125–132 DOI: 10.1089/hum.2008.928.

(7)

Mendyk, A.; Tuszyński, P. K.; Polak, S.; Jachowicz, R. Generalized in vitro-in vivo relationship (IVIVR) model based on artificial neural networks. Drug Des. Devel. Ther. 2013, 7, 223–232 DOI: 10.2147/DDDT.S41401.

(8)

Zuckerman, J. E.; Gritli, I.; Tolcher, A.; Heidel, J. D.; Lim, D.; Morgan, R.; Chmielowski, B.; Ribas, A.; Davis, M. E.; Yen, Y. Correlating animal and human phase Ia/Ib clinical data with CALAA-01, a targeted, polymer-based nanoparticle containing siRNA. Proc. Natl. Acad. Sci. U. S. A. 2014, 111

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(31), 11449–11454 DOI: 10.1073/pnas.1411393111. (9)

Ferruti, P.; Bettelli, A.; Fere, A. High polymers of acrylic and methacrylic esters of Nhydroxysuccinimide as polyacrylamide and polymethacrylamide precursors. Polymer (Guildf). 1972, 13 (10), 462–464 DOI: 10.1016/0032-3861(72)90084-5.

(10)

Godwin, A.; Hartenstein, M.; Müller, A. H. E.; Brocchini, S. Narrow molecular weight distribution precursors for polymer - Drug conjugates. Angew. Chemie - Int. Ed. 2001, 40 (3), 594–597 DOI: 10.1002/1521-3773(20010202)40:33.0.CO;2-P.

(11)

Monge, S.; Haddleton, D. M. Synthesis of precursors of poly(acryl amides) by copper mediated living radical polymerization in DMSO. Eur. Polym. J. 2004, 40 (1), 37–45 DOI: 10.1016/j.eurpolymj.2003.08.003.

(12)

Wong, S. Y.; Putnam, D. Overcoming limiting side rxns associated w NHS-activated precursor polymethacrylamide-based polymer. Bioconjugate Chem. 2007, 18 (3), 970–982 DOI: 10.1021/bc0603790.

(13)

Xue, L.; Ingle, N. P.; Reineke, T. M. Highlighting the role of polymer length, carbohydrate size, and nucleic acid type in potency of glycopolycation agents for pDNA and siRNA delivery. Biomacromolecules 2013, 14 (11), 3903–3915 DOI: 10.1021/bm401026n.

(14)

Zhang, Y.; Lundberg, P.; Diether, M.; Porsch, C.; Janson, C.; Lynd, N. a.; Ducani, C.; Malkoch, M.; Malmström, E.; Hawker, C. J.; et al. Histamine-functionalized copolymer micelles as a drug delivery system in 2D and 3D models of breast cancer. J. Mater. Chem. B 2015, 3 (12), 2472–2486 DOI: 10.1039/C4TB02051K.

(15)

Siegwart, D. J.; Whitehead, K. A.; Nuhn, L.; Sahay, G.; Cheng, H.; Jiang, S.; Ma, M. Combinatorial synthesis of chemically diverse core-shell nanoparticles for intracellular delivery. 2011 DOI: 10.1073/pnas.1106379108//DCSupplemental.www.pnas.org/cgi/doi/10.1073/pnas.1106379108.

ACS Paragon Plus Environment

Page 32 of 50

Page 33 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

(16)

Nascimento, A. V.; Singh, A.; Bousbaa, H.; Ferreira, D.; Sarmento, B.; Amiji, M. M. CombinatorialDesigned Epidermal Growth Factor Receptor-Targeted Chitosan Nanoparticles for Encapsulation and Delivery of Lipid-Modified Platinum Derivatives in Wild-Type and Resistant Non-Small-Cell Lung Cancer Cells. Mol. Pharm. 2015, 12 (12), 4466–4477 DOI: 10.1021/acs.molpharmaceut.5b00642.

(17)

Pelet, J. M.; Putnam, D. A combinatorial library of Bi-functional polymeric vectors for siRNA delivery in vitro. Pharm. Res. 2013, 30 (2), 362–376 DOI: 10.1007/s11095-012-0876-4.

(18)

Jones, C. H.; Chen, C.-K.; Ravikrishnan, A.; Rane, S.; Pfeifer, B. a. Overcoming nonviral gene delivery barriers: perspective and future. Mol. Pharm. 2013, 10 (11), 4082–4098 DOI: 10.1021/mp400467x.

(19)

Dandekar, P.; Jain, R.; Keil, M.; Loretz, B.; Muijs, L.; Schneider, M.; Auerbach, D.; Jung, G.; Lehr, C.M.; Wenz, G. Cellular delivery of polynucleotides by cationic cyclodextrin polyrotaxanes. J. Control. Release 2012, 164 (3), 387–393 DOI: 10.1016/j.jconrel.2012.06.040.

(20)

Liu, X.; Liu, C.; Laurini, E.; Posocco, P.; Pricl, S.; Qu, F.; Rocchi, P.; Peng, L. Efficient delivery of sticky siRNA and potent gene silencing in a prostate cancer model using a generation 5 triethanolamine-core PAMAM dendrimer. Mol. Pharm. 2012, 9 (3), 470–481 DOI: 10.1021/mp2006104.

(21)

Han, L.; Tang, C.; Yin, C. Effect of binding affinity for siRNA on the in vivo antitumor efficacy of polyplexes. Biomaterials 2013, 34 (21), 5317–5327 DOI: 10.1016/j.biomaterials.2013.03.060.

(22)

Zheng, M.; Pavan, G. M.; Neeb, M.; Schaper, A. K.; Danani, A.; Klebe, G.; Merkel, O. M.; Kissel, T. Targeting the blind spot of polycationic nanocarrier-based siRNA delivery. ACS Nano 2012, 6 (11), 9447–9454 DOI: 10.1021/nn301966r.

(23)

Karatasos, K.; Posocco, P.; Laurini, E.; Pricl, S. Poly(amidoamine)-based dendrimer/siRNA complexation studied by computer simulations: effects of pH and generation on dendrimer

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

structure and siRNA binding. Macromol. Biosci. 2012, 12 (2), 225–240 DOI: 10.1002/mabi.201100276. (24)

Ulasov, A. V; Khramtsov, Y. V; Trusov, G. a; Rosenkranz, A. a; Sverdlov, E. D.; Sobolev, A. S. Properties of PEI-based polyplex nanoparticles that correlate with their transfection efficacy. Mol. Ther. 2011, 19 (1), 103–112 DOI: 10.1038/mt.2010.233.

(25)

Naeye, B.; Deschout, H.; Caveliers, V.; Descamps, B.; Braeckmans, K.; Vanhove, C.; Demeester, J.; Lahoutte, T.; De Smedt, S. C.; Raemdonck, K. In vivo disassembly of IV administered siRNA matrix nanoparticles at the renal filtration barrier. Biomaterials 2013, 34 (9), 2350–2358 DOI: 10.1016/j.biomaterials.2012.11.058.

(26)

Zuckerman, J. E.; Choi, C. H. J.; Han, H.; Davis, M. E. Polycation-siRNA nanoparticles can disassemble at the kidney glomerular basement membrane. Proc. Natl. Acad. Sci. U. S. A. 2012, 109 (8), 3137–3142 DOI: 10.1073/pnas.1200718109.

(27)

Tenzer, S.; Docter, D.; Kuharev, J.; Musyanovych, A.; Fetz, V.; Hecht, R.; Schlenk, F.; Fischer, D.; Kiouptsi, K.; Reinhardt, C.; et al. Rapid formation of plasma protein corona critically affects nanoparticle pathophysiology. Nat. Nanotechnol. 2013, 8 (10), 772–781 DOI: 10.1038/nnano.2013.181.

(28)

Zhong, D.; Jiao, Y.; Zhang, Y.; Zhang, W.; Li, N.; Zuo, Q.; Wang, Q.; Xue, W.; Liu, Z. Effects of the gene carrier polyethyleneimines on structure and function of blood components. Biomaterials 2013, 34 (1), 294–305 DOI: 10.1016/j.biomaterials.2012.09.060.

(29)

Li, C.; Zhong, D.; Zhang, Y.; Tuo, W.; Li, N.; Wang, Q.; Liu, Z.; Xue, W. The effect of the gene carrier material polyethyleneimine on the structure and function of human red blood cells in vitro. J. Mater. Chem. B 2013, 1 (14), 1885 DOI: 10.1039/c3tb00024a.

(30)

Perry, J. L.; Reuter, K. G.; Kai, M. P.; Herlihy, K. P.; Jones, S. W.; Luft, J. C.; Napier, M.; Bear, J. E.; DeSimone, J. M. PEGylated PRINT nanoparticles: the impact of PEG density on protein binding,

ACS Paragon Plus Environment

Page 34 of 50

Page 35 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

macrophage association, biodistribution, and pharmacokinetics. Nano Lett. 2012, 12 (10), 5304– 5310 DOI: 10.1021/nl302638g. (31)

Kunath, K.; von Harpe, A.; Petersen, H.; Fischer, D.; Voigt, K.; Kissel, T.; Bickel, U. The Structure of PEG-Modified Poly ( Ethylene Imines ) Influences Biodistribution and Pharmacokinetics of Their Complexes with NF- kapa B Decoy in Mice. Pharm. Res. 2002, 19 (6), 810–817.

(32)

Zheng, M.; Librizzi, D.; Kılıç, A.; Liu, Y.; Renz, H.; Merkel, O. M.; Kissel, T. Enhancing in vivo circulation and siRNA delivery with biodegradable polyethylenimine-graft-polycaprolactoneblock-poly(ethylene glycol) copolymers. Biomaterials 2012, 33 (27), 6551–6558 DOI: 10.1016/j.biomaterials.2012.05.055.

(33)

Zhao, Y.; Qin, Y.; Liang, Y.; Zou, H.; Peng, X.; Huang, H.; Lu, M.; Feng, M. Salt-induced stability and serum-resistance of polyglutamate polyelectrolyte brushes/nuclear factor-κB p65 siRNA Polyplex enhance the apoptosis and efficacy of doxorubicin. Biomacromolecules 2013, 14 (6), 1777–1786 DOI: 10.1021/bm400177q.

(34)

Salvati, A.; Pitek, A. S.; Monopoli, M. P.; Prapainop, K.; Bombelli, F. B.; Hristov, D. R.; Kelly, P. M.; Åberg, C.; Mahon, E.; Dawson, K. a. Transferrin-functionalized nanoparticles lose their targeting capabilities when a biomolecule corona adsorbs on the surface. Nat. Nanotechnol. 2013, 8 (2), 137–143 DOI: 10.1038/nnano.2012.237.

(35)

Wang, F.; Yu, L.; Monopoli, M. P.; Sandin, P.; Mahon, E.; Salvati, A.; Dawson, K. a. The biomolecular corona is retained during nanoparticle uptake and protects the cells from the damage induced by cationic nanoparticles until degraded in the lysosomes. Nanomedicine 2013, 9 (8), 1159–1168 DOI: 10.1016/j.nano.2013.04.010.

(36)

Kanasty, R. L.; Whitehead, K. a; Vegas, A. J.; Anderson, D. G. Action and reaction: the biological response to siRNA and its delivery vehicles. Mol. Ther. 2012, 20 (3), 513–524 DOI: 10.1038/mt.2011.294.

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(37)

Whitehead, K. a; Dahlman, J. E.; Langer, R. S.; Anderson, D. G. Silencing or stimulation? siRNA delivery and the immune system. Annu. Rev. Chem. Biomol. Eng. 2011, 2, 77–96 DOI: 10.1146/annurev-chembioeng-061010-114133.

(38)

Forsbach, A.; Müller, C.; Montino, C.; Kritzler, A.; Curdt, R.; Benahmed, A.; Jurk, M.; Vollmer, J. Impact of delivery systems on siRNA immune activation and RNA interference. Immunol. Lett. 2012, 141 (2), 169–180 DOI: 10.1016/j.imlet.2011.10.001.

(39)

Moyano, D. F.; Goldsmith, M.; Solfiell, D. J.; Landesman-Milo, D.; Miranda, O. R.; Peer, D.; Rotello, V. M. Nanoparticle hydrophobicity dictates immune response. J. Am. Chem. Soc. 2012, 134 (9), 3965–3967 DOI: 10.1021/ja2108905.

(40)

Olejniczak, M.; Galka-marciniak, P.; Polak, K.; Fligier, A. RNAimmuno : A database of the nonspecific immunological effects of RNA interference and microRNA reagents. 2012, 930–935 DOI: 10.1261/rna.025627.110.2009.

(41)

Takemoto, H.; Miyata, K.; Hattori, S.; Ishii, T.; Suma, T.; Uchida, S.; Nishiyama, N.; Kataoka, K. Acidic pH-Responsive siRNA Conjugate for Reversible Carrier Stability and Accelerated Endosomal Escape with Reduced IFNα-Associated Immune Response. Angew. Chemie 2013, 125 (24), 6338– 6341 DOI: 10.1002/ange.201300178.

(42)

Guk, K.; Lim, H.; Kim, B.; Hong, M.; Khang, G.; Lee, D. Acid-cleavable ketal containing poly(βamino ester) for enhanced siRNA delivery. Int. J. Pharm. 2013, 453 (2), 541–550 DOI: 10.1016/j.ijpharm.2013.06.021.

(43)

Kulkarni, A.; DeFrees, K.; Hyun, S.-H.; Thompson, D. H. Pendant polymer:amino-βcyclodextrin:siRNA guest:host nanoparticles as efficient vectors for gene silencing. J. Am. Chem. Soc. 2012, 134 (18), 7596–7599 DOI: 10.1021/ja300690j.

(44)

Dong, D.-W.; Xiang, B.; Gao, W.; Yang, Z.-Z.; Li, J.-Q.; Qi, X.-R. pH-responsive complexes using prefunctionalized polymers for synchronous delivery of doxorubicin and siRNA to cancer cells.

ACS Paragon Plus Environment

Page 36 of 50

Page 37 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

Biomaterials 2013, 34 (20), 4849–4859 DOI: 10.1016/j.biomaterials.2013.03.018. (45)

Kim, J. S.; Oh, M. H.; Park, J. Y.; Park, T. G.; Nam, Y. S. Protein-resistant, reductively dissociable polyplexes for in vivo systemic delivery and tumor-targeting of siRNA. Biomaterials 2013, 34 (9), 2370–2379 DOI: 10.1016/j.biomaterials.2012.12.004.

(46)

Saito, G.; Swanson, J. a; Lee, K.-D. Drug delivery strategy utilizing conjugation via reversible disulfide linkages: role and site of cellular reducing activities. Adv. Drug Deliv. Rev. 2003, 55 (2), 199–215.

(47)

Dunn, S. S.; Tian, S.; Blake, S.; Wang, J.; Galloway, A. L.; Murphy, A.; Pohlhaus, P. D.; Rolland, J. P.; Napier, M. E.; DeSimone, J. M. Reductively responsive siRNA-conjugated hydrogel nanoparticles for gene silencing. J. Am. Chem. Soc. 2012, 134 (17), 7423–7430 DOI: 10.1021/ja300174v.

(48)

Namgung, R.; Kim, W. J. A highly entangled polymeric nanoconstruct assembled by siRNA and its reduction-triggered siRNA release for gene silencing. Small 2012, 8 (20), 3209–3219 DOI: 10.1002/smll.201200496.

(49)

Lin, D.; Jiang, Q.; Cheng, Q.; Huang, Y.; Huang, P.; Han, S.; Guo, S.; Liang, Z.; Dong, A. Polycationdetachable nanoparticles self-assembled from mPEG-PCL-g-SS-PDMAEMA for in vitro and in vivo siRNA delivery. Acta Biomater. 2013, 9 (8), 7746–7757 DOI: 10.1016/j.actbio.2013.04.031.

(50)

Tan, C.; Wang, Y.; Fan, W. Exploring polymeric micelles for improved delivery of anticancer agents: recent developments in preclinical studies. Pharmaceutics 2013, 5 (1), 201–219 DOI: 10.3390/pharmaceutics5010201.

(51)

Wang, H.-X.; Xiong, M.-H.; Wang, Y.-C.; Zhu, J.; Wang, J. N-acetylgalactosamine functionalized mixed micellar nanoparticles for targeted delivery of siRNA to liver. J. Control. Release 2013, 166 (2), 106–114 DOI: 10.1016/j.jconrel.2012.12.017.

(52)

Wooddell, C. I.; Rozema, D. B.; Hossbach, M.; John, M.; Hamilton, H. L.; Chu, Q.; Hegge, J. O.; Klein, J. J.; Wakefield, D. H.; Oropeza, C. E.; et al. Hepatocyte-targeted RNAi therapeutics for the

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 38 of 50

treatment of chronic hepatitis B virus infection. Mol. Ther. 2013, 21 (5), 973–985 DOI: 10.1038/mt.2013.31. (53)

Wang, J.; Dou, B.; Bao, Y. Efficient targeted pDNA/siRNA delivery with folate-low-molecularweight polyethyleneimine-modified pullulan as non-viral carrier. Mater. Sci. Eng. C. Mater. Biol. Appl. 2014, 34, 98–109 DOI: 10.1016/j.msec.2013.08.035.

(54)

Dohmen, C.; Fröhlich, T.; Lächelt, U.; Röhl, I.; Vornlocher, H.-P.; Hadwiger, P.; Wagner, E. Defined Folate-PEG-siRNA Conjugates for Receptor-specific Gene Silencing. Mol. Ther. Nucleic Acids 2012, 1 (November 2011), e7 DOI: 10.1038/mtna.2011.10.

(55)

Arima, H.; Yoshimatsu, A.; Ikeda, H.; Ohyama, A.; Motoyama, K.; Higashi, T.; Tsuchiya, A.; Niidome, T.; Katayama, Y.; Hattori, K.; et al. Folate-PEG-appended dendrimer conjugate with αcyclodextrin as a novel cancer cell-selective siRNA delivery carrier. Mol. Pharm. 2012, 9 (9), 2591– 2604 DOI: 10.1021/mp300188f.

(56)

Cheng, D.; Cao, N.; Chen, J.; Yu, X.; Shuai, X. Multifunctional nanocarrier mediated co-delivery of doxorubicin and siRNA for synergistic enhancement of glioma apoptosis in rat. Biomaterials 2012, 33 (4), 1170–1179 DOI: 10.1016/j.biomaterials.2011.10.057.

(57)

Novo, L.; Mastrobattista, E.; van Nostrum, C. F.; Hennink, W. E. Targeted Decationized Polyplexes for Cell Specific Gene Delivery. Bioconjug. Chem. 2014 DOI: 10.1021/bc500074a.

(58)

Cheng, C. J.; Saltzman, W. M. Enhanced siRNA delivery into cells by exploiting the synergy between targeting ligands and cell-penetrating peptides. Biomaterials 2011, 32 (26), 6194–6203 DOI: 10.1016/j.biomaterials.2011.04.053.

(59)

Zhang, C. Y.; Kos, P.; Müller, K.; Schrimpf, W.; Troiber, C.; Lächelt, U.; Scholz, C.; Lamb, D. C.; Wagner, E. Native chemical ligation for conversion of sequence-defined oligomers into targeted pDNA and siRNA carriers. J. Control. Release 2014, 180, 42–50 DOI: 10.1016/j.jconrel.2014.02.015.

ACS Paragon Plus Environment

Page 39 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

(60)

Nakagawa, O.; Ming, X.; Carver, K.; Juliano, R. Conjugation with Receptor-Targeted Histidine-Rich Peptides Enhances the Pharmacological Effectiveness of Antisense Oligonucleotides. Bioconjug. Chem. 2013 DOI: 10.1021/bc400500h.

(61)

Hoyer, J.; Neundorf, I. Knockdown of a G protein-coupled receptor through efficient peptidemediated siRNA delivery. J. Control. Release 2012, 161 (3), 826–834 DOI: 10.1016/j.jconrel.2012.05.017.

(62)

Hayashi, Y.; Mori, Y.; Higashi, T.; Motoyama, K.; Jono, H.; Sah, D. W. Y.; Ando, Y.; Arima, H. Systemic delivery of transthyretin siRNA mediated by lactosylated dendrimer/α-cyclodextrin conjugates into hepatocyte for familial amyloidotic polyneuropathy therapy. Amyloid 2012, 19 Suppl 1 (March), 47–49 DOI: 10.3109/13506129.2012.674581.

(63)

Berguig, G. Y.; Convertine, A. J.; Shi, J.; Palanca-Wessels, M. C.; Duvall, C. L.; Pun, S. H.; Press, O. W.; Stayton, P. S. Intracellular delivery and trafficking dynamics of a lymphoma-targeting antibody-polymer conjugate. Mol. Pharm. 2012, 9 (12), 3506–3514 DOI: 10.1021/mp300338s.

(64)

Samarajeewa, S.; Ibricevic, A.; Gunsten, S. P.; Shrestha, R.; Elsabahy, M.; Brody, S. L.; Wooley, K. L. Degradable cationic shell cross-linked knedel-like nanoparticles: synthesis, degradation, nucleic acid binding, and in vitro evaluation. Biomacromolecules 2013, 14 (4), 1018–1027 DOI: 10.1021/bm3018774.

(65)

Juliano, R. L.; Carver, K.; Cao, C.; Ming, X. Receptors, endocytosis, and trafficking: the biological basis of targeted delivery of antisense and siRNA oligonucleotides. J. Drug Target. 2013, 21 (1), 27–43 DOI: 10.3109/1061186X.2012.740674.

(66)

Zhu, L.; Torchilin, V. P. Stimulus-responsive nanopreparations for tumor targeting. Integr. Biol. (Camb). 2013, 5 (1), 96–107 DOI: 10.1039/c2ib20135f.

(67)

Alam, M. R.; Ming, X.; Nakagawa, O.; Jin, J.; Juliano, R. L. Covalent conjugation of oligonucleotides with cell-targeting ligands. Bioorg. Med. Chem. 2013, 21 (20), 6217–6223 DOI:

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

10.1016/j.bmc.2013.05.037. (68)

Yao, Y.; Sun, T.; Huang, S.; Dou, S.; Lin, L.; Chen, J.; Ruan, J.; Mao, C.; Yu, F.; Zeng, M.; et al. Targeted delivery of PLK1-siRNA by ScFv suppresses Her2+ breast cancer growth and metastasis. Sci. Transl. Med. 2012, 4 (130), 130ra48 DOI: 10.1126/scitranslmed.3003601.

(69)

Lu, H.; Wang, D.; Kazane, S.; Javahishvili, T.; Tian, F.; Song, F.; Sellers, A.; Barnett, B.; Schultz, P. G. Site-specific antibody-polymer conjugates for siRNA delivery. J. Am. Chem. Soc. 2013, 135 (37), 13885–13891 DOI: 10.1021/ja4059525.

(70)

Chiu, R. Y. T.; Tsuji, T.; Wang, S. J.; Wang, J.; Liu, C. T.; Kamei, D. T. Improving the systemic drug delivery efficacy of nanoparticles using a transferrin variant for targeting. J. Control. Release 2014, 180C, 33–41 DOI: 10.1016/j.jconrel.2014.01.027.

(71)

Ren, Y.; Hauert, S.; Lo, J. H.; Bhatia, S. N. Identi fi cation and Characterization of Receptor-Speci fi c Peptides for siRNA Delivery. 2012, No. 10, 8620–8631.

(72)

Pelkmans, L.; Helenius, A. Endocytosis via caveolae. Traffic 2002, 3 (5), 311–320.

(73)

Doherty, G. J.; McMahon, H. T. Mechanisms of endocytosis. Annu. Rev. Biochem. 2009, 78 (March), 857–902 DOI: 10.1146/annurev.biochem.78.081307.110540.

(74)

Khalil, I. A.; Kogure, K.; Akita, H.; Harashima, H. Uptake Pathways and Subsequent Intracellular Trafficking in Nonviral Gene Delivery. 2006, 58 (1), 32–45 DOI: 10.1124/pr.58.1.8.32.

(75)

Torchilin, V. P. Recent Approaches To Intracellular Delivery of Drugs and Dna and Organelle Targeting. Annu. Rev. Biomed. Eng. 2006, 8 (1), 343–375 DOI: 10.1146/annurev.bioeng.8.061505.095735.

(76)

Gary, D. J.; Lee, H.; Sharma, R.; Lee, J.-S.; Kim, Y.; Cui, Z. Y.; Jia, D.; Bowman, V. D.; Chipman, P. R.; Wan, L.; et al. Influence of nano-carrier architecture on in vitro siRNA delivery performance and in vivo biodistribution: polyplexes vs micelleplexes. ACS Nano 2011, 5 (5), 3493–3505 DOI: 10.1021/nn102540y.

ACS Paragon Plus Environment

Page 40 of 50

Page 41 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

(77)

Kim, J.; Sunshine, J. C.; Green, J. J. Differential Polymer Structure Tunes Mechanism of Cellular Uptake and Transfection Routes of Poly(β-amino ester) Polyplexes in Human Breast Cancer Cells. Bioconjug. Chem. 2013 DOI: 10.1021/bc4002322.

(78)

Lühmann, T.; Rimann, M.; Bittermann, A. G.; Hall, H. Cellular uptake and intracellular pathways of PLL-g-PEG-DNA nanoparticles. Bioconjug. Chem. 2008, 19 (9), 1907–1916 DOI: 10.1021/bc800206r.

(79)

Van Der Aa, M. A. E. M.; Huth, U. S.; Häfele, S. Y.; Schubert, R.; Oosting, R. S.; Mastrobattista, E.; Hennink, W. E.; Peschka-Süss, R.; Koning, G. A.; Crommelin, D. J. A. Cellular uptake of cationic polymer-DNA complexes via caveolae plays a pivotal role in gene transfection in COS-7 cells. Pharm. Res. 2007, 24 (8), 1590–1598 DOI: 10.1007/s11095-007-9287-3.

(80)

Zhang, Y.; Zhou, C.; Kwak, K. J.; Wang, X.; Yung, B.; Lee, L. J.; Wang, Y.; Wang, P. G.; Lee, R. J. Efficient siRNA delivery using a polyamidoamine dendrimer with a modified pentaerythritol core. Pharm. Res. 2012, 29 (6), 1627–1636 DOI: 10.1007/s11095-012-0676-x.

(81)

Benfer, M.; Kissel, T. Cellular uptake mechanism and knockdown activity of siRNA-loaded biodegradable DEAPA-PVA-g-PLGA nanoparticles. Eur. J. Pharm. Biopharm. 2012, 80 (2), 247–256 DOI: 10.1016/j.ejpb.2011.10.021.

(82)

Rejman, J.; Oberle, V.; Zuhorn, I. S.; Hoekstra, D. Size-dependent internalization of particles via the pathways of clathrin- and caveolae-mediated endocytosis. Biochem. J. 2004, 377 (Pt 1), 159– 169 DOI: 10.1042/BJ20031253.

(83)

Harush-Frenkel, O.; Debotton, N.; Benita, S.; Altschuler, Y. Targeting of nanoparticles to the clathrin-mediated endocytic pathway. Biochem. Biophys. Res. Commun. 2007, 353 (1), 26–32 DOI: 10.1016/j.bbrc.2006.11.135.

(84)

Lerch, S.; Dass, M.; Musyanovych, A.; Landfester, K.; Mailänder, V. Polymeric nanoparticles of different sizes overcome the cell membrane barrier. Eur. J. Pharm. Biopharm. 2013, 84 (2), 265–

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

274 DOI: 10.1016/j.ejpb.2013.01.024. (85)

Kim, D.; Hong, J.; Moon, H.-H.; Nam, H. Y.; Mok, H.; Jeong, J. H.; Kim, S. W.; Choi, D.; Kim, S. H. Anti-apoptotic cardioprotective effects of SHP-1 gene silencing against ischemia-reperfusion injury: use of deoxycholic acid-modified low molecular weight polyethyleneimine as a cardiac siRNA-carrier. J. Control. Release 2013, 168 (2), 125–134 DOI: 10.1016/j.jconrel.2013.02.031.

(86)

Chae, S. Y.; Kim, H. J.; Lee, M. S.; Jang, Y. L.; Lee, Y.; Lee, S. H.; Lee, K.; Kim, S. H.; Kim, H. T.; Chi, S.-C.; et al. Energy-independent intracellular gene delivery mediated by polymeric biomimetics of cell-penetrating peptides. Macromol. Biosci. 2011, 11 (9), 1169–1174 DOI: 10.1002/mabi.201100088.

(87)

Gratton, S. E. a; Ropp, P. a; Pohlhaus, P. D.; Luft, J. C.; Madden, V. J.; Napier, M. E.; DeSimone, J. M. The effect of particle design on cellular internalization pathways. Proc. Natl. Acad. Sci. U. S. A. 2008, 105 (33), 11613–11618 DOI: 10.1073/pnas.0801763105.

(88)

Tagalakis, A. D.; Kenny, G. D.; Bienemann, A. S.; McCarthy, D.; Munye, M. M.; Taylor, H.; Wyatt, M. J.; Lythgoe, M. F.; White, E. a; Hart, S. L. PEGylation improves the receptor-mediated transfection efficiency of peptide-targeted, self-assembling, anionic nanocomplexes. J. Control. Release 2013, 174C, 177–187 DOI: 10.1016/j.jconrel.2013.11.014.

(89)

Kim, H.-J.; Jeon, H.-K.; Cho, Y. J.; Park, Y. A.; Choi, J.-J.; Do, I.-G.; Song, S. Y.; Lee, Y.-Y.; Choi, C. H.; Kim, T.-J.; et al. High galectin-1 expression correlates with poor prognosis and is involved in epithelial ovarian cancer proliferation and invasion. Eur. J. Cancer 2012, 1–8 DOI: 10.1016/j.ejca.2012.02.005.

(90)

Guo, J.; Cheng, W. P.; Gu, J.; Ding, C.; Qu, X.; Yang, Z.; O’Driscoll, C. Systemic delivery of therapeutic small interfering RNA using a pH-triggered amphiphilic poly-L-lysine nanocarrier to suppress prostate cancer growth in mice. Eur. J. Pharm. Sci. 2012, 45 (5), 521–532 DOI: 10.1016/j.ejps.2011.11.024.

ACS Paragon Plus Environment

Page 42 of 50

Page 43 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

(91)

Kim, H. J.; Oba, M.; Pittella, F.; Nomoto, T.; Cabral, H.; Matsumoto, Y.; Miyata, K.; Nishiyama, N.; Kataoka, K. PEG-detachable cationic polyaspartamide derivatives bearing stearoyl moieties for systemic siRNA delivery toward subcutaneous BxPC3 pancreatic tumor. J. Drug Target. 2012, 20 (1), 33–42 DOI: 10.3109/1061186X.2011.632010.

(92)

Silica, D.; Peg, D. Smart Multilayered Assembly for Biocompatible siRNA Delivery Featuring. 2012, No. 8, 6693–6705.

(93)

Rose, L.; Aliabadi, H. M.; Uludağ, H. Gelatin coating to stabilize the transfection ability of nucleic acid polyplexes. Acta Biomater. 2013, 9 (7), 7429–7438 DOI: 10.1016/j.actbio.2013.03.029.

(94)

Gouda, N.; Miyata, K.; Christie, R. J.; Suma, T.; Kishimura, A.; Fukushima, S.; Nomoto, T.; Liu, X.; Nishiyama, N.; Kataoka, K. Silica nanogelling of environment-responsive PEGylated polyplexes for enhanced stability and intracellular delivery of siRNA. Biomaterials 2013, 34 (2), 562–570 DOI: 10.1016/j.biomaterials.2012.09.077.

(95)

Sizovs, A.; Xue, L.; Tolstyka, Z. P.; Ingle, N. P.; Wu, Y.; Cortez, M.; Reineke, T. M. Poly(trehalose): sugar-coated nanocomplexes promote stabilization and effective polyplex-mediated siRNA delivery. J. Am. Chem. Soc. 2013, 135 (41), 15417–15424 DOI: 10.1021/ja404941p.

(96)

Järver, P.; Coursindel, T.; Andaloussi, S. El; Godfrey, C.; Wood, M. J.; Gait, M. J. Peptide-mediated Cell and In Vivo Delivery of Antisense Oligonucleotides and siRNA. Mol. Ther. Nucleic Acids 2012, 1 (March), e27 DOI: 10.1038/mtna.2012.18.

(97)

Fröhlich, T.; Edinger, D.; Russ, V.; Wagner, E. Stabilization of polyplexes via polymer crosslinking for efficient siRNA delivery. Eur. J. Pharm. Sci. 2012, 47 (5), 914–920 DOI: 10.1016/j.ejps.2012.09.006.

(98)

Lee, S.-Y.; Huh, M. S.; Lee, S.; Lee, S. J.; Chung, H.; Park, J. H.; Oh, Y.-K.; Choi, K.; Kim, K.; Kwon, I. C. Stability and cellular uptake of polymerized siRNA (poly-siRNA)/polyethylenimine (PEI) complexes for efficient gene silencing. J. Control. Release 2010, 141 (3), 339–346 DOI:

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

10.1016/j.jconrel.2009.10.007. (99)

Troiber, C.; Edinger, D.; Kos, P.; Schreiner, L.; Kläger, R.; Herrmann, A.; Wagner, E. Stabilizing effect of tyrosine trimers on pDNA and siRNA polyplexes. Biomaterials 2013, 34 (5), 1624–1633 DOI: 10.1016/j.biomaterials.2012.11.021.

(100) Zhou, J.; Liu, J.; Shi, T.; Xia, Y.; Luo, Y.; Liang, D. Phase separation of siRNA–polycation complex and its effect on transfection efficiency. Soft Matter 2013, 9 (7), 2262 DOI: 10.1039/c2sm27498a. (101) Creusat, G.; Rinaldi, A.-S.; Weiss, E.; Elbaghdadi, R.; Remy, J.-S.; Mulherkar, R.; Zuber, G. Proton sponge trick for pH-sensitive disassembly of polyethylenimine-based siRNA delivery systems. Bioconjug. Chem. 2010, 21 (5), 994–1002 DOI: 10.1021/bc100010k. (102) Zaghloul, E. M.; Viola, J. R.; Zuber, G.; Smith, C. I. E.; Lundin, K. E. Formulation and delivery of splice-correction antisense oligonucleotides by amino acid modified polyethylenimine. Mol. Pharm. 2010, 7 (3), 652–663 DOI: 10.1021/mp900220p. (103) Akinc, A.; Zumbuehl, A.; Goldberg, M.; Leshchiner, E. S.; Busini, V.; Hossain, N.; Bacallado, S. A.; Nguyen, D. N.; Fuller, J.; Alvarez, R.; et al. A combinatorial library of lipid-like materials for delivery of RNAi therapeutics. Nat. Biotechnol. 2008, 26 (5), 561–569 DOI: 10.1038/nbt1402. (104) Love, K. T.; Mahon, K. P.; Levins, C. G.; Whitehead, K. A.; Querbes, W.; Dorkin, J. R.; Qin, J.; Cantley, W.; Qin, L. L.; Racie, T.; et al. Lipid-like materials for low-dose, in vivo gene silencing. Proc. Natl. Acad. Sci. U. S. A. 2010, 107 (5), 1864–1869 DOI: 10.1073/pnas.0910603106. (105) Lee, S. H.; Castagner, B.; Leroux, J.-C. Is there a future for cell-penetrating peptides in oligonucleotide delivery? Eur. J. Pharm. Biopharm. 2013, 85 (1), 5–11 DOI: 10.1016/j.ejpb.2013.03.021. (106) Lehto, T.; Kurrikoff, K.; Langel, Ü. Cell-penetrating peptides for the delivery of nucleic acids. Expert Opin. Drug Deliv. 2012, 9 (7), 823–836 DOI: 10.1517/17425247.2012.689285. (107) Gooding, M.; Browne, L. P.; Quinteiro, F. M.; Selwood, D. L. siRNA delivery: from lipids to cell-

ACS Paragon Plus Environment

Page 44 of 50

Page 45 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

penetrating peptides and their mimics. Chem. Biol. Drug Des. 2012, 80 (6), 787–809 DOI: 10.1111/cbdd.12052. (108) Biology, M.; Baltimore, M. Cellular uptake of the tat protein from human immunodeficiency virus. Dis. Markers 1988, 8 (1), 33–34. (109) Derossi, D.; Joliot, a H.; Chassaing, G.; Prochiantz, a. The third helix of the Antennapedia homeodomain translocates through biological membranes. J. Biol. Chem. 1994, 269 (14), 10444– 10450. (110) Mo, R. H.; Zaro, J. L.; Shen, W.-C. Comparison of cationic and amphipathic cell penetrating peptides for siRNA delivery and efficacy. Mol. Pharm. 2012, 9 (2), 299–309 DOI: 10.1021/mp200481g. (111) Carter, E.; Lau, C. Y.; Tosh, D.; Ward, S. G.; Mrsny, R. J. Cell penetrating peptides fail to induce an innate immune response in epithelial cells in vitro: implications for continued therapeutic use. Eur. J. Pharm. Biopharm. 2013, 85 (1), 12–19 DOI: 10.1016/j.ejpb.2013.03.024. (112) Moschos, S. A.; Jones, S. W.; Perry, M. M.; Williams, A. E.; Erjefalt, J. S.; Turner, J. J.; Barnes, P. J.; Sproat, B. S.; Gait, M. J.; Lindsay, M. A. Lung delivery studies using siRNA conjugated to TAT(4860) and penetratin reveal peptide induced reduction in gene expression and induction of innate immunity. Bioconjug. Chem. 18 (5), 1450–1459 DOI: 10.1021/bc070077d. (113) Karagiannis, E. D.; Urbanska, A. M.; Sahay, G.; Pelet, J. M.; Jhunjhunwala, S.; Langer, R.; Anderson, D. G. Rational design of a biomimetic cell penetrating peptide library. ACS Nano 2013, 7 (10), 8616–8626 DOI: 10.1021/nn4027382. (114) Liu, X. Q.; Sun, C. Y.; Yang, X. Z.; Wang, J. Polymeric-micelle-based nanomedicine for siRNA delivery. Part. Part. Syst. Charact. 2013, 30 (3), 211–228 DOI: 10.1002/ppsc.201200061. (115) Shim, M. S.; Kwon, Y. J. Stimuli-responsive polymers and nanomaterials for gene delivery and imaging applications. Adv. Drug Deliv. Rev. 2012, 64 (11), 1046–1059 DOI:

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

10.1016/j.addr.2012.01.018. (116) Wang, M.; Li, X.; Ma, Y.; Gu, H. Endosomal escape kinetics of mesoporous silica-based system for efficient siRNA delivery. Int. J. Pharm. 2013, 448 (1), 51–57 DOI: 10.1016/j.ijpharm.2013.03.022. (117) Li, X.; Chen, Y.; Wang, M.; Ma, Y.; Xia, W.; Gu, H. A mesoporous silica nanoparticle--PEI--fusogenic peptide system for siRNA delivery in cancer therapy. Biomaterials 2013, 34 (4), 1391–1401 DOI: 10.1016/j.biomaterials.2012.10.072. (118) Åmand, H. L.; Nordén, B.; Fant, K. Functionalization with C-terminal cysteine enhances transfection efficiency of cell-penetrating peptides through dimer formation. Biochem. Biophys. Res. Commun. 2012, 418 (3), 469–474 DOI: 10.1016/j.bbrc.2012.01.041. (119) Wagner, E. Polymers for siRNA Delivery: Inspired by Viruses to be Targeted , Dynamic , and Precise. Acc. Chem. Res. 2012, 45 (7), 1005–1013. (120) Akita, H.; Kogure, K.; Moriguchi, R.; Nakamura, Y.; Higashi, T.; Nakamura, T.; Serada, S.; Fujimoto, M.; Naka, T.; Futaki, S.; et al. Nanoparticles for ex vivo siRNA delivery to dendritic cells for cancer vaccines: programmed endosomal escape and dissociation. J. Control. Release 2010, 143 (3), 311–317 DOI: 10.1016/j.jconrel.2010.01.012. (121) Chen, C.-J.; Wang, J.-C.; Zhao, E.-Y.; Gao, L.-Y.; Feng, Q.; Liu, X.-Y.; Zhao, Z.-X.; Ma, X.-F.; Hou, W.J.; Zhang, L.-R.; et al. Self-assembly cationic nanoparticles based on cholesterol-grafted bioreducible poly(amidoamine) for siRNA delivery. Biomaterials 2013, 34 (21), 5303–5316 DOI: 10.1016/j.biomaterials.2013.03.056. (122) Fröhlich, T.; Edinger, D.; Kläger, R.; Troiber, C.; Salcher, E.; Badgujar, N.; Martin, I.; Schaffert, D.; Cengizeroglu, A.; Hadwiger, P.; et al. Structure-activity relationships of siRNA carriers based on sequence-defined oligo (ethane amino) amides. J. Control. Release 2012, 160 (3), 532–541 DOI: 10.1016/j.jconrel.2012.03.018. (123) ur Rehman, Z.; Hoekstra, D.; Zuhorn, I. S. Mechanism of polyplex- and lipoplex-mediated delivery

ACS Paragon Plus Environment

Page 46 of 50

Page 47 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

of nucleic acids: real-time visualization of transient membrane destabilization without endosomal lysis. ACS Nano 2013, 7 (5), 3767–3777 DOI: 10.1021/nn3049494. (124) Shrestha, R.; Elsabahy, M.; Florez-Malaver, S.; Samarajeewa, S.; Wooley, K. L. Endosomal escape and siRNA delivery with cationic shell crosslinked knedel-like nanoparticles with tunable buffering capacities. Biomaterials 2012, 33 (33), 8557–8568 DOI: 10.1016/j.biomaterials.2012.07.054. (125) Parmar, R. G.; Busuek, M.; Walsh, E. S.; Leander, K. R.; Howell, B. J.; Sepp-Lorenzino, L.; Kemp, E.; Crocker, L. S.; Leone, A.; Kochansky, C. J.; et al. Endosomolytic bioreducible poly(amido amine disulfide) polymer conjugates for the in vivo systemic delivery of siRNA therapeutics. Bioconjug. Chem. 2013, 24 (4), 640–647 DOI: 10.1021/bc300600a. (126) Benjaminsen, R. V; Mattebjerg, M. a; Henriksen, J. R.; Moghimi, S. M.; Andresen, T. L. The possible “proton sponge ” effect of polyethylenimine (PEI) does not include change in lysosomal pH. Mol. Ther. 2013, 21 (1), 149–157 DOI: 10.1038/mt.2012.185. (127) Xia, W.; Wang, P.; Lin, C.; Li, Z.; Gao, X.; Wang, G.; Zhao, X. Bioreducible polyethyleniminedelivered siRNA targeting human telomerase reverse transcriptase inhibits HepG2 cell growth in vitro and in vivo. J. Control. Release 2012, 157 (3), 427–436 DOI: 10.1016/j.jconrel.2011.10.011. (128) Klein, P. M.; Wagner, E. Bioreducible Polycations as Shuttles for Therapeutic Nucleic Acid and Protein Transfection. Antioxid. Redox Signal. 2014, 00 (00), 1–14 DOI: 10.1089/ars.2013.5714. (129) Human, P.; Cancer, B. Bioreducible Cationic Polymer-Based Environmentally Triggered Cytoplasmic siRNA Delivery to. 2014, No. Xx. (130) Lin, D.; Cheng, Q.; Jiang, Q.; Huang, Y.; Yang, Z.; Han, S.; Zhao, Y.; Guo, S.; Liang, Z.; Dong, A. Intracellular cleavable poly(2-dimethylaminoethyl methacrylate) functionalized mesoporous silica nanoparticles for efficient siRNA delivery in vitro and in vivo. Nanoscale 2013, 5 (10), 4291–4301 DOI: 10.1039/c3nr00294b.

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(131) Wu, C.; Wang, S.; Brülisauer, L.; Leroux, J.-C.; Gauthier, M. A. Broad Control of Disulfide Stability through Microenvironmental Effects and Analysis in Complex Redox Environments. Biomacromolecules 2013, 14 (7), 2383–2388 DOI: 10.1021/bm400501c. (132) Naito, M.; Ishii, T.; Matsumoto, A.; Miyata, K.; Miyahara, Y.; Kataoka, K. A phenylboronatefunctionalized polyion complex micelle for ATP-triggered release of siRNA. Angew. Chem. Int. Ed. Engl. 2012, 51 (43), 10751–10755 DOI: 10.1002/anie.201203360. (133) Lee, H.; Kim, I. K.; Park, T. G. Intracellular trafficking and unpacking of sirna/quantum dot-pei complexes modified with and without cell penetrating peptide: Confocal and flow cytometric fret analysis. Bioconjug. Chem. 2010, 21 (2), 289–295 DOI: 10.1021/bc900342p. (134) Endres, T.; Zheng, M.; Kılıç, A.; Turowska, A.; Beck-Broichsitter, M.; Renz, H.; Merkel, O. M.; Kissel, T. Amphiphilic Biodegradable PEG-PCL-PEI Triblock Copolymers for FRET-Capable in Vitro and in Vivo Delivery of siRNA and Quantum Dots. Mol. Pharm. 2014, 11 (4), 1273–1281 DOI: 10.1021/mp400744a. (135) Li, Y.; Stroberg, W.; Lee, T.-R.; Kim, H. S.; Man, H.; Ho, D.; Decuzzi, P.; Liu, W. K. Multiscale modeling and uncertainty quantification in nanoparticle-mediated drug/gene delivery. Comput. Mech. 2013, 53 (3), 511–537 DOI: 10.1007/s00466-013-0953-5. (136) Tian, W.; Ma, Y. Theoretical and computational studies of dendrimers as delivery vectors. Chem. Soc. Rev. 2013, 42 (2), 705–727 DOI: 10.1039/c2cs35306g. (137) Meneksedag-Erol, D.; Tang, T.; Uludağ, H. Molecular modeling of polynucleotide complexes. Biomaterials 2014, 1–9 DOI: 10.1016/j.biomaterials.2014.04.103. (138) Meneksedag-erol, D.; Sun, C.; Tang, T.; Uludag, H. Intracellular Delivery II; Prokop, A., Iwasaki, Y., Harada, A., Eds.; Fundamental Biomedical Technologies; Springer Netherlands: Dordrecht, 2014; Vol. 7. (139) Sun, C.; Tang, T.; Uludag, H. A molecular dynamics simulation study on the effect of lipid

ACS Paragon Plus Environment

Page 48 of 50

Page 49 of 50

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

substitution on polyethylenimine mediated siRNA complexation. Biomaterials 2013, 34 (11), 2822–2833 DOI: 10.1016/j.biomaterials.2013.01.011. (140) Ding, H.; Ma, Y. Role of physicochemical properties of coating ligands in receptor-mediated endocytosis of nanoparticles. Biomaterials 2012, 33 (23), 5798–5802 DOI: 10.1016/j.biomaterials.2012.04.055. (141) Posocco, P.; Liu, X.; Laurini, E.; Marson, D.; Chen, C.; Liu, C.; Fermeglia, M.; Rocchi, P.; Pricl, S.; Peng, L. Impact of siRNA overhangs for dendrimer-mediated siRNA delivery and gene silencing. Mol. Pharm. 2013, 10 (8), 3262–3273 DOI: 10.1021/mp400329g. (142) Kim, T.; Afonin, K. a; Viard, M.; Koyfman, A. Y.; Sparks, S.; Heldman, E.; Grinberg, S.; Linder, C.; Blumenthal, R. P.; Shapiro, B. a. In Silico, In Vitro, and In Vivo Studies Indicate the Potential Use of Bolaamphiphiles for Therapeutic siRNAs Delivery. Mol. Ther. Nucleic Acids 2013, 2 (October 2012), e80 DOI: 10.1038/mtna.2013.5. (143) Kim, M.; Kim, H. R.; Chae, S. Y.; Larson, R. G.; Lee, H.; Park, J. C. Effect of arginine-rich peptide length on the structure and binding strength of siRNA-peptide complexes. J. Phys. Chem. B 2013, 117 (23), 6917–6926 DOI: 10.1021/jp402868g.

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

For table of contents use only

Strategies Used to Deliver siRNA in Clinical Trials Naked siRNA (n=25)

3.3% 1.6% 4.9% 4.9% 6.6%

41.0%

6.6%

Lipid Nanoparticles (n=19) Polymers (LODER & Cyclodextrin) (n=4) siRNA-GalNAc Conjugates (n=4) Ex Vivo (n=3) Dynamic PolyConjugates (n=3) Self-Delivering (n=2) Viral (n=1)

31.1% TOC Graphic: A breakdown of the number of national clinical trials that employ each siRNA delivery strategy – source: clinicaltrials.gov Manuscript Title: Polymers for siRNA delivery: A Critical Assessment of Current Technology Prospects for Clinical Application Authors: Bailey M. Cooper & David Putnam

ACS Paragon Plus Environment

Page 50 of 50