Review pubs.acs.org/CR

Mechanisms of Drug Release in Nanotherapeutic Delivery Systems Pamela T. Wong and Seok Ki Choi* Michigan Nanotechnology Institute for Medicine and Biological Sciences, Department of Internal Medicine, University of Michigan, Ann Arbor, Michigan 48109, United States 3.1. Pore Size 3.2. Effective Volume 3.3. Cross-Links 3.4. Conformation 3.5. Microbubbles 4. Conclusions 4.1. Summary 4.1.1. Ester Hydrolysis 4.1.2. Amide Hydrolysis 4.1.3. Hydrazone Hydrolysis 4.1.4. Disulfide Exchange 4.1.5. Hypoxia Activation 4.1.6. Mannich Base 4.1.7. Self-Immolation 4.1.8. Photochemistry 4.1.9. Thermolysis 4.1.10. Azo Reduction 4.1.11. Encapsulation 4.2. Status of Nanotherapeutic Development 4.3. Challenges in Nanotherapeutic Linker Design 4.3.1. Synthetic Chemistry 4.3.2. Binding Avidity 4.3.3. Physicochemical Properties 4.3.4. Pharmacokinetics (PK) 4.3.5. Tissue Penetration Author Information Corresponding Author Notes Biographies Acknowledgments Abbreviations References

CONTENTS 1. Introduction 1.1. Nanotherapeutics 1.2. Nanocarriers 1.3. Biomarkers 1.4. Aims and Scope 2. Mechanisms of Drug Release via Linker Cleavage 2.1. Ester Hydrolysis 2.1.1. Biological Mechanism 2.1.2. Chemical Mechanism 2.1.3. Representative Mechanisms of Drug Release 2.2. Amide Hydrolysis 2.2.1. Biological Mechanism 2.2.2. Chemical Mechanism 2.2.3. Representative Mechanisms of Drug Release 2.3. Hydrazone Hydrolysis 2.3.1. Mechanism 2.3.2. Synthetic Methods 2.3.3. Release Kinetics 2.4. Disulfide Exchange 2.4.1. Chemical Mechanism. Endogenous Thiols 2.4.2. Design of Disulfide-Tethered Drug Molecules 2.5. Hypoxia Activation 2.5.1. Biological Mechanism 2.5.2. Drug Release via Bioreductive Mechanisms 2.6. Mannich Base 2.7. Self-Immolation 2.8. Photochemistry 2.9. Thermolysis 2.10. Azo Reduction 3. Mechanisms of Drug Release via Control of Carriers

© 2015 American Chemical Society

3388 3389 3389 3389 3390 3391 3391 3391 3393 3393 3398 3398 3399 3400 3402 3402 3402 3404 3405

3413 3414 3415 3415 3415 3416 3416 3416 3416 3416 3416 3416 3416 3416 3416 3416 3416 3416 3417 3417 3417 3417 3417 3417 3418 3418 3418 3418 3418 3418 3418 3419

1. INTRODUCTION Targeted drug delivery refers to an established therapeutic strategy that develops platforms and nanoscale devices for selective delivery of small drug molecules and therapeutic genes to cells of interest.1−10 Molecular strategies to develop such delivery systems vary to a large extent, but all utilize nanometersized entities or other forms of nanocarriers to deliver therapeutic payloads to their targeted cells. The concept of nanotherapeutic delivery relies on the combination of three key mechanistic elements, each thought to play an essential role for efficient delivery (Figure 1): (i) specific cellular binding, (ii) intracellular uptake of drug-carrying nanomaterials by targeted cells, and (iii) controlled release of carried drug molecules in an

3405 3405 3406 3406 3407 3408 3409 3410 3412 3412

Received: August 22, 2014 Published: April 27, 2015

3413 3388

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

anticancer therapeutics where almost all chemotherapeutic agents suffer from one or more of such issues. For example, doxorubicin (DOX), which is highly efficacious in treating various types of tumors, suffers from the drawback of serious cardiotoxicity11 due to the lack of tumor-specific cytotoxicity. Many other anticancer drugs such as paclitaxel (PTX), 10hydroxycamptothecin, 5-fluorouracil (5-FU), and tamoxifen present with the difficulties of making aqueous formulations due to the hydrophobic nature of these agents. Nanotechnology for drug delivery provides a diverse array of delivery platforms, and is uniquely suited for improving the safety and efficacy profiles of current chemotherapeutic agents. 1.2. Nanocarriers

Nanoscale entities can take various shapes such as globular particles, tubes, and rods, and can serve as versatile nanocarrier platforms that have the potential to also provide multifunctional properties. They have been actively investigated for the development of targeted nanotherapeutics and diagnostic devices for applications in cancer, as well as in inflammatory, infectious, and autoimmune diseases.3,4,12−15 Currently, there are many classes of such nanocarriers as summarized in Table 1.2 They include organic-based nanomaterials such as dendrimer nanoparticles (NPs),16−27 polymers and polymerbased micelles,28−38 carbon nanotubes,39−43 iron oxide NPs (IONPs),26,44,45 and gold NPs (AuNPs).46 Each of these nanomaterials is unique in its chemical and physical aspects such as synthetic methods, surface functionality and modification, core−shell architecture, size, and shape. Understanding the physicochemical properties of these materials and their interactions with biological systems is highly important in developing therapeutic applications, as some of these nanomaterials are known to cause undesired effects due to their intrinsic toxicity47,48 or immunogenicity49 often due to suboptimal surface modification.50 For example, cationic nanoparticles are cytotoxic due to their ability to disrupt cellular membranes as illustrated by unmodified PEI and PAMAM dendrimers.51−53 However, the cytotoxicity of these nanomaterials is reduced or fully eliminated by modification of the surface with neutral or anionic groups.54 This surface modification is also important for developing certain inorganic classes of nanoparticles which are known to be intrinsically toxic, such as cobalt−chromium nanoparticles and multiwalled nanotubes (MWNTs) that can damage DNA strands48 or suppress immune function,47 respectively. Various aspects of nanotoxicity related to therapeutic applications have been reviewed thoroughly in recent articles.55−63 Thus, the design and engineering of nanomaterials for drug delivery purposes necessitates a variety of approaches. In the main sections to follow, we will discuss selected methods for nanomaterial conjugation with drug molecules.

Figure 1. Schematic for cell targeted delivery of therapeutic agents carried by a nanoparticle (NP). The process is comprised of three sequential steps: (i) NP binding to target cells via multivalent receptor−ligand interactions, (ii) intracellular uptake of the NP via receptor-mediated endocytosis, and (iii) intracellular drug release or action.

active form. Thus, it is important that the third step involving the release of drug should occur in a precisely controlled manner in order for the drug to display its biological activity in the targeted cell only. Despite being a critical aspect in the design of nanotherapeutic delivery systems, the various mechanisms that have been developed for drug release have not yet been systematically reviewed from a molecular point of view. This review thoroughly surveys release mechanisms as classified on the basis of linker type for various drugs conjugated to nanocarriers, and discusses how such mechanisms are designed and incorporated specifically into the delivery platform to achieve specific targeted release in diseased cells. 1.1. Nanotherapeutics

Small molecule therapeutic agents often cause unwanted adverse events and systemic toxicity. There are many factors that play into these clinical issues; however, in most cases, the suboptimal activities and unwanted side effects observed for these drugs can be attributed to their pharmacokinetics and pharmacodynamics (PK/PD). PK/PD parameters that have hampered drug effectiveness include off-target activity that leads to a narrow therapeutic index (TI) (TI = lethal dose (LD50) ÷ effective dose (ED50)), low aqueous solubility due to drug hydrophobicity, rapid clearance and extensive metabolism of the drugs in vivo, and nonspecific tissue accumulation. Nanotherapeutic delivery strategies aim to overcome such problems and show great potential, in particular, in the area of

1.3. Biomarkers

Delivery of drug molecules by a nanocarrier is generally achieved through two approaches as summarized in Figure 2: (i) nontargeted delivery (passive targeting) and (ii) targeted delivery (active targeting). Passive targeting takes advantage of the increased permeability of the endothelial blood microvasculature in tumors where interstitial gaps between neighboring endothelial cells are larger (∼200−1200 nm)68 than in normal tissue. Thus, the enhanced permeability of these blood microvessels allows for easier extravasation of drugloaded nanocarriers into tumors, leading to greater accumulation and a longer duration of drug exposure in the tumor due 3389

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Table 1. Selected Nanometer-Sized Carriers Used for Drug Conjugation in Targeted Drug Deliverya class PAMAM16,17

PEI, PPI (linear, branched)19,23,24,26 poly(triazine)22,64 poly(lysine)25 poly(glycerol)18 poly(ester)65 PLGA, PEG−PLGA28−31 HPMA32 HA33,34 dextran35−38 SWNT39−42 MWNT43 IONP26,44,45 QD66,67 AuNP46

size (d, nm)

shape

Dendrimer Nanoparticle (NP) 3.6 (G3) S 4.5 (G4) 5.4 (G5) 6.7 (G6) ≤10 S ≤10−50 S − S − S − S Polymer and Polymeric Micelle ≤200 S − S, L − S, L − S, L Carbon Nanotube (NT) 1−2b T 2−25 T Inorganic Nanoparticle (NP) 5−50 S 1−10 S 8−70 S

surface functionality

conjugation chemistry

NH2, CO2H, CO2Me

Am, Es, thiol−alkene

NH2 NH2, CO2H NH2 OH CO2R

Am Am, Es Am Es, Et Am, Es

CO2H, OH OH CO2H, OH CO2H, OH

Am, Es Es Am, Es

CO2H, OH CO2H, OH

Am Am

NH2 CO2H Au

Am Am Au−S

a Gn = nth generation; S = spherical; L = linear; T = tubular; d = diameter; Am = amide; Es = ester; Et = ether; PEI = poly(ethylene imine); PPI = poly(propylene imine); PLGA = poly[(D/L)-lactic acid-co-glycolic acid]; PEG = poly(ethylene glycol); HPMA = 2-hydroxypropyl methacrylate; HA = hyaluronic acid; SWNT = single-walled carbon nanotube; MWNT = multiwalled carbon nanotube; IONP = iron oxide (Fe3O4) nanoparticle; QD = quantum dots; AuNP = gold nanoparticle. bLength = 200−500 nm.

to limited elimination.69−71 Even without targeting ligands on the nanocarrier surface, such nanotherapeutic agents are passively recruited to the tumors by this enhanced permeation and retention (EPR) effect.69−71 On the other hand, active targeting aims to both take advantage of the EPR effect and achieve specific cancer cell targeting. The concept for this targeted approach relies on modification of the nanocarrier surface through covalent attachment of targeting ligands specific for a cell surface biomarker or receptor molecule that is overexpressed in the cancerous cells (Figure 1). As summarized in Table 2, cancer-associated biomarker molecules include the following: families of vitamin receptors such as the folic acid (FA, vitamin B9) receptors (FAR-α, FARβ),72−74 riboflavin (vitamin B2) receptor75,76 and biotin receptor; the αvβ3 integrin receptor;77−79 prostate-specific membrane antigen (PSMA) receptor;80 a group of growth factor receptors including Her2,81 epidermal growth factor receptor (EGFR)81−83 and fibroblast growth factor receptor (FGFR);84 insulin and insulin-like receptors;85 a family of selectin protein molecules;86−88 and transferrin receptor (TfR89). In conjugating the ligand to the nanocarrier, the targeting ligand is attached typically in multiple copies to the nanocarrier surface, as the mechanism for endocytic uptake of such targeted nanocarriers requires multiple simultaneous interactions to occur collectively at the interface of multiple surface receptor−ligand pairs.77,90 This multivalent binding mechanism91 confers tight adhesion of the nanocarrier to the targeted cell surface, and is considered to be highly important during receptor-mediated uptake (endocytosis) of nanocarriers by the cell. Without it, the nanocarriers are bound weakly and have too short of a residence time to get taken up through the formation of coated pits.91−94 Therefore, each nanocarrier is covalently conjugated with multiple copies of a targeting ligand

on its periphery in order to maximize these multivalent effects.12,95−97 1.4. Aims and Scope

The scope of this article is primarily focused on nanodelivery systems that carry therapeutic molecules attached through covalent linkers (“conjugated”). Other delivery systems that carry unmodified drug molecules through complexation or encapsulation (such as within polymeric liposomes) are already extensively reviewed9,137−142 and are discussed only to a limited extent at the end of this review. Here, we focus on the rational design of the linker as it plays a primary role in not only covalent attachment for carrying the drug molecule but also in providing a mechanism for controlled drug release. The therapeutic agents to be discussed are summarized and categorized according to their therapeutic areas, ranging from anticancer to antimicrobial (viral, bacterial and fungal) to immunomodulatory agents (Table 3). However, greater focus will be placed on anticancer therapeutics because more applications have been investigated in this particular area. The conditions that control drug release by triggering linker cleavage involve pathophysiological features and subcellular properties specific to diseased cells. Select triggering mechanisms to be discussed include tumor hypoxia (low oxygen levels due to increased metabolic rates in tumor cells), low intracellular pH (endosomes and lysosomes where targeted nanomaterials are taken up),143 lowered extracellular pH for tumor cells,144 tumor-specific enzymes (matrix metalloproteinase, prostate-specific membrane antigen) overexpressed on the cell membrane,144 and upregulation of glutathione.145 Such mechanisms of drug release will be discussed according to the structure and function of the linker type. Linker type chemistries to be discussed include ester, amide/peptide, disulfide, hydrazone, hypoxia-activated, and self-immolative 3390

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

linkages (Table 4). Finally, an emerging strategy that uses photochemistry or thermolysis for triggering drug release in an actively controlled manner will be discussed for its mechanism and applications. In summary, we aim to provide comprehensive coverage of the various aspects important to linker design and of the specific molecular mechanisms that are employed to achieve controlled release of therapeutic agents only under specific conditions. We believe this review article will serve as an invaluable source of information for understanding and developing innovative controlled drug release systems.

2. MECHANISMS OF DRUG RELEASE VIA LINKER CLEAVAGE 2.1. Ester Hydrolysis

2.1.1. Biological Mechanism. The use of carboxylic ester based linkers in targeted release applications is primarily influenced by their convenience and by the wide applicability of ester linker chemistry to many therapeutic molecules. As summarized in Table 3, having a functional group such as a carboxylic acid or alcohol allows the use of a conjugation strategy using ester-based drug attachment to the nanocarrier. The resulting ester bond subsequently opens a route for drug release due to its general susceptibility to hydrolysis when exposed to physiological conditions in vivo. The ester linker is readily cleaved by the hydrolytic reactions catalyzed by acids,182,183 bases,183,184 metal ions,183,185 and hydrolytic proteins such as human serum albumin186 and esterases.187,188 Despite its contribution to drug release, such physiological instability by itself is not effective and specific enough to trigger drug release in a controlled manner. In particular, metal ion catalyzed hydrolysis of ester bonds has been reported with certain metal ions such as Cu2+,183,185 but significant rate acceleration has only been observed in model substrates that are designed such that metal ions are chelated proximal to the carboxylate functional group.185,189,190 Additional mechanisms

Figure 2. Schematic illustration for tumor-directed delivery of drugcarrying nanoparticles (NP) via two mechanisms: (i) an enhanced permeation and retention (EPR) effect in tumors (passive targeting); (ii) the receptor-specific binding and uptake by a tumor cell (active targeting).

Table 2. List of Surface Biomarkers (Cell Surface Receptors, Antigens) Overexpressed in Cancers and Inflammatory Diseasesa biomarker receptor

cellular function

MTX (KD = ∼20−100 nM),33,99−101 quinazoline102 MTX quinacrine (KD = 264 nM)76

B, O Mc B, P

− peptide (GE11111) peptide113

O B, HN B

FGF (KD = 0.3 nM)115 Arg-Gly-Asp120 (KD = 0.2 μM)121

− cilengitide,116 c(RGDyK),119 RGD4C, ATN-161117

B, L, M B, M

peptide metabolism

N-Ac-Asp-Glu

P

glucose metabolism iron uptake

insulin (KD = 1.2 nM)131 transferrin (KD = 35 nM)134

urea-based (Lys-CO-Glu),80 aptamers,31 phosphoramidate124 benzoquinones (EC50 = 300 nM)126,127 −

B, P, O B, Br

cell adhesion cell adhesion blood coagulation

sLex, sLea epitope lectins factor VIIA

carbohydrate, glycopeptide mimics − −

P I, L HN

FA uptake FA uptake riboflavin uptake105

biotin receptor106−108 EGFR (Her1)82,109 Her281,112

biotin uptake cell growth cell growth, differentiation tissue homeostasis cell adhesion

insulin receptor126−130 transferrin receptor89,123,132,133 selectins (E, L, P)69−71,120 Ley135 tissue factor136

related tissues

synthetic ligand

FA (vitamin B9; KD ≈ 0.4 nM)99 FA riboflavin (vitamin B2; KD = 5 nM)76 biotin (Km = 4−22 μM)106 EGF (KD = 1−10 nM),110 TGF-α unknown

FAR-α12,72,90,97,98 FAR-β103,104 riboflavin receptor58,59,88−90

FGFR84,114 integrin receptor (αvβ3)78,116−119 PSMA (GCPII)31,80,122−125

endogenous ligand

a

FA = folic acid; FAR = folic acid receptor; MTX = methotrexate; EGF(R): epidermal growth factor (receptor); EGF = epidermal growth factor; FGF(R) = fibroblast growth factor (receptor); PSMA = prostate specific membrane antigen; B = breast; Br = brain; HN = head and neck; I = intestine; L = lung; M = melanoma; Mc = macrophage activated; O = ovary; P = prostate. 3391

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Table 3. List of Therapeutic Agents and Conjugation Chemistry Developed for Drug Release Mechanismsa drug name auristatin E (MMAE) bleomycin camptothecin calicheamicin cisplatin colchicine doxorubicin methotrexate mitomycin C maytansine paclitaxel α-TOS thapsigargin tamoxifen vancomycin erythromycin amphotericin B saquinavir ribavirin adenosine arabinoside amprenavir SDC-1721 cyclosporine FK506

mode of action

linkage functionality

Anticancer Therapeutics inhibition of tubulin polymerization NH-Me (Val), OH DNA fragmentation NH2 DNA topoisomerase I C20-OH DNA strand cleavage trisulfide DNA base cross-linking CO2H, NH2 inhibition of tubulin polymerization HNC(O)R DNA intercalation CO (C13), NH2 (daunosamine) DHFR, thymidylate synthase CO2H (Glu) DNA cross-linking carbamate, NH (aziridine) inhibition of tubulin polymerization N-Me (Ala) microtubule stabilization C2′-OH, C7-OH, C10 apoptosis succinate inhibition of Ca2+-ATPase C8-OH estrogen receptor NMe2 Antimicrobial Therapeutics inhibition of bacterial cell wall biosynthesis CO2H, NH2 inhibition of bacterial rRNA complex OH pore formation (fungal) CO2H, NH2 protease inhibitor (HIV) C26-OH inhibition of RNA metabolism (HCV) C5′-OH inhibition of viral DNA synthesis C5′-OH protease inhibitor (HIV) NH2 (Ph) fusion inhibitor (HIV) SH Immunomodulatory Therapeutics inhibitory regulation of calcineurin OH inhibitory regulation of calcineurin OH (C32)

conjugation chemistry Am,87 Hz146 Am,147 R174 Am,148 E149 Am,135 Hz150 Es,40,151 Hz152 Am153 Am,154,155 Hz,156 Mb,116 Cb157 Am,43,158,159 Es11,79,157−161 SS162 SS163,164 Am,165 Es,166 SS167 Es, Am168,169 Es32 Au−S170 Am171−174 Es175 Am42 Es176 P177 P178 Am179 S-Au180 Es181 Es38

α-TOS = α-tocopheryl succinate; Am = amide; Cb = carbamate; SS = disulfide; Es = ester; Hz = hydrazone; Mb = Mannich base; P = phosphoramidate; R = reductive amination. a

Table 4. Summary of Release Mechanisms by Linker Conjugation Chemistry linker ester amide hydrazone indolequinone Mannich base self-immolative (aminobenzyl) o-nitrobenzyl, coumarinyl diazo

release mechanism

therapeutic agents

enzymatic (hydrolases, alkaline phosphatase, carboxylesterase); chemical hydrolysis (low or high pH, metal cations) enzymatic (peptidases, matrix metalloproteinase, collagenase, prostate specific membrane antigen, plasmin) hydrolysis (pH 5.5−6.5) hypoxia (low O2) hydrolysis (acid) enzymatic, bioreduction, hydrolysis, thiol exchange

taxol, methotrexate, cisplatin, SN38, FK506, cyclosporine A, erythromycin methotrexate, doxorubicin, SN38, auristatin, α-TOS

light (UV, vis)

doxorubicin, methotrexate, 5-FU, taxol, tamoxifen, chlorambucil doxorubicin, aminosalicylic acid, 9aminocamptothecin

thermolysis, enzymatic reduction

doxorubicin, taxol naloxone, cisplatin, mustard agent doxorubicin doxorubicin, taxol, calicheamicin, naloxone

environment created by endosomes (pH 5.0−6.0)194 and lysosomes (pH 4.8)192,193 such that drug release occurs more rapidly and selectively after uptake than in other neutral environments such as the plasma and cytoplasm (pH 7.4) (Table 5). Thus, the primary mechanism of drug release from ester-based drug linkers on nanomaterials is attributable to specific uptake by a target cell and intracellular linker hydrolysis by the action of acid hydrolases occupying the acidic compartments. In addition to the intracellular activity of hydrolytic enzymes involved in controlled drug release, there are other diseaseassociated hydrolytic enzymes which are overexpressed in the extracellular membrane or plasma (soluble). Because of the upregulation in their expression due to certain abnormalities in diseased cells under pathological conditions, the activity of

that make a greater contribution to the controlled release relate to cellular events that occur during the uptake and processing of the drug-carrying nanomaterials by a target cell. As illustrated in Figure 1, the path proposed for nanomaterial uptake occurs primarily via receptor-mediated endocytosis in which nanomaterials are taken up into the cytosol through coated pits and end up in the compartment of an early endosome that is then sorted to late endosomes before being fused with lysosomes.191 The interiors of such cellular vesicular compartments contain numerous acid hydrolases192,193 including cholesteryl ester acid hydrolase, aryl sulfatase, acid phosphatase, N-acetylglucosaminidase, and cathepsin D.188 These hydrolases are typically involved in the ester or amide hydrolysis of a diverse set of substrates including ester-linked drug conjugates.183 Such enzymes display more optimal catalytic activity in the acidic 3392

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

>8.0

peroxisome (8.2)198 plasma/extracellular membrane (7.4);197 cytoplasm (7.3);192 endoplasmic reticulum (7.1);192 mitochondrial matrix (7.9−8.0);195 nucleus (7.7)192

7.0−8.0

these enzymes can be exploited for the release of drugs from ester-linked NP−drug conjugates.187,188,199,200 Alkaline phosphatasea Zn2+/Mg2+-dependent membrane-bound metalloenzymehas been shown to be able to release a quinolinone drug molecule by catalyzing the hydrolysis of a drug−Ophosphate bond.201 Carboxylesterase 2 (CE-2) is a serinedependent esterase localized on the membrane, and it constitutes one of the biomarker proteins expressed abundantly in liver and colon tumors. This esterase catalyzes the hydrolysis of endogenous lipid esters and is involved in drug release such as in the cleavage of carbamate-based prodrug substrates including the anticancer agent irinotecan (CPT-11202). This enzyme cleaves the carbamate-based linker moiety of irinotecan, releasing SN38 as an active metabolite which serves as a potent inhibitor of topoisomerase I.202,203 Carboxylesterase 1 (CE-1), an isozyme of CE-2, also cleaves the carbamate linker and is involved in the first step of the metabolic pathway of capecitabine that leads to the release of 5-fluorouracil (5-FU). Such esterase-triggered drug release could also be achieved through the activity of other esterases including acetylcholine esterase (AChE), butyrylcholine esterase (BuChE), and paraoxonase 1 (PON1). However, each of these enzymes has not been clearly validated as a tumor-specific or other diseasespecific biomarker. Nevertheless, their utility as enzymatic tools for the control of drug release is illustrated in the hydrolytic activation of butyryl acyclovir (AChE),204 irinotecan (BuChE), and prulifloxacin (PON1205). 2.1.2. Chemical Mechanism. The ester bond used for drug conjugation is cleavable via chemical mechanisms such as hydrolytic reactions catalyzed by acids,182,183 bases,183,184 and metal ions. 183,185 Despite the general perception that compounds would be more stable under neutral conditions (pH 7.4), most esters, in fact, show greater stability in slightly acidic media (pH 5−6) because ester hydrolysis via acid or base catalysis has minimal impact at this pH range.183 Therefore, chemical hydrolysis of ester-linked drug molecules carried by nanoconjugates may play only a limited role in facilitating drug release even after uptake into acidic endosomes and lysosomes, or upon exposure to the extracellular matrix (ECM) of a tumor which is more acidic (pH 6.2−6.9) than the plasma (Table 5). However, the ester linker is more susceptible to base-catalyzed hydrolysis, which opens up the possibility for controlled release under conditions that target basic subcellular compartments such as mitochondrial matrixes (7.9−8.0)195 and peroxisomes (8.2).198 For example, the ester linkers in the ester conjugates of paclitaxel are hydrolyzed at a rate 10-fold faster with each increase in pH unit, such as observed for a change from pH 7 to 8.182 In summary, ester linkages show stability in mildly acidic conditions such that acid-catalyzed hydrolysis alone plays only a minor role in controlled drug release of these nanoconjugates. 2.1.3. Representative Mechanisms of Drug Release. 2.1.3.1. Release Mechanism of Taxol. The taxane class of anticancer agents includes paclitaxel and docetaxel (taxotere), and are classified as microtubule-stabilizing agents due to their ability to bind and stabilize cytosolic microtubules.206 Microtubule stabilization results in cell growth inhibition by interfering with the dynamic processes of cellular division and mobility which depend largely on the ability of microtubules to assemble and disassemble. Each microtubule is a hollow cylindrical polymer made up of repeating units of a heterodimer of α-tubulin and β-tubulin. The taxol molecule binds selectively to the β-tubulin unit on the luminal side207 of the tubular structure (outer diameter ≈ 25 nm; inner diameter ≈ 12

Golgi apparatus (6.6);195 extracellular matrix (tumor: 6.2−6.9)196

6.0−7.0 100-fold higher activity as compared to the free platinum(IV) prodrug agent in testicular cancer cells. Such enhanced activity is attributable to more efficient endocytic uptake and facilitated activation of the platinum(IV) prodrug carried by the SWNT carrier. 2.1.3.4. Release Mechanism of SN38. SN38 is an active metabolite of irinotecan. Despite its potent antitumor activity, it is poorly soluble in water and would thus benefit greatly from better targeted methods of delivery. Ghandehari et al.149 used carboxylic acid terminated PAMAM (G3.5) as a carrier for SN38 which was attached at the C20-OH position via an ester linker with a glycine or a β-alanine spacer (Figure 7). While

Figure 8. Structures of immunosuppressive agent (FK506, cyclosporine) conjugated macromolecules in which each drug is carried by a carboxylic acid terminated dextran or poly(lactide) and conjugated through an ester bond.

is 2 orders of magnitude more potent than cyclosporine in vitro, though FK506 is rapidly cleared in plasma following intravenous administration and is extensively metabolized by the liver cytochrome enzyme P450 3A. Hashida et al.38 developed a dextran-based delivery system for FK506 in order to control its pharmacokinetics and enhance its therapeutic efficacy. The drug molecule was conjugated through an ester linkage (Figure 8). Release studies of FK506 from its dextran conjugate indicated that the ester linker was fairly stable, and accordingly, the rate of drug release was slow in a phosphate buffer (pH 7.4) with a half-life of ∼150 h. Due to such a slow rate of release, the FK506−dextran conjugate showed an extended circulation time in blood with a clearance rate ∼1800fold lower than that of free FK506. Cyclosporine A (CsA) suffers from a lack of selectivity and a narrow therapeutic index, both of which cause major adverse events such as nephrotoxicity upon administering chronic doses. Abdi et al.181 employed a targeted drug delivery system to carry and release cyclosporine to the diseased organ in a sustained manner. They prepared a cyclosporine-tethered poly(lactide) (Figure 8), and coprecipitated it with PEG-linked poly(lactide) to prepare polymeric nanoparticles (∼100 nm).

Figure 7. Structure of an SN38 anticancer molecule carried by a carboxylic acid terminated PAMAM dendrimer (G3.5) where the drug is conjugated through a glycine or β-alanine ester at the C20 position.

these two spacers differ only by a methylene group, such differences in linkers could impact the rate of drug release and the activity of the conjugates. Both conjugates were stable in PBS with less than 5% of the drug being released after 24 h. When incubated in a rat plasma (50%) medium, release of SN38 was higher from the glycine (∼12%) than the β-alanine (∼8%) spacer. This difference suggests involvement of enzymes in the drug release since esterases are present in the plasma. Both conjugates were active in inhibiting proliferation of human colorectal cancer cells. However, the conjugate made with the glycine spacer (IC50 = 129 nM) was ∼3-fold more 3397

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

ester linkage used for drug conjugation was quite susceptible to hydrolysis even in PBS buffer (pH 7.4), with about 90% of the drug being released within 10 h. The authors noted that the ester bond in the 2′ position was optimal for effective drug release as compared to the other ester bond located proximal to the dendrimer periphery. Such ester instability to hydrolysis is remarkable compared to other examples reviewed earlier, but is likely attributable to the involvement of a neighboring tertiary amine group that acts as an autocatalyst for the hydrolysis of the ester link.

They showed that this delivery system released CsA via cleavage of its ester linker in a sustained manner, and resulted in suppression of T-cell proliferation as potently as free CsA. It is notable that the CsA-linked poly(lactide) nanoparticles were internalized in dendritic cells and that these nanoparticleloaded cells migrated to lymph nodes where CsA was released in a sustained manner. 2.1.3.6. Release Mechanism of Erythromycin. As a member of the macrolide class of antibacterial molecules, erythromycin potently inhibits bacterial growth by blocking a rRNA complex involved in protein biosynthesis. It also has additional activities related to host cells including an anti-inflammatory effect and gastrointestinal pro-motility effect. Kannan et al.175 explored a dendrimer-based conjugate system for achieving sustained release of erythromycin. In the design, erythromycin was attached to a hydroxyl-terminated G4 PAMAM dendrimer through an ester bond by performing a regioselective reaction at its 2′-OH position with the glutarate spacer (Figure 9). The

2.2. Amide Hydrolysis

2.2.1. Biological Mechanism. Drug conjugation to the nanocarrier through an amide link is widely applicable for many therapeutic molecules because each drug molecule is commonly functionalized with a carboxylic acid or amine group. The resulting amide bond is much more stable and less susceptible to chemical hydrolysis than the corresponding ester bond. Unless designed with a special functional group, the amide linker is rarely cleaved by chemical hydrolysis under physiological conditions, and its chemical cleavage requires harsh conditions such as much higher temperatures in combination with the presence of strong acid or base catalysts.183 Such general stability can provide certain benefits for achieving improved pharmacokinetic profiles by extending the duration of circulation in the blood. However, its use for drug release requires more consideration including other nonchemical methods of amide cleavage. Generally, these methods are based on enzymatic mechanisms and are carried out by hydrolytic proteases 183 such as serine proteases,32,155,229−232 cysteine proteases,148 and zinc-dependent endopeptidases (Table 7).144,158,233 Each of these enzymes is localized in one or more sites in the cell ranging from the extracellular environment, to the cellular membrane, to intracellular lysosomes, and thus their site-specific action is directly related to the site of drug release. As an illustration, matrix metalloproteinases (MMPs) such as collagenases are zinc-dependent endopeptidases secreted primarily into the ECM environment.235 MMPs are normally involved in the remodeling of the extracellular matrixes through their degradation of ECM proteins. Several MMPs are overexpressed in a variety of tumor cells and are implicated in the dysregulation of angiogenesis leading to tumor growth and metastasis.236 Thus, use of an MMP-specific peptide as a cleavable linker serves as a strategy for controlling drug release at the tumor ECM. This is illustrated by MMP-mediated release of methotrexate linked to a poly(lysine) dendrimer through a peptide linker containing the MMP specific cleavage sequence, PVG↓LIG (Table 7).158 MTX release upon incubation with MMP 2 or MMP 9 was demonstrated in a time-dependent manner with ≤20% efficiency at 24 h.

Figure 9. Structure of a fourth generation (G4) PAMAM dendrimer conjugated with erythromycin, each drug attached through an ester at a D-desosamine sugar residue (C2′-OH) to a carboxylic acid terminated spacer to hydroxylated G4 dendrimer. Below is shown a proposed mechanism of drug release at neutral pH (7.4) that occurs via C2′ ester hydrolysis catalyzed by a neighboring tertiary amine at the C3′ position.

Table 7. List of Proteases Catalyzing Amide Hydrolysis for Anticancer Drug Releasea peptidase 144

MMP collagenase (MMP 1, 8, 13)158,233 PSA32,229,230 PSMA (GCPII)122,234 plasmin155,231,232 a

enzyme classification EC EC EC EC EC

3.4.24.x 3.4.24.7 (34) 3.4.21.77 3.4.17.21 3.4.21.7

enzyme function

localization

substrate sequence

zinc endopeptidase zinc endopeptidase serine protease zinc peptidase serine protease

extracellular; matrix, cell surface; lysosome extracellular; matrix, cell surface, lysosome extracellular matrix, serum, lysosome cell surface blood, lysosome

PQG↓LAG; PVG↓LIG LS↓N; HSSKLQ↓L; SSKYQ↓L Ac-D↓E; folate↓E (Y/F)(R/K)↓SR AFK↓

MMP = matrix metalloproteinase; ↓ = cleavage site; PSA = prostate-specific antigen; PSMA = prostate-specific membrane antigen. 3398

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

DE). Incorporation of such substrates into a linker framework opens a potential approach for targeted drug release in prostate tumors. Plasmin is another example of a protease that is utilizable for the release of a drug molecule linked through a plasmin-susceptible peptide bond.155,231,232 It is also a serine protease that is localized in the plasma and in lysosomes. Therefore, this enzyme provides a method for controlled drug release both during circulation in blood and after intracellular uptake of drug conjugates in the lysosomes. 2.2.2. Chemical Mechanism. Cleavage of amide linkers by chemical hydrolysis does not constitute a general mechanism for drug release; however, certain classes of specialized linkers attached to drug molecules through an amide linker have been shown to be responsive to low pH environments. These linkers are composed of a group of maleic acid based frameworks such as citraconyl, cis-aconityl, and maleyl groups (Figure 11). Their

The MMP-cleavable peptide linker has been used for fluorescence imaging of primary tumors and metastases in vivo that overexpress MMPs through the use of ratiometric activatable cell-penetrating peptides (RACPPs).237−240 This imaging technology employs a protease-specific substrate peptide as a spacer that tethers two cell penetrating peptides, each terminated with a Cy5 and Cy7 reporter dye molecule, respectively (Figure 10).238,241,242 In noncancerous cells where

Figure 10. Concept of ratiometric activatable cell-penetrating peptide (RACPP). The protease-cleavable linker located in the middle of this RACPP is cleaved by specific proteases overexpressed in a pathologic cell, which leads to abolishment of intramolecular FRET emission (Cy5 to Cy7) and instead to emission of Cy5 alone.

MMP expression is relatively low, the cleavage of this RACPP is slow and the two tethered dye molecules undergo fluorescence resonance energy transfer (FRET) (λex,Cy5 = 620 nm; λem,Cy7 = 780 nm) due to their spatial proximity. However, in tumor cells that overexpress MMP 2 and MMP 9, the two dye molecules are separated upon cleavage of the MMP-specific peptide spacer, and FRET no longer occurs, resulting in emission at only the Cy5 wavelength (λex,Cy5 = 620 nm; λem,Cy5 = 670 nm). Application of this imaging technology has been broadly demonstrated for in vivo detection methods and for monitoring certain proteases (MMPs,237−239,243 thrombin242,244) in cancer and other therapeutic areas including asthma.243 Prostate-specific antigen (PSA) is another enzyme target that is applicable for controlled drug release. It belongs to the family of serine proteases and is expressed at high levels in prostate tumor tissue. The sequences of the peptide substrates most susceptible to PSA are comprised of serine (S)-glutamine (Q) and glutamine (Q)-leucine (L).32,229,230 Doxorubicin conjugated to a peptide spacer containing the SQ dipeptide sequence at its aminoglycoside was readily released via peptide cleavage by PSA.229,230 Prostate-specific membrane antigen (PSMA) constitutes the other enzyme biomarker highly expressed in prostate cancer cells.107,210 It is a zinc endopeptidase that acts as a glutamate carboxypeptidase (GCPII) on small molecule substrates including folate-γpolyglutamate (FA-En), methotrexate-γ-polyglutamate (MTXEn), and the neuropeptide N-acetyl-L-aspartyl-L-glutamate (Ac-

Figure 11. Structure of pH-susceptible amide linkers (top) and illustration of a proposed acid-catalyzed cleavage mechanism for a cisaconityl amide linker, leading to doxorubicin release (bottom).

utility is illustrated by amide conjugation of the cis-aconityl group to doxorubicin at its daunosamine sugar.245 When incubated in citrate−phosphate buffers at various pHs and 37 °C, this amide linker was hydrolyzed to release free doxorubicin with half-lives at pH 4, 5, and 6 of aconityl > maleyl amide).248 2.2.3. Representative Mechanisms of Drug Release. 2.2.3.1. Methotrexate. Several strategies have been reported for cancer targeted delivery of methotrexate conjugated through a peptide-based amide linker.33,35,158,159,249,262 In each of these approaches, drug conjugation is achieved through a peptidasesusceptible amide spacer, and drug release relies on a mechanism controlled by peptidases overexpressed on the cancer cell surface. Langer et al. designed a dextran-based delivery system for MTX which is conjugated through a peptide linker based on PVG↓LIG, the sequence susceptible to cleavage by matrix metalloproteinases (MMPs) (Figure 12, entry 1).35,249,250 Incubation of this conjugate with MMP 2 led to drug release with efficiencies of 67% (1 h) and 89% (24 h).250 The cleavage site of the peptide spacer (Figure 12) was determined by mass spectrometric detection of MTX-PVG after MMP treatment. Drug release efficiency was sensitive to the MMP subtype and the type of peptide spacer. Treatment with MMP 9 resulted in a lower level of release of 61% (24 h), and MTX molecules conjugated to the dextran polymer directly or through a random peptide sequence were not released. These results provide evidence supportive of MMP-controlled MTX release, and suggest that drug conjugation through a MMP-susceptible peptide spacer will be applicable for other types of anticancer agents as well. Porter et al. employed this MMP-specific peptide spacer for FAR-targeted MTX delivery using PEG−poly(lysine)-grafted 3400

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

more favored for MTX molecules linked through a longer spacer.268 2.2.3.2. Doxorubicin. Delivery systems for doxorubicin attached through an amide or peptide-based linker are illustrated in Figure 14. Each of these conjugates is made

the MTX conjugate made through a peptide with the sequence NFF was active in inhibiting growth of synovial fibroblast cells. Collectively, these studies demonstrate a correlation between MTX release and the sensitivity of the peptide linker to different MMP subtypes. This linker design based on such peptide linkage can play a critical role in disease-targeted drug delivery. For example, certain subtypes of MMP 2 and MMP 9 are implicated in promoting cancer metastasis in ovarian and breast cancers, while other subtypes like MMP 1 and MMP 3 are more involved in the progression of rheumatoid arthritis.265−267 In the delivery approach discussed above, conjugated MTX serves as a cytotoxic agent upon release. However, MTX can also play a role as a cancer targeting ligand because it belongs to the antifolate class of drugs, and its high structural homology to FA confers a reasonable affinity to FAR (KD = ∼20−100 nM).99−101 Recent studies reported by Baker et al.268 and Thomas et al.98,252 explored this delivery concept based on a dual-acting MTX in which conjugation of MTX alone to a nanocarrier without FA led to both FAR targeting activity and cytotoxicity. This dual-action approach is illustrated by G5MTXn conjugates (Figure 12; entries 5, 6) in which MTX is not releasable due to its conjugation through a peptidase-insensitive linker. One of the stable linkers used for testing this strategy was cyclooctyne-based amide, which was prepared by conjugation chemistry based on the copper-free click reaction between MTX−linker−azide and cyclooctyne-presenting dendrimer (Figure 13).251 As designed, this G5-MTXn conjugate

Figure 13. Synthesis of PAMAM dendrimer conjugated with methotrexate (MTX) attached through an amide linkage at its γcarboxylic acid position. The conjugation is based on copper-free click reaction of the dendrimer presenting cycloalkyne groups with γ-azidoMTX.

Figure 14. Nanocarriers conjugated with doxorubicin (DOX), each attached through an amide linkage at its daunosamine sugar residue.

was stable in serum, was selectively taken up by FAR-positive cells, and proved to be cytotoxic.98 The other spacer, which was based on repeating units of oligo(ethylene glycol), was explored for its applicability in large scale synthesis and as a tool for determining the relationship between the linker length and subsequent cytotoxicity (Figure 12, entry 6).252 This study suggests that the cytotoxicity of the G5-MTXn conjugate is significantly affected by the linker length and that MTX attached through an elongated spacer is more potent. This structure−activity correlation is consistent with the theoretical model in which inhibition of the DHFR enzyme is sterically

through attachment of the daunosamine sugar residue to the linker to be cleaved by peptidase action or low pH. Jones et al.229 demonstrated release of doxorubicin attached to a peptide containing the LSQ sequence (Figure 14, entry 1) which is designed to be hydrolyzed at the S↓Q site by prostate-specific antigen (PSA) (↓ denotes a bond to be cleaved). Drug release yielded both free doxorubicin and doxorubicin−leucine when exposed to prostate tumor cells that secrete prostate-specific antigen. As a result, this peptide-conjugated doxorubicin was more cytotoxic to PSA-secreting cancer cells than those normal 3401

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

for cancer cell specific drug release if conjugated through an active targeting system. As introduced earlier in section 2.2.2, the controlled release of an amide linked molecule can occur through a nonenzymatic reaction mechanism as illustrated by the maleimide-linked doxorubicin conjugates (Figure 14, entries 3, 4).245,258,259 Each of these amide linkers is unstable at pH 5 or 6; Table 5) but can occur faster after entering a tumor cell via endocytosis and exposure to the acidic environments of endosomes and lysosomes (pH 6.5−5.5). 2.3.2. Synthetic Methods. Hydrazone-based linkers refer to a class of linkers terminated with acyl hydrazone,131,135,238−240,242,243,278−281 alkoxy carbonyl hydrazone,152,260,270,271,282 and benzenesulfonyl hydrazone.25 Two synthetic approaches have been developed for hydrazone conjugation of a carbonyl group (ketone, aldehyde)-containing drug molecule to nanocarriers as summarized in Figure 17. First, a hydrazine-terminated linker is integrated onto the nanocarrier prior to the conjugation reaction with the drug molecule (method 1).152,156,270,274,277,279,280 This approach is illustrated by doxorubicin conjugation through its ketone.156,270,272,274,279,280 This method is preferred for certain types of drug molecules that have structures which contain nucleophilic functional groups such as amines as found in the daunosamine aminoglycoside of doxorubicin. If the nanocarrier is terminated with a nonhydrazine functional group such as a carboxylic acid instead, the amine from the drug will act competitively in the conjugation reaction with the nanocarrier, resulting in an amide linkage. However, the amide linker is chemically too stable and the high fraction of drug molecules linked through an amide linkage is subsequently not releasable even upon exposure to acidic subcellular compartments. Second, a hydrazone linker is installed first on the drug molecule by reaction with a hydrazine linker and the resulting drug linker is coupled to the carrier by a reaction chemoselective to the linker (method 2).25,282,283 This method is illustrated by paclitaxel conjugation in which a bifunctional linker comprised of an acyl hydrazine and maleimido group is used.276 The latter reacts selectively with the thiol presented on

Figure 15. Plasmin-catalyzed peptide cleavage that triggers doxorubicin release in a self-immolation process.

2.2.3.3. α-TOS. α-Tocopheryl succinate (α-TOS) is a vitamin E analogue that displays a potent apoptotic activity and has potential as an adjuvant for cancer immunotherapy.269 The structure of α-TOS consists of a chromanol substituted with a long hydrophobic tail which makes it poorly soluble in water. This physicochemical property reduces its biocompatibility and limits its application in cancer chemotherapy. In order to address this problem, Shi et al. employed FA-conjugated G5 PAMAM dendrimers as a nanoplatform for FAR-targeted delivery of α-TOS.168,169 In their approach, α-TOS was conjugated to the dendrimer through a succinyl amide bond at a density of 5−10 α-TOS molecules per dendrimer NP. The drug-conjugated dendrimers were water-soluble and further engineered to carry AuNPs partly encapsulated as a probe for imaging studies. In vitro and in vivo studies suggest that these drug-conjugated dendrimers are taken up by FAR(+) cancer cells and cause apoptosis as the result of α-TOS release. The release mechanisms are believed to occur by cleavage at one or two linkage points that include the ester bond to α-tocopherol and the amide bond to the dendrimer. Other amide linkers used for controlled drug release include peptides such as oligo(glycine)148 for camptothecin (SN38) and the valine−citruline peptide87,146 for auristatin. Each linker is cleavable by less selective proteases such as cathepsin B which is abundant in lysosomes (Table 8). However, drug conjugation through these peptide linkers can also open a route 3402

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Table 8. Amide Cleavage for Drug Releasea targeting ligand; nanocarrier

target receptor

drug delivered

linker chemistry

release mechanism

MTX; carboxymethyl dextran35,249,250 PEG−poly(Lys) G4 dendrimer158,159 MWNT43 HA33 PAMAM dendrimer (G5)98,251 PAMAM dendrimer (G5)252 PAMAM dendrimer (G5)109 albumin253 poly(L-lysine)254 antibody255,256 HPMA copolymer154 peptide scaffold229 RGD-4C155 PLGA257 pullulan258 chitosan245,259 PEG liposome260 dextran148

FAR

MTX

γ-amide; PVG↓LIG

MMP

FAR

MTX

MMP

FAR FAR FAR FAR cetuximab-EGFR lactose receptor FAR mammary tumor antigen human IgG PSA αvβ3 integrin receptor − FAR − − −

MTX MTX MTX MTX MTX MTX MTX MTX DOX DOX DOX DOX DOX DOX DOX SN38

α/γ-amide; PVG↓ LIG α/γ-amide; GFLG α-amide; N↓FF γ-amide α/γ-amide α/γ-amide α/γ-amide α/γ-amide α/γ-amide; LGLG peptide; GFLG Peptide; SGcHQSL peptide-carbamate carbamate maleamide cis-aconitamide cis-aconitamide (Gly)n

PAMAM dendrimer G5/ AuNP168,169 immunoconjugate87 immunoconjugate146

FAR

α-TOS

succinyl amide

protease MMP 1, 13 nd nd nd nd nd lysosomal proteases (cathepsins) protease PSA plasmin, protease nd pH 5 pH 100 h. There was no strong correlation between the linker types and the halflives, though half of the acyl and alkoxycarbonyl hydrazone linkers gave half-lives of less than 10 h. Furthermore, no correlation was observed between half-life and nanocarrier type, suggesting minimal effect of the carrier structure, shape, and size on drug release. Thus, drug conjugation via hydrazone linkage allows for controlled release in response to pH stimuli inside the cell. However, it is noteworthy that hydrazone hydrolysis occurs rather slowly even at pH 5, a condition more

Figure 18. Comparison of half-life values (t1/2, h) in doxorubicin release among three types of hydrazone linkers with variation in nanocarriers, each measured at pH 5. The half-life t1/2 is defined as the time required for half of the initial amount of a doxorubicin− hydrazone linker to be hydrolyzed for drug release. Each data point comes from the reference cited in the text.

3404

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Figure 19. Representative endogenous thiols that are involved in disulfide exchange reactions for drug release in the cytoplasm via disulfide cleavage.

distribution is localized primarily in the cytoplasm (2−10 mM; ≤85%), mitochondria (≤30%), and nucleus (≤10%).289 Such high cytoplasmic concentrations of GSH are due to its involvement in cellular detoxification mechanisms by forming GSH conjugates with cytotoxic xenobiotics including chemotherapeutic agents and by redox reactions with genotoxic reactive oxygen species. In addition, its cellular expression level is inducible and can be increased to levels as high as 10−14 mM. In malignant cells, GSH is often exploited as a means to reduce the cytotoxicity of anticancer therapeutics and has been implicated in the resistance of these cells toward certain drugs. However, such higher expression and reactivity of GSH in cancer cells can also serve as a mechanism for the controlled release of drugs targeted toward these cancer cells. 2.4.2. Design of Disulfide-Tethered Drug Molecules. Drug delivery via disulfide linkage has been extensively investigated for a group of cytotoxic compounds including those used currently in the clinic as summarized in Table 9. Unlike direct conjugation via amide or hydrazone linkage, the disulfide attachment of most anticancer drug molecules to a nanocarrier or cancer specific antibody is performed by indirect methods since each drug molecule lacks a free thiol or disulfide functional group in its chemical structure. Design and incorporation of such a linker in the spacer requires a rational approach in order to achieve the mechanism of controlled drug release. This is illustrated by paclitaxel-disulfide linked molecules (PTX-SS-1,292 PTX-SS-2,292 PTX-SS-3212) and taxoid-linker molecules (PTX-SS-4,210 PTX-SS-5210) in Figure 20. Here the disulfide spacer is attached to paclitaxel at its C-7, C-10, or C-13 side chain attached via a carbonate or ester

acidic than any found in subcellular compartments. In summary, the rate of drug release appears to be determined primarily by the intrinsic rates of hydrazone hydrolysis and is less influenced by the type of linker construct used and the size or shape of the nanocarriers. 2.4. Disulfide Exchange

2.4.1. Chemical Mechanism. Endogenous Thiols. Disulfide linkers constitute one of the primary linkers that have been employed for drug conjugation in targeted drug delivery. Unlike those linkers based on amide, ester, and hydrazone chemistry, the disulfide bond is not subject to hydrolytic cleavage but is cleaved through an electrochemical reduction reaction to yield the respective thiol or through disulfide exchange reactions.285−287 This cleavage reaction occurs via a chemical mechanism triggered by an endogenous thiol molecule such as cysteine, homocysteine, N-acetylcysteine, glutathione (L-γ-glutamyl-L-cysteinyl-L-glycine; GSH), other cysteine-containing peptides, and thiolglycolic acid (Figure 19).288 This disulfide linker is stable enough for targeted intracellular uptake of the drug-linked nanoparticles because the disulfide cleavage occurs primarily in the cytoplasm after their endocytosis. In addition, unless otherwise modified, drug release via cleavage of the disulfide linker is not contributed by redox machinery located on the cell surface but by the thioltriggered intracellular mechanisms.285 2.4.1.1. Cellular Distribution of GSH. Of those endogenous thiols, glutathione (GSH) plays the most significant role in cancer cell specific drug release. GSH exists predominantly in the reduced form inside the cell, and its total cellular 3405

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

(PTX-SS4, PTX-SS5), each a close analogue to ortataxel (BAY59-8862). These taxoid-conjugated mAb molecules were 2−3 orders of magnitude more potent than paclitaxel and docetaxel. In vivo studies with these taxoid immunoconjugates demonstrated strong antitumor efficacy in a paclitaxel-resistant tumor model. Conjugate design via disulfide linkage has been applied to other important classes of anticancer chemotherapeutic agents including cisplatin,295 doxorubicin,295 mitomycin C,162 gemcitabine, maytansine,163,164,296 and calicheamicin294 as summarized in Figure 21. Drug release from each of these drug-linker conjugates is positively demonstrated based on its cytotoxicity in treated cancer cells in vitro or in vivo. However, beyond the cell-based evidence, its release kinetics in solution remains to be determined in a quantitative manner for rigorous evaluation of those disulfide linkers.

Table 9. Disulfide Linkers Designed for Drug Release of Anticancer Therapeutic Agents

a

nanocarrier; targeting ligand

drug linkera

targeting mechanism

FA mAb mAb triazine dendrimer FA FA mAb mAb

MMC-SS-1 DOX-SS-1, 2 PTX-SS-4, 5 PTX-SS-3 PTX-SS-1, 2 GEM-SS-1 MYT-SS-1 CM-SS-1

FAR Ley EGFR − FAR FAR cancer antigen polyepithelial mucin

drug delivered mitomycin C290 doxorubicin291 taxoid210 paclitaxel212 paclitaxel292 gemcitabine293 maytansine163,164 calicheamicin294

For structures of drug-linker code names, see Figures 19−21.

functionality. This combination of functional groups in the spacer is designed for enabling the mechanism of free drug release since a disulfide exchange reaction with GSH triggers the release of only a drug molecule terminated with a thiol moiety. However, since the ester or carbonate is reactive to the nucleophilic thiol, the transient drug intermediate subsequently undergoes a thiol-mediated intramolecular cyclization reaction, and free paclitaxel is released as a result of the formation of 2oxathiolone (PTX-SS-1) or a five-membered thiolactone (PTXSS-2). Ojima et al.210 employed this release mechanism for the design of an anti-EGFR mAb conjugated with a taxoid molecule

2.5. Hypoxia Activation

2.5.1. Biological Mechanism. Hypoxia is defined as the state of abnormally low oxygen supply in tissues and cells.297,298 Like acidosis (pH 6.5−6.9), it constitutes one of the hallmarks of solid tumors, as poor perfusion due to the growth of new immature vessels results in oxygen deprivation (Figure 22). Since tumor microenvironments are associated with low oxygen levels, the equilibrium of the enzymatic activities of oxidoreductases are shifted toward favoring the reduction of substrates. Thus, tumor cells can be selectively targeted by

Figure 20. Design of disulfide-tethered taxol constructs, and mechanisms for GSH-triggered, self-immolative release of paclitaxel or taxoid molecules. 3406

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Figure 21. Design of disulfide-tethered drug molecules for doxorubicin, mitomycin C, gemcitabine, maytansine, and calicheamicin.

several classes depending on the core functionality of the substrate including quinone-trimethyl lock systems,299,306 indolequinone,307 nitroaromatic heterocyclics (nitroimidazole,308 nitrofuran,309 nitrothiofuran310), and N-oxides (tirapazamine311,312) as summarized in Table 10. Representative examples of such hypoxia linkers include the quinone-trimethyl linker system.313 Drug release of this linker has been demonstrated by incubation with chemical agents such as NaBH4 and Na2S2O4, each able to trigger the release by a two electron reduction mechanism.314,315 However, use of potential reducing agents in vivo such as endogenous glutathione (GSH) fails to reduce the quinone to the hydroquinone precursor required to trigger the release, perhaps due to the weaker reduction potential of GSH.316

using substrates and prodrugs that are only activated by such enzymes which are activated in hypoxic conditions as the oxidoreductase (referred to as hypoxia-specific enzymes).200,299,300 Such enzymatic activation has been well documented in the discovery and development of numerous anticancer therapeutic agents such as mitomycin C,301 apaziquone (EO9),302 TH-302,303 banoxantrone (AQ4N), PR104A,304 and RH1.305 2.5.2. Drug Release via Bioreductive Mechanisms. Reductive activation of prodrugs catalyzed by oxidoreductases introduced above has also been employed in the design of certain linker-drug constructs in which the linker is cleaved in response to hypoxia, resulting in drug release in the tumor. Such linker types, dubbed as hypoxia linkers, are grouped into 3407

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

heteroaromatic linkers based on imidazole, (thio)furan, or pyrrole core can also serve as hypoxia-susceptible linkers. Drug release is triggered via an electron push−pull mechanism in which the electron-withdrawing nitro group is reduced by nitroquinone oxidoreductase (NQO) to an electron-donating hydroxyl or amine. This nitro reduction mechanism is also applicable for a conjugated nitrobenzene system as demonstrated by release of paclitaxel upon reduction of the nitro group in a model reaction by chemical treatment.319 Finally, unlike the examples above, certain Pt-based anticancer drug conjugates are designed for release without the involvement of any hypoxic linkers but rather on the basis of the Pt oxidation state. This mechanism is illustrated by Pt(IV)Cl2(NH3)2, an oxidized form of cisplatin attached to a carbon nanotube carrier through its carboxylate axial ligand.28,40,226 This oxidized cisplatin is released in its active form, Pt(II), after exposure to the reductive environment of the cytoplasm. The efficiency of hypoxia-triggered drug release is dictated not only by the design and chemistry of the linker framework but also by physiological factors. For example, the hypoxic conditions found in tumors are the result of poor blood perfusion within the tumor, which also subsequently reduces the amount of exposure of the drug conjugates to the tumor cells. Second, hypoxia can also occur in normal tissue under certain conditions in which oxygen levels are low, which may lead to the reduction in drug specificity to tumor cells.298 However, such physiological factors which lower the selectivity of hypoxia-triggered drug release could be addressed by combining the use of a hypoxia-triggered linker with tumor cell specific targeted nanoparticles that enable cell-specific uptake of the nanoconjugate.

Figure 22. Hypoxia-triggered reductive activation of drug linker in tumor cells. L = linker.

Table 10. Bioreductive Mechanisms for Drug Release in Hypoxiaa redox linker indolequinone307 benzoquinone267,273,274 nitroaromatics

272,276−278,283

release mechanism POR; DT diaphorase (2e) NQO1 (2e); NQO2 (2e) POR

N-oxide311,312

POR

Pt(IV)-carboxylate40,226

reductase; thiol

drug release or activity DNA alkylation; naloxone amines

2.6. Mannich Base

The Mannich reaction refers to the aldehyde-mediated condensation between an amine molecule (primary or secondary) and a nucleophilic molecule including amines, carboxamides, phenols, and ketones (Figure 24). The product of this reaction, commonly called a Mannich base, serves as a prodrug for the parent drug molecule and is effective for improving its solubility and pharmacokinetics. This method is illustrated in the design of a tumor targeted doxorubicin conjugate in which the doxorubicin is conjugated through a formaldehyde-derived linkage to a small Arg-Gly-Asp (RGD) peptide ligand targeting an αvβ3 integrin receptor.116 Another example is doxorubicin linked to imidazolinedione, a nonsteroidal ligand which has micromolar affinity to the androgen receptor (Table 11).320 Koch et al. investigated the release mechanism for doxorubicin delivery in prostate cancer cells.320 Their study showed that this Mannich linker undergoes hydrolytic cleavage to its Schiff base (“imine”) form in the cytosol which undergoes subsequent hydrolysis to free doxorubicin. Interestingly, this androgen-receptor-targeted doxorubicin showed greater cytotoxicity than unmodified doxorubicin in both drug sensitive and drug resistant tumor cells. This remarkable activity was attributed to its additional mode of actioncovalent modification of DNA base pairs by the released doxorubicin Schiff base. As illustrated above, the Mannich base mediated release system serves as an important strategy for targeted drug delivery but has been employed less frequently than other methods (Table 11).116,320 However, this method provides a number of advantages from a synthetic point of view, such as

mustard alkylating agent oxidative DNA damage cisplatin

a

NQO = nitroquinone oxidoreductase; POR = cytochrome P450 oxidoreductase.

Instead, cleavage of this linker in vivo is mediated by cytochrome P450 oxidoreductase (POR) via a two electron (2e−) reduction mechanism, and the resulting hydroquinone intermediate undergoes an intramolecular six-membered-ring cyclization, leading to release of the drug molecule as a leaving group (Figure 23). The role played by three methyl groups involved as part of the extended spacer was originally conceived by Cohen et al. in their linker design for stereopopulation control,317 and aims to lock the conformation of the drug linker in a more favorable position for intramolecular attack by the phenol group and, thus, facilitate the rate of drug release. Another example of the hypoxia linkers is based on the reductive property of indolequinone. As a subgroup of the quinone system, the indolequinone linker is proposed to be activated in part by NAD(P)H:quinone oxidoreductase (DT diaphorase)318 via a two electron mechanism. This system lacks a trimethyl lock, but facilitates drug release via an electron push mechanism. Huang et al. studied this linker for controlled release of naloxone, an opioid receptor antagonist, and demonstrated that the drug molecule is released more selectively and rapidly under hypoxia (t50% = 2.5 h) than normoxia (t5% ≥ 24 h). In addition, nitro-substituted 3408

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Figure 23. Mechanisms of drug release triggered by reductases. POR = cytochrome P450 oxidoreductase; NQO1 = nitroquinone oxidoreductase 1.

2.7. Self-Immolation

A self-immolative reaction process refers to a cascade of spontaneous, intramolecular reactions that occur in response to an applied external stimulus trigger. The self-immolative linker has been considered for applications in controlled drug release.321 As illustrated in Figure 25, this self-immolative

Figure 24. Structure of a doxorubicin molecule tethered to an ArgGly-Asp (RGD) targeting ligand through a Mannich linker based on salicylamide.

chemo-selective conjugation to amines, and tolerance of the conjugation reaction to the presence of diverse functional groups which negates the need for orthogonal protection of these groups. Lastly, the reaction is performed in aqueous− alcoholic conditions which is important for the solubilization of many drug molecules which are often polar and charged.

Figure 25. Schematic illustration of self-immolative release mechanisms: (A) direct release; (B) self-immolative release of a single drug unit; (C) self-immolative release of multiple drug units via application of a single trigger.

Table 11. Doxorubicin-Targeting Ligand Conjugates for Drug Release via Hydrolysis of Mannich Basesa

a

targeting ligand

targeting receptor

drug

linker chemistry

release mechanism

RGD peptide116 imidazolinedione320

αvβ3 integrin androgen receptor

doxorubicin doxorubicin

formaldehyde−salicylamide formaldehyde−salicylamide

acid hydrolysis acid hydrolysis

RGD peptide = CDCRGDCFC; cyclic(N-Me-VRGDf-NH), where “f” refers to D-Phe. 3409

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Table 12. Self-Immolative Linkers for Controlled Drug Release trigger−spacer (Figure 26)

triggering mechanism

immolative reaction type

glycoside−carbamate (A)

β-glucuronidase,322 β-galactosidase323

cephalosporin−carbamate (B) peptide−carbamate; peptide−carbonate (C) disulfide−ester (D) disulfide−hydrazide (E) indolequinone−carbamate (F)

β-lactamase325,326 plasmin,155,165,232 cathepsin B,165 peptidase324 GSH, thiol low pH reductase (DT diaphorase)307

1,6-elimination, cyclization to carbamate 1,4-elimination cyclization to urea, 1,6-elimination mercapto-cyclization327 cyclization to thiadiazinanone135 1,4-elimination; cyclization to urea

linker incorporates dual functional features in its structural design with a triggering moiety linked to a self-immolative spacer rather than directly to the drug molecule. Thus, the drug release mechanism differs from that of direct release linkers due to the involvement of the intervening spacer. Numerous triggering mechanisms have been successfully integrated with an array of self-immolative spacers (Table 12). Each of these trigger moieties is rationally designed for specific cleavage in response to the application of certain external stimuli that include enzymes (glycosidases, 322,323 plasmin,155,165,232 cathepsin B,165 other peptidase,324 β-lactamase325,326), thiol−disulfide exchange,292,327 low pH,135 bioreduction,307 and light.328 Once the triggering moiety is removed from the terminus of the linker, the free spacer group is activated and undergoes spontaneous cyclization or electronic cascade reactions, leading to drug release. Each of such cascade reactions is based on elimination (1,4-; 1,6-) reactions, cyclization of an amine-terminated spacer to a fivemembered urea and carbamate fragment, and cyclization of a mercapto ester or “trimethyl lock”324 spacer to a lactone moiety (Figure 26). One of the primary advantages offered uniquely by selfimmolative release mechanisms relates to the amplification of release frequency through repeated release of multiple drug units per reaction cascade (Figure 25). Proof of concept studies for such amplification have been conducted as demonstrated by full breakdown of multiple branches of a dendritic molecule in response to the single cleavage event of a trigger moiety.328,329 Other beneficial features of these linkers come from the extension of the spacer length. Introduction of such additional spacer length provides relief from steric congestion164,330 between the trigger moiety and a bulkier drug molecule, and is essential for efficient cleavage of the trigger group by macromolecular enzymes such as glycosidases, peptidases, and bioreductive DT diaphorases (Table 12). Furthermore, direct attachment of a certain trigger group such as O-glycoside to a potentially unstable drug molecule can be highly challenging from a synthetic aspect. This synthetic problem can be addressed by synthesizing the trigger-terminated self-immolative linker prior to conjugation to the drug molecule in two separate steps. Given these beneficial features, self-immolation chemistry constitutes an important strategy for drug-linker conjugation in targeted drug delivery.

drug released doxorubicin,322 duocarmycin323 paclitaxel,326 doxorubicin325 paclitaxel,165,232 doxorubicin155,165 taxotere,327 paclitaxel292 calicheamicin135 naloxone307

2.8. Photochemistry

Figure 26. Types of self-immolative linkers and spontaneous release reactions triggered by glycosidase (A), cephalosporinase or βlactamase (B), peptidase (C), glutathione (D), low pH (E), and bioreduction (F).

An important aspect in drug release chemistry is the ability to control the timing of release after uptake of the drug conjugate by the targeted cell. Most of the release mechanisms discussed so far rely on either chemical or enzymatic cleavage of the linker that tethers the drug molecule. Such release reactions occur passively under the influence of specific factors or stimuli. Photochemistry, however, enables an orthogonal release

approach in which light is applied to actively trigger drug release (Figures 27 and 28). This approach is largely based on the concept of photocaging,331,332 in which a drug molecule is temporarily inactivated by derivatization with a photoncleavable trigger group (photocage). This photocaged molecule releases its parent drug molecule in an actively controlled 3410

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Table 13. Light-Controlled Drug Releasea photon-cleavable linker

drug released

light wavelength (nm)

o-nitrobenzyl (ONB)

methotrexate doxorubicin 5-FU tegafur tamoxifen camptothecin doxycycline paclitaxel chlorambucil

365,160,161 254161 365,157 980 (UCN)355 365341 365345 365338,339,343,349,356,357̀

coumarin a

UV348 365342 355,346 365,346 430347 800344 (two-photon)

UCN = upconversion nanocrystal; 5-FU = 5-fluorouracil.

Figure 27. Schematic illustrating the concept of photon-triggered drug release.

Figure 28. Reaction mechanisms for photon-triggered release of a drug molecule linked to o-nitrobenzyl (A) and coumarin (B) linkers.

manner only when the photocage is cleaved by light irradiation (Figure 28). Applications of such photochemical means to control drug release have been demonstrated in numerous types of biological systems331−337 and for drug molecules157,160,161,338−345 including doxorubicin,157 methotrexate,160,161 5-fluorouracil,341 paclitaxel,346,347 camptothecin,348 doxycycline,342 and tamoxifen (Table 13).338,339,343,349 Photon-cleavable linkers that are amenable to release applications are based on those UV-light-responsive aromatic rings comprised of o-nitrobenzyl (ONB),332,335,337,340,350 coumarin,333,335,350 quinoline,351 xanthene,334 and benzophenone.352 Figure 29 illustrates the structures of these photoncleavable linkers. Of these, the ONB group has been frequently applied for drug conjugation to nanocarrriers including PAMAM dendrimers157,160,161 and AuNP.341 This ONB linker allows for high flexibility in its aromatic ring substitution and further derivatization, and allows for synthetic modifications for use in linker installation and drug attachment. ONB is rapidly cleaved by one-photon (254−365 nm157,160,161,341,343) and twophoton (710 nm,337 750 nm340) excitation mechanisms. Coumarin-based linkers represent another major class of

Figure 29. Examples of photocontrolled release of drug molecules.

photon-cleavable linkers in which each drug molecule is covalently attached to a methyl group located at the C-4 position of 7-dialkylaminocoumarin or 6-bromo-7-hydroxycoumarin. These linkers are cleavable by one-photon (365 nm,333,353 475 nm354) as well as by two-photon (740 nm,333,353 800 nm354) mechanisms in which they are reported to provide a greater cross section of two-photon absorption for uncaging for more efficient drug release than the ONB class.333 In addition to light wavelengths, absorption mechanisms (one photon, two photon), and exposure time, the pH condition of the media where the drug release occurs is 3411

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

types of nanocarriers. Liu et al.348 reported shell cross-linked (SCL) micelles with hydrophobic cores conjugated with photocaged camptothecin (ONB-CPT) and coronas functionalized with the FA ligand for tumor cell targeting. These micelles were taken up more efficiently in FAR expressing cancer cells than FAR deficient cells, and showed ∼10-fold enhanced cytotoxicity after UV exposure in the cancer cells. Rotello et al. employed gold nanoparticles (AuNP; hydrodynamic diameter = 10 nm) as the nanocarrier for phtocaged 5fluorouracil (5-FU).341 The AuNP modified with photocaged 5FU on the surface displayed light (365 nm) dependent cytotoxicity in MCF-7 cells which increased as a function of exposure time. In summary, light serves as an effective trigger mechanism for drug release. This photochemical mechanism is applicable in an active manner in contrast to other passive release mechanisms we discussed in earlier sections. It is potentially applicable for those therapeutic and diagnostic applications in vivo that require noninvasive or spatiotemporal drug/probe activation. The technical challenges facing such applications in vivo relate to the limited tissue penetration of UV or visible light, which makes it less efficient than in vitro. However, recent advancements in upconversion nanocrystals (UCNs) open up promising opportunities for the photochemical control of drug release. Certain classes of UCNs show unique optical properties which allow them to emit light in the UV range upon excitation by near-infrared (NIR) light at 980 nm.360−362 In contrast to UV light, NIR has a much deeper range of tissue penetration depth.363 Exposure of photocaged drug molecules carried by the UCNs to NIR light provided a mechanism for efficient doxorubicin release in deep tissues which are otherwise unreachable by UV−vis light.355

considered to be important for determining the rate of photochemical drug release. The release rates are slightly greater in acidic (pH 5.0) or basic (pH 9.0) conditions than in neutral conditions (pH 7.4).161,331,358,359 Such dependency is attributable to pH-dependent changes in the molar absorptivity (ε) of the ONB or coumarin group, the leaving group effect of the drug itself,331,358,359 or the combination of these two. The photocontrolled targeted delivery of anticancer therapeutics was demonstrated by a fifth-generation (G5) PAMAM dendrimer conjugated with a folate (FA) ligand as the targeted carrier for a photocaged doxorubicin (G5(FA)(ONB-Dox; Figure 30).157 This dendrimer conjugate was taken up by folic

2.9. Thermolysis

Chemical bonds for certain types of drug linkers are unstable to thermal stimulus and undergo thermal cleavage, providing a mechanism for drug release. These linkers include an Au−S bond364 and a diazo (NN)365 bond-containing spacer group. Application of heat to a targeted tissue, however, is difficult to control and can cause serious damage at temperatures above a physiological range. One approach to circumvent this problem is to apply the irradiation-induced thermal effect by specialized nanoparticles which release heat in response to irradiation stimuli. This approach enables controlled release and localization of heat around the drug-conjugated nanoparticles after they are taken up or bound by the diseased cells. Iron oxide nanoparticles (IONPs) facilitate such mechanisms due to their ability to generate heat when exposed to an alternating magnetic field (AMF). Riedinger et al.365 demonstrated this release approach by conjugating doxorubicin to the IONP through an azo linker (Figure 31). AMF triggered release of the drug in a cytotoxicity assay on KB cancer cells.

Figure 30. Examples of photocontrolled drug delivery using nanocarriers. G5 PAMAM dendrimer conjugated with the folate ligand and photocaged doxorubicin (A) or photocaged methotrexate (B). Gold nanoparticle (AuNP) conjugated with photocaged 5fluorouracil (5-FU) (C).

acid receptor (FAR)-overexpressing KB cancer cells via FARmediated endocytosis, but was not cytotoxic prior to UV light exposure. Exposure to UV light led to inhibition of cell proliferation as a function of exposure time and a maximum level of the cytotoxicity (70% inhibition) was comparable to that by free doxorubicin (85% inhibition). This result demonstrates that FAR-targeted dendrimer carrying photocaged doxorubicin is only cytotoxic to the target cells once it is exposed to UV light. In a similar approach, targeted photocontrolled delivery of methotrexate (MTX) was demonstrated in KB cells by G5 PAMAM dendrimer conjugated with a photocaged MTX (G5(FA)(ONB-MTX; Figure 30).160 Lighttriggered drug release was also demonstrated by using other

2.10. Azo Reduction

In addition to cleavage via a thermolytic mechanism, the azo linker is also cleavable via a biological mechanism (Figure 32).366,367 This cleavage mechanism has been employed for the activation of an anti-inflammatory class of azo-linked prodrugs of 5-aminosalicylic acid (ASA; mesalazine)368,369 that includes sulfasalazine, balsalazide, ipsalazide, and olsalazine. Sulfasalazine, for example, undergoes reductive cleavage of its azo linker by bacterial enzymes in the colon to an active metabolite, ASA.370 This mechanism has been further applied for the design of colon-targeted polymer therapeutics based on 3412

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Figure 31. Thermal cleavage of an azo linker for doxorubicin release.

physical encapsulation in a container375,376 and play significant roles in drug delivery. The kinetics of drug release is controllable by altering various design features of the nanocarrier such as particle size,377 shape,378 and surface modification chemistry.379,380 Particle size,377,381,382 geometry,383,384 and shape378 are physical properties; however, they are known to play a significant role in dictating the biological activity of the nanoparticles. Processes such as cell surface adhesion, phagocytosis, and degradation have all been shown to be influenced by the size, geometry, and shape of the nanoparticles, as reviewed thoroughly in numerous articles.56,377,378 Surface modification of nanoparticles is one of the fundamental approaches used for altering their biocompatibility56,385,386 as well as for creating sites for host−guest interactions and drug complexation.379,387−397 Thus, the release rates of drug complexes vary in response to the physicochemical properties (charge, hydrophobicity, hydrophilicity) and the density of the drug binding sites.379,398−401 This latter topic has been extensively reviewed for the principles of drug complexation and release kinetics in various nanocarriers such as dendrimers,9,137−139,402 polymers,140 and liposomes.141,142 Here we will briefly discuss selected examples of release mechanisms involved in drug encapsulation that occur through the control of drug diffusion rates by varying pore sizes of encapsulating containers375 and the controlled gating (open/closed) mechanism of the container via alterations in its effective volume, cross-links, and conformation, and the generation of microbubbles. Each of these release mechanisms occurs in response to physicochemical stimuli376 including changes in pH/ion gradient,403−406 reaction with thiols,407−413 exposure to ultrasound,414,415 elevated temperature,376,412,416−420 and light421,422 (Figures 33 and 34, Table 14).

Figure 32. Biological cleavage of an azo linker for drug release.

poly(vinyl amine)371 and 2-hydroxypropyl methacrylate (HPMA) polymer.372,373 Poly(vinyl amine) conjugates with azo-linked ASA and HPMA conjugates with 9-aminocamptothecin have been well-studied.374

3.1. Pore Size

Controlling the size of pores in the delivery vehicle serves as a mechanism of controlled release of guest molecules. This type of release control is illustrated by hollow mesoporous silica nanoparticles (HMSNPs) designed with various pore sizes (diameter = 3.2, 6.4, 12.6 nm) (Figure 33a).375 Li et al. demonstrated that this drug carrier shows differential release rates of its doxorubicin payloads in which the rate of drug diffusion is proportional to the pore size. Accordingly, the duration of drug release can be adjusted by selecting a certain pore size of the vehicle.

3. MECHANISMS OF DRUG RELEASE VIA CONTROL OF CARRIERS The focus of this review is on compiling various release mechanisms of covalently conjugated drug molecules through controlled cleavage of drug linkers. Drug loading and release can also occur through noncovalent mechanisms such as through complexation interactions with the carrier139,140 or 3413

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Figure 33. Mechanisms for alteration of carrier properties for controlled drug release: (a) pore-size-dependent diffusion control; (b) pH-dependent alteration in effective volume, porosity, and permeability; (c) alternative magnetic field (AMF) stimulated uncapping of magnetic nanoparticles (MNPs). Figure 34. Mechanisms of controlled drug release: (a) enhanced permeability via disulfide exchange; (b) configuration-dependent alteration in permeability; (c) ultrasound-mediated perfluorocarbon (PFC) evaporation or vesicle shrinkage; (d) thermally induced generation of bubbles in liposomes.

Compared to mesoporous silica nanoparticles (MSNs) which have been extensively studied,423−428 other types of porous materials have been identified but studied only limitedly for controlled drug release.382,429 Selected examples of these include poly(lactide-co-glycolide) polymer microspheres430 and mesoporous particles made of aluminosilicates (MCM41 431 ). Each of these studies tested piroxicam 430 or ibuprofen431 as the drug payload and suggests that the rate of drug release is inversely proportional to the pore size.

in response to the pH condition of the solution. Voit et al. demonstrated that the release of encapsulated doxorubicin from cross-linked polymersomes of a block copolymer (PEG-bP(DEAEMA-BMA)) is tunable by pH-dependent permeability of the polymer brushes.403 At pH 5 the release is on state due to polymer expansion, while at pH 7.4 it is off state due to polymer collapse. Magnetic nanoparticles have the ability to produce heat when subjected to a localized alternating magnetic field (AMF). Such magnetic hyperthermia constitutes a mechanism used for controlled release. Ruiz-Hernández et al.433 reported on the design of mesoporous silica nanoparticles (MSN; diameter = 100 nm) loaded with a fluorescein payload (Figure 33c). In this system, the surface of each MSN is premodified by conjugation with multiple residues of a single-stranded cDNA. This cDNA oligo serves as a linker for noncovalent attachment of an iron oxide magnetic nanoparticle conjugated with the complementary oligo. Each magnetic NP is thus, in essence, capped in the area proximal to the pore entrance, blocking payload release. When an AMF is applied, the magnetic NPs are heated to temperatures of ≥42 °C, which is high enough to melt the cDNA linker. As a result of DNA dehybridization, the magnetic

3.2. Effective Volume

Another method to control release is to alter the effective volume of the vehicle available for drug encapsulation by applying external stimuli. For example, after cellular uptake by cancer cells, the loading capacity of the drug encapsulating nanosystem can be lowered by a decrease in pH (endosomes, lysosomes),403−406 ultrasound stimulation,414 and temperature alteration.376,416,417,419,420 This release mechanism frequently is used for polymers152 and polymer grafted core−shell systems based on liposomes404 and inorganic nanoparticles414,417,432 because the local density of certain polymer residues on the liposome surface increases with the protonation of the polymer pendant groups, triggering the collapse of the liposomal membrane and release of its payloads. This is illustrated by poly(acrylic acid)-grafted liposomes loaded with arsenic trioxide (As2O6) as an anticancer agent (Figure 33b).404 However, with the use of a different type of polymer sensor, a mechanism for drug release can be designed in a different way 3414

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Table 14. Mechanisms of Drug Release via Control of Carrier Properties for Payload Encapsulation trigger

carrier

mechanism

mesoporous silica NP375,453,454 PAA-grafted liposome,404 PEG cross-linked nanogel,455 polymersome,403,456 mesoporous silica NP,457 lipid/polymer conjugate,458 AuNP-covered mesoporous silica NP459 disulfide cross-linked PEG hydrogel,408,455,460 cationic polymer,439 polymer micelle,440,441,461,462 AuNP-coated polymer micelle,463 mesoporous silica NP412

diffusion rate altered porosity

light

azobenzene-coated silica NP,422 liposome,464 azobenzene block polymer,465 graphene oxide466

temperature

cellulose-grafted magnetic particle,418 liposome,467 PMAA polymer468

cis−trans isomerization, altered porosity altered porosity

pore size pH gradient thiol (GSH)

ultrasound nanoemulsion,415 polymer micelle,414 mesoporous silica particle469 magnetic field iron oxide NP (IONP),433,434 zinc-doped IONP436

NPs are uncapped, and the fluorescein molecules are released through the open pore. This magnetic control of hyperthermia mediated release has been similarly applied in other systems as well,434−436 including a nanovalve437 which has been shown to be effective for controlled gating.

disulfide exchange

altered porosity localized hyperthermia

payload released doxorubicin,375,454 trypsin inhibitor453 arsenic trioxide (As2O6),404 doxorubicin,403,455,458,459 rutheniumpyridyl,457 prednisolone456 DNA plasmid,408,439 doxorubicin,412,440,460−462 methotrexate,441 paclitaxel463 irinotecan,422 doxorubicin464,466 gemcitabine,418 doxorubicin,419,420,467 daunorubicin468 paclitaxel,415 doxorubicin414,469 fluorescein,433,434 doxorubicin,436 methylene blue435

conformational switch as well.2 Typically these polymers or their micelles are composed of pH-responsive block copolymers such as poly(histidine-b-PEG-folate) and poly(lactic acidb-PEG-folate).446 Other mechanisms frequently used for conformational control are based on ultrasound irradiation414,449 and temperature alteration.376,416,417,419,420,450,451

3.3. Cross-Links

3.5. Microbubbles

Alteration of the effective volume and internal permeability of a delivery vehicle is also controllable by reduction of disulfide cross-links used for stabilizing the interior of nanoparticles and for payload encapsulation.407−411,413,438,439 This release mechanism is applied for DNA delivery using a PEG-based hydrogel in which exposure to GSH led to cleavage of the disulfide crosslinks, and facilitated DNA release after its intracellular uptake (Figure 34a). It serves as a route for tumor targeted gene modulation because of the higher expression levels of GSH in tumor cells.408 This mechanism has been applied for the controlled release of doxorubicin and methotrexate encapsulated in disulfide cross-linked polymer micelles.440,441

Ultrasound plays an active role in the controlled release of drug payloads.414,415 A number of nanocarriers including nanoemulsions415,452 and polymer vesicles414 have been investigated for this release mechanism.376 While exact release mechanisms vary with the specific design features of each nanocarrier, the release is attributable to the ability of ultrasound to physically stimulate individual structural components that comprise the nanocarrier. This is illustrated by a nanoemulsion droplet system in which paclitaxel is entrapped in combination with perfluorocarbon (PFC) as an ultrasound-responsive gas molecule (Figure 34c).415 Exposure of the nanodroplet to ultrasound irradiation triggered the generation and escape of microbubbles filled with paclitaxel, resulting in an increase in the effective volume of the nanodroplet. Application of ultrasound for controlled drug release is also demonstrated by using polymer vesicles prepared by the self-assembly of the amphiphilic block copolymer (PEO-b-P(DEA-stat-TMA) carrying doxorubicin as the payload.414 Ultrasound irradiation was selected for this application due to its ability to alter the structural morphology of the vesicles. In particular, ultrasound was able to alter the size of the nanoparticles, decreasing the size (mean hydrodynamic diameter = 377 nm) 3 times (124 nm) following irradiation. The authors suggest that this shrinkage accounts for the release of drug molecules because it directly affects the effective volume of the hydrophobic core housing the payload. Thermal-stimulus-triggered release mechanisms have been investigated using temperature-responsive liposomes as the drug carrier. Advantages of using liposomes for temperature control include the ability to simultaneously encapsulate several types of molecules including a thermally responsive component. Chen et al.419 encapsulated doxorubicin and ammonium bicarbonate together in liposomes (Figure 34d). Ammonium bicarbonate is thermally unstable and begins to decompose at an elevated temperature (42 °C), resulting in the generation of bubbles made up of CO2 and ammonia gas. These bubbles create defects in the lipid bilayer structure resulting in the release of entrapped doxorubicin molecules. Another liposome system is reported by Al-Ahmady et al.420 employing an

3.4. Conformation

Light serves an effective tool to alter the internal configuration and permeability of a delivery system.421,442 Unlike use of a photocleavable linker to a drug molecule, the linker is incorporated as the building block of the delivery system such as in lipid membranes.421 Best et al. designed ONBincorporated lipid and demonstrated light triggered release of dye molecules loaded in the liposomes made of the photocleavable lipid.421 In addition to irreversible cleavage of the photosensitive linker, light triggers a reversible change in the linker configuration such as from trans-to-cis isomerization of azobenzene, and this change leads to a decrease in the volume available for drug encapsulation as demonstrated by the faster rate of release of irinotecan encapsulated in azobenzenecoated silica nanoparticles (Figure 34b).422 In addition to light, there are other types of stimuli that have been investigated for controlled drug release through conformational changes of nanocarriers.2,376 Some of these mechanisms were already discussed in section 3.2.152,405,414,416 One of the mechanisms is control by pH, as illustrated by Abraxane, the albumin-bound paclitaxel delivery system.443,444 Its mode of action is believed to involve albumin receptor-mediated intracellular uptake and subsequent drug release triggered by low pH-induced protein denaturation in the acidic endosomal compartment (Table 5). Nonprotein carriers that include sterically stabilized liposomes,445 polymer nanoparticles, or polymeric micelles446−448 are believed to utilize a pH-triggered 3415

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

localized primarily in the cytoplasm (2−10 mM; ≤85%) and mitochondria (≤30%). 4.1.5. Hypoxia Activation. Certain linkers based on indolequinone and nitroaromatics are reductively cleaved by hypoxia-activated enzymes. Thus, use of such linkers allows tumor-targeted drug release. However, their tumor selectivity can be compromised because hypoxia can also occur in normal tissues under certain conditions in which oxygen levels are low. 4.1.6. Mannich Base. Drug conjugation via a Mannich base link is applicable for certain classes of drug molecules functionalized with amines, carboxamides, phenols, or ketones. The release mechanism for this Mannich conjugated drug molecule involves hydrolytic cleavage to its Schiff base form in acidic cellular compartments followed by hydrolysis to release the parent drug. In contrast to amide and ester linkers, drug conjugation through a Mannich base can be performed in aqueous solutions without the need to protect the other functional groups. 4.1.7. Self-Immolation. Release of a drug molecule conjugated via a self-immolative linker begins by activation of a triggering moiety linked to the self-immolative spacer. Numerous triggering mechanisms have been designed to occur in response to external stimuli that include enzymes, low pH, bioreduction, and light. The hallmark of this release mechanism relates to the amplification of release frequency through repeated release of multiple drug units per reaction cascade. However, conjugation through a self-immolative linker for targeted delivery requires convergent synthetic approaches and thus can be challenging to integrate with existing carrier platforms. 4.1.8. Photochemistry. Release of a drug molecule attached through a photocleavable linkage is triggered by light exposure and serves as an effective mechanism for spatiotemporal control of drug release. Photochemical control is challenging for applications in subepidermal cells and tissues due to the limited light penetration of radiation sources such as UV light. However, these limitations are currently being addressed by the use of two-photon excitation which has a deeper penetration range, as it utilizes visible light (740−800 nm). In addition, recent developments in biophotonic nanomaterials have opened up several promising applications for photon-controlled drug release. In particular, the use and development of upconversion nanocrystals (UCNs)360−362 which are excited by IR radiation (980 nm)363 and emit UV light has great potential for photochemical release applications in deep tissue. 4.1.9. Thermolysis. Thermolysis relies on a thermal stimulus to trigger cleavage of the chemical bond that constitutes the linker such as an Au−S bond or a diazo (N N) bond. Despite its simple mechanism of drug release, application of this mechanism is challenging due to the lack of effective methods for localized heat exposure to only the targeted tissue. However, this issue is addressed in part by integration of the linker with specialized magnetic nanoparticles which generate a localized thermal effect upon exposure to an alternating magnetic field (AMF). 4.1.10. Azo Reduction. Biological cleavage of an azo linker by enzymes in the colon allows for colon-targeted drug delivery, giving this linker great potential for use in extended release of drug molecules associated with gastrointestinal cancers and irritable bowel diseases. 4.1.11. Encapsulation. Parent drug molecules can also be loaded through noncovalent complexation with and/or physical

amphiphilic leucine zipper peptide. Incorporation of this peptide in the lipid bilayer occurs through a self-assembly process that results in a stable bundle of peptides inserted in the membrane. When the liposomes are exposed to an elevated temperature, the peptides in the bundle become highly disordered in conformation and open pores are formed in the lipid bilayer large enough for drug release.

4. CONCLUSIONS 4.1. Summary

Targeted drug delivery using nanometer-sized carriers aims to deliver therapeutic payloads precisely to targeted cells. Key design elements that play an essential role for enabling this strategy are comprised of selective cell targeting by multivalent ligand design and subsequent cell inhibition by timely release of carried drug molecules. The release of drug needs to occur in a precisely controlled manner in order to deliver the cytotoxic payload to the targeted cell only. Here we systematically reviewed various mechanisms of drug release developed for targeted delivery from a chemical and biochemical point of view. These release mechanisms are classified on the basis of drug linker functionalities, and are summarized below with regard to rationale, linker properties, and their inherent limitations. We believe this review article will help give a better understanding of the molecular basis of each of these release mechanisms and provide important insights for the rational design of drug release systems. 4.1.1. Ester Hydrolysis. Drug attachment through a carboxylic ester linker is widely used due to its synthetic convenience and general applicability to numerous therapeutic molecules, as most of these molecules have a carboxylic acid or alcohol functional group. Controlled drug release through cleavage of the ester bond is facilitated by (i) esterases and hydrolases in acidic endosomes and lysosomes, (ii) certain hydrolytic enzymes overexpressed in the extracellular membrane or plasma, and, (iii) to a lesser extent, acid- or basecatalyzed hydrolysis. 4.1.2. Amide Hydrolysis. Like the ester linker, drug attachment via an amide linker is widely applicable due to synthetic convenience and functional compatibility. In comparison to the ester bond, the amide bond is much more stable and less susceptible to chemical hydrolysis unless certain classes of specialized linkers are employed such as citraconyl, cis-aconityl, and maleyl groups. Thus, release of amide-linked drug molecules occurs via enzymatic mechanisms through the activity of hydrolytic proteases localized in the plasma, ECM, plasma membrane, and intracellular lysosomes. 4.1.3. Hydrazone Hydrolysis. Drug conjugation via a hydrazone linkage is applicable for drug molecules containing a ketone or aldehyde functional moiety. This linker is fairly stable under physiological conditions but undergoes rapid cleavage in acidic conditions (pH ≤5), allowing for pH-regulated drug release. Thus, acid-responsive release occurs after internalization into the acidic environments of endosomes and lysosomes (pH 5−6) rather than in the extracellular environment (pH >6). 4.1.4. Disulfide Exchange. Unlike the ester, amide, and hydrazine linkers, the disulfide bond is not cleavable by hydrolysis but is cleaved through reduction or disulfide exchange reactions. Thus, drug release is triggered by endogenous thiol molecules including glutathione (GSH) 3416

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Table 15. Polymer-Based Nanotherapeutics Approved or Advanced to Clinical Studiesa conjugate name

polymers

drug (therapeutic area)

linker chemistry (release mechanism)

Xyotax (CT-2103)211,470−472 PNU16645473,474 CT-2106472 XMT-1001475,476 MAG-CPT (PNU166148)477 CRLX101 (IT-101)478 Delimotecan (T-2513)479 PK1, PK2480,481 NK911447,482−484 ProLindac (AP5280)475,485,486 HAS-MTX487,488 XMT-1107475,489 MOVANTIK (naloxegol; NKTR-118)490

poly(L-Glu) HPMA poly(L-Glu) PHF (Fleximer) HPMA CD−PEG carboxymethyl dextran HPMA PEG−poly(L-Asp) HPMA HSA PHF (Fleximer) PEG

paclitaxel (cancer) paclitaxel (cancer) camptothecin (cancer) camptothecin (cancer) camptothecin (cancer) camptothecin (cancer) camptothecin (cancer) doxorubicin (cancer) doxorubicin (cancer) cisplatin (cancer) methotrexate (cancer) fumagillol (cancer) naloxone (OIC)

Glu ester (cathepsin B211) GFLG ester (lysosomal peptidases) Gly ester (cathepsin B211) GGFG amide (intramolecular transacylation) Gly ester Gly ester amide GFLG amide (lysosomal peptidases481) Asp amide malonate-Pt (hydrolysis) lysine amide (lysosomal peptidases487) carbamate (plasma489) ether (C6)

HPMA = N-(2-hydroxypropyl)methacrylate; CD−PEG = β-cyclodextrin−poly(ethylene glycol) copolymer; HSA = human serum albumin; PHF = poly(1-hydroxymethylethylene hydroxymethylformal); OIC = opioid-induced constipation. a

4.3. Challenges in Nanotherapeutic Linker Design

encapsulation in the nanocarrier. Release then occurs in a diffusion controlled manner. Systems designed for such diffusion control include variation in pore sizes of the encapsulating containers and controlled gating of the container that is triggered in response to such stimuli as pH/ion gradient, thiol, ultrasound, temperature, and light.

A number of challenges exist in the optimal design of nanotherapeutics. We address certain aspects that need to be considered in the drug linker design. 4.3.1. Synthetic Chemistry. Drug conjugation through certain linker types such as photocleavable, hypoxia, and disulfide linkers require prederivatization of each drug molecule with a custom-tailored linker. Synthesis of such linkers and subsequent drug modification require multiple steps of synthetic operation and can be more challenging to nonchemists in particular during the scale-up process. This opens up an area of multidisciplinary collaboration among synthetic chemists, biologists, and biomedical engineers. 4.3.2. Binding Avidity. Co-attachment of drug molecules with targeting ligands on the surface of a nanocarrier increases steric congestion491,492 around the ligands, and thus can interfere with binding avidity of the nanoconjugate to the targeted cell, resulting in lower cell selectivity. Use of a dualacting drug molecule that can serve as a targeting ligand as well can overcome such a potential issue in nanotherapeutic design. Methotrexate is a good example of such dual-acting molecules because of its affinity to FAR.98,252,268,493 This unavoidable design issue is also addressable by optimizing the length and conformational flexibility of a spacer to the drug or ligand molecule. Use of a shorter spacer for the drug molecule reduces steric hindrance around the ligand. 4.3.3. Physicochemical Properties. Heavy payloads of hydrophobic drug molecules carried by a single nanoparticle result in low aqueous solubility. Hydrophilic groups in the spacer that have high polarity and charge such as amide, ethylene oxide (EO), carboxylic acid, and amine are more effective in improving the solubility of the conjugate than hydrophobic alkane spacers. 4.3.4. Pharmacokinetics (PK). In vivo efficacy of a nanotherapeutic is also determined by its PK properties during systemic circulation such as metabolic stability, low clearance rate, and minimal level of premature drug release. Prediction of these PK properties is difficult but can be evaluated in part by performing in vitro PK studies through measurement of conjugate stability and nonspecific drug release kinetics in PBS and serum solutions. Optimization of PK properties is achievable by varying the structural features and chemical compositions (size, shape, polarity, charge) of a nanocarrier, the linker, and the spacer. Excessive charge is often linked to

4.2. Status of Nanotherapeutic Development

Drug delivery in nanotechnology has served as a promising strategy for improving safety and efficacy profiles of existing therapeutic agents which are otherwise associated with poor aqueous solubility, plasma instability, generally high toxicity, and serious side effects.382,390,393,394,474,475,490 Development of such nanotherapeutic agents have been focused primarily on cancer treatment, but recently has been expanding to other therapeutic applications including opioid-induced constipation.490 Currently there are dozens of polymer-based nanotherapeutics which have been approved or evaluated in clinical trials (Table 15). Most clinical candidates have been focused on improving the therapeutic performance of anticancer compounds with poor aqueous solubilities such as paclitaxel and camptothecin, as well as those causing serious systemic toxicity such as doxorubicin and cisplatin. However, the progress of these clinical candidates toward regulatory approval has been slow, resulting in only one approved drug (Xyotax),190,381−383 with many others still in clinical trial stages (eg., ProLindac, XMT-1001). Several candidates resulted in clinical failures,473,474,477 though such poor outcomes are, in principle, attributable to their suboptimal design features due to lack of targeting ligands, nonspecific release mechanisms, or a combination of these factors. Many of the current candidates that have been developed rely on a passive mechanism of tumor targeting, the EPR effect,69−71 which is known to occur in leaky tumor vasculatures (section 1.3). Recent candidates, in contrast, incorporate an active targeting mechanism as illustrated by a hepatic tumor targeting conjugate in which multivalent presentation of a galactosamine ligand enables binding to liver cells with high avidity.480,481 Looking forward, the combination of active tumor targeting with precise control of drug release through the mechanisms discussed in this review is expected to play a more crucial role in disease-specific drug delivery and personalized medicine. 3417

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

adsorption of serum proteins to the nanocarrier, and large sizes result in prolonged tissue retention, slow renal clearance (hydrodynamic diameter >8 nm),494 and systemic toxicity.494,495 4.3.5. Tissue Penetration. Targeting cells through an ECM or dense layer of tissues adds another challenge compared to endothelial cells. Such conjugate diffusion is a slower process in which premature drug release can occur via the metabolic breakdown of the drug linker by enzymes secreted in the ECM including MMPs. Use of a nonpeptidic drug linker inert to MMP activity is needed for achieving intracellular drug release. However, such an MMP-selective release mechanism is also considered as a route for tumor targeting since it enables localized drug exposure to tumor cells overexpressing MMPs (section 2.2.1).141,213 In summary, nanocarriers are a very powerful tool for drug delivery when combined with precise mechanisms for spatially and temporally controlling drug release. Conjugation of currently used drugs to nanocarriers will likely lead to better strategies for improving the therapeutic index of these drugs and significantly enhance our ability to treat several human diseases.

Seok Ki Choi earned his B.S. and M.S. degrees at Seoul National University, Korea, and Ph.D. degree at Columbia University, New York, NY (advisor Prof. Koji Nakanishi), and then completed his postdoctoral training at the laboratory of Prof. George M. Whitesides, Harvard University, Cambridge, MA. In mid-1997, he joined Theravance, Inc. (then, Advanced Medicine), South San Francisco, CA, where he developed expertise in translation of the polyvalent concept for small molecule drug discovery. In late 2008, he moved to the University of Michigan Medical School, Ann Arbor, MI, and currently serves as a faculty member in the Department of Internal Medicine with a joint appointment at the Michigan Nanotechnology Institute for Medicine and Biological Sciences. His research interests are focused on biologic nanotechnology, multivalency, photogenetic probes, and biophotonic nanomaterials.

AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. Tel.: (734) 647-0052. Fax: (734) 936-2990.

ACKNOWLEDGMENTS The authors thank the Michigan Nanotechnology Institute for Medicine and Biological Sciences for support, and the University of Michigan Office of the Vice President for Research and Shanghai Jiao Tong University (SKC) for funding.

Notes

The authors declare no competing financial interest. Biographies

ABBREVIATIONS AMF alternating magnetic field AuNP gold nanoparticle CD β-cyclodextrin DHFR dihydrofolate reductase DOX doxorubicin ECM extracellular matrix EGF epidermal growth factor EGFR epidermal growth factor receptor EPR enhanced permeation and retention FA folic acid FAR folate receptor FGF fibroblast growth factor FGFR fibroblast growth factor receptor 5-FU 5-fluorouracil GSH glutathione HA hyaluronic acid HSA human serum albumin HMSNP hollow mesoporous silica nanoparticle HPMA N-(2-hydroxypropyl)methacrylamide IONP iron oxide nanoparticle MMP matrix metalloproteinase MTX methotrexate MWNP multiwalled carbon nanotube NIR near infrared NP nanoparticle

Pamela T. Wong received her Ph.D. in biological chemistry from the University of Michigan, Ann Arbor, MI, in 2009 under the guidance of Prof. Ari Gafni, where her research focused on the biochemical and biophysical characterization of amyloid-beta oligomerization and interaction with liposomal membrane surfaces. Currently, she is a postdoctoral fellow with Prof. James R. Baker, Jr., at the University of Michigan Medical School Nanotechnology Institute for Medicine and Biological Sciences. Her research interests focus on the design and biological characterization of nanoemulsion based adjuvants for subunit vaccines and on nanoparticle based targeted delivery and release systems including dendrimer, AuNP, and upconversion nanocrystal platforms. 3418

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews NQO ONB PAMAM PEG PEI PHF PPI PLGA POR PSA PSMA PTX RF RFR SWNT α-TOS UCN UV TI

Review

(16) Tomalia, D. A.; Naylor, A. M.; William, A.; Goddard, I. Starburst Dendrimers: Molecular-Level Control of Size, Shape, Surface Chemistry, Topology, and Flexibility from Atoms to Macroscopic Matter. Angew. Chem., Int. Ed. 1990, 29, 138−175. (17) Tomalia, D. A.; Baker, H.; Dewald, J.; Hall, M.; Kallos, G.; Martin, S.; Roeck, J.; Ryder, J.; Smith, P. A New Class of Polymers: Starburst-Dendritic Macromolecules. Polym. J. (Tokyo, Jpn.) 1985, 17, 117−132. (18) Zimmerman, S. C.; Quinn, J. R.; Burakowska, E.; Haag, R. Cross-Linked Glycerol Dendrimers and Hyperbranched Polymers as Ionophoric, Organic Nanoparticles Soluble in Water and Organic Solvents. Angew. Chem., Int. Ed. 2007, 46, 8164−8167. (19) Zhou, Z.; Shen, Y.; Tang, J.; Jin, E.; Ma, X.; Sun, Q.; Zhang, B.; Van Kirk, E. A.; Murdoch, W. J. Linear polyethyleneimine-based charge-reversal nanoparticles for nuclear-targeted drug delivery. J. Mater. Chem. 2011, 21, 19114−19123. (20) Fischer, M.; Vö gtle, F. Dendrimers: From Design to ApplicationA Progress Report. Angew. Chem., Int. Ed. 1999, 38, 884−905. (21) Bosman, A. W.; Janssen, H. M.; Meijer, E. W. About Dendrimers: Structure, Physical Properties, and Applications. Chem. Rev. (Washington, DC, U. S.) 1999, 99, 1665−1688. (22) Zhang, W.; Nowlan, D. T.; Thomson, L. M.; Lackowski, W. M.; Simanek, E. E. Orthogonal, Convergent Syntheses of Dendrimers Based on Melamine with One or Two Unique Surface Sites for Manipulation. J. Am. Chem. Soc. 2001, 123, 8914−8922. (23) Sideratou, Z.; Kontoyianni, C.; Drossopoulou, G. I.; Paleos, C. M. Synthesis of a folate functionalized PEGylated poly(propylene imine) dendrimer as prospective targeted drug delivery system. Bioorg. Med. Chem. Lett. 2010, 20, 6513−6517. (24) Paleos, C. M.; Tsiourvas, D.; Sideratou, Z.; Tziveleka, L. Acidand Salt-Triggered Multifunctional Poly(propylene imine) Dendrimer as a Prospective Drug Delivery System. Biomacromolecules 2004, 5, 524−529. (25) Kaminskas, L. M.; Kelly, B. D.; McLeod, V. M.; Sberna, G.; Owen, D. J.; Boyd, B. J.; Porter, C. J. H. Characterisation and tumour targeting of PEGylated polylysine dendrimers bearing doxorubicin via a pH labile linker. J. Controlled Release 2011, 152, 241−248. (26) Chertok, B.; David, A. E.; Yang, V. C. Polyethyleneiminemodified iron oxide nanoparticles for brain tumor drug delivery using magnetic targeting and intra-carotid administration. Biomaterials 2010, 31, 6317−6324. (27) Chen, H.-T.; Neerman, M. F.; Parrish, A. R.; Simanek, E. E. Cytotoxicity, Hemolysis, and Acute in Vivo Toxicity of Dendrimers Based on Melamine, Candidate Vehicles for Drug Delivery. J. Am. Chem. Soc. 2004, 126, 10044−10048. (28) Dhar, S.; Gu, F. X.; Langer, R.; Farokhzad, O. C.; Lippard, S. J. Targeted delivery of cisplatin to prostate cancer cells by aptamer functionalized Pt(IV) prodrug-PLGA-PEG nanoparticles. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 17356−17361. (29) Eniola, A. O.; Hammer, D. A. Artificial polymeric cells for targeted drug delivery. J. Controlled Release 2003, 87, 15−22. (30) Farokhzad, O. C.; Cheng, J.; Teply, B. A.; Sherifi, I.; Jon, S.; Kantoff, P. W.; Richie, J. P.; Langer, R. Targeted nanoparticle-aptamer bioconjugates for cancer chemotherapy in vivo. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 6315−6320. (31) Gu, F.; Zhang, L.; Teply, B. A.; Mann, N.; Wang, A.; RadovicMoreno, A. F.; Langer, R.; Farokhzad, O. C. Precise Engineering of Targeted Nanoparticles by Using Self-Assembled Biointegrated Block Copolymers. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 2586−2591. (32) Chandran, S. S.; Nan, A.; Rosen, D. M.; Ghandehari, H.; Denmeade, S. R. A prostate-specific antigen-activated N-(2-hydroxypropyl) methacrylamide copolymer prodrug as dual-targeted therapy for prostate cancer. Mol. Cancer Ther. 2007, 6, 2928−2937. (33) Homma, A.; Sato, H.; Okamachi, A.; Emura, T.; Ishizawa, T.; Kato, T.; Matsuura, T.; Sato, S.; Tamura, T.; Higuchi, Y.; Watanabe, T.; Kitamura, H.; Asanuma, K.; Yamazaki, T.; Ikemi, M.; Kitagawa, H.; Morikawa, T.; Ikeya, H.; Maeda, K.; Takahashi, K.; Nohmi, K.; Izutani, N.; Kanda, M.; Suzuki, R. Novel hyaluronic acid-methotrexate

nitroquinone oxidoreductase o-nitrobenzyl poly(amidoamine) poly(ethylene glycol) poly(ethylene imine) poly(1-hydroxymethylethylene hydroxymethylformal) poly(propylene imine) poly(D/L-lactic acid-co-glycolic acid) P450 oxidoreducatse prostate-specific antigen prostate-specific membrane antigen paclitaxel riboflavin riboflavin receptor single-walled carbon nanotube α-tocopheryl succinate upconversion nanocrystal ultraviolet therapeutic index

REFERENCES (1) Farokhzad, O. C.; Langer, R. Impact of Nanotechnology on Drug Delivery. ACS Nano 2009, 3, 16−20. (2) Peer, D.; Karp, J. M.; Hong, S.; Farokhzad, O. C.; Margalit, R.; Langer, R. Nanocarriers as an emerging platform for cancer therapy. Nat. Nanotechnol. 2007, 2, 751−760. (3) Low, P. S.; Henne, W. A.; Doorneweerd, D. D. Discovery and Development of Folic-Acid-Based Receptor Targeting for Imaging and Therapy of Cancer and Inflammatory Diseases. Acc. Chem. Res. 2008, 41, 120−129. (4) Majoros, I. J.; Williams, C. R.; Becker, A.; Baker, J. R., Jr. Methotrexate Delivery via Folate Targeted Dendrimer-Based Nanotherapeutic Platform. Wiley Interdiscip. Rev.: Nanomed. Nanobiotechnol. 2009, 1, 502−510. (5) Chari, R. V. J. Targeted Cancer Therapy: Conferring Specificity to Cytotoxic Drugs. Acc. Chem. Res. 2008, 41, 98−107. (6) Ojima, I. Guided Molecular Missiles for Tumor-Targeting Chemotherapy; Case Studies Using the Second-Generation Taxoids as Warheads. Acc. Chem. Res. 2008, 41, 108−119. (7) Brigger, I.; Dubernet, C.; Couvreur, P. Nanoparticles in cancer therapy and diagnosis. Adv. Drug Delivery Rev. 2002, 54, 631−651. (8) Svenson, S.; Tomalia, D. A. Dendrimers in biomedical applicationsreflections on the field. Adv. Drug Delivery Rev. 2005, 57, 2106−2129. (9) Medina, S. H.; El-Sayed, M. E. H. Dendrimers As Carriers for Delivery of Chemotherapeutic Agents. Chem. Rev. (Washington, DC, U. S.) 2009, 109, 3141−3157. (10) Patri, A. K.; Majoros, I. J.; Baker, J. R. Dendritic polymer macromolecular carriers for drug delivery. Curr. Opin. Chem. Biol. 2002, 6, 466−471. (11) Gharib, M. I.; Burnett, A. K. Chemotherapy-induced cardiotoxicity: current practice and prospects of prophylaxis. Eur. J. Heart Failure 2002, 4, 235−242. (12) Kukowska-Latallo, J. F.; Candido, K. A.; Cao, Z.; Nigavekar, S. S.; Majoros, I. J.; Thomas, T. P.; Balogh, L. P.; Khan, M. K.; Baker, J. R., Jr. Nanoparticle Targeting of Anticancer Drug Improves Therapeutic Response in Animal Model of Human Epithelial Cancer. Cancer Res. 2005, 65, 5317−5324. (13) Myc, A.; Majoros, I. J.; Thomas, T. P.; Baker, J. R. DendrimerBased Targeted Delivery of an Apoptotic Sensor in Cancer Cells. Biomacromolecules 2007, 8, 13−18. (14) Majoros, I. J.; Williams, C. R.; Baker, J., Jr. Current Dendrimer Applications in Cancer Diagnosis and Therapy. Curr. Top. Med. Chem. 2008, 8, 1165−1179. (15) Pan, X.; Lee, R. J. Tumour-selective drug delivery via folate receptor-targeted liposomes. Expert Opin. Drug Delivery 2004, 1, 7−17. 3419

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

conjugates for osteoarthritis treatment. Bioorg. Med. Chem. 2009, 17, 4647−4656. (34) Luo, Y.; Prestwich, G. D. Synthesis and Selective Cytotoxicity of a Hyaluronic Acid−Antitumor Bioconjugate. Bioconjugate Chem. 1999, 10, 755−763. (35) Chau, Y.; Dang, N. M.; Tan, F. E.; Langer, R. Investigation of targeting mechanism of new dextran-peptide-methotrexate conjugates using biodistribution study in matrix-metalloproteinase-overexpressing tumor xenograft model. J. Pharm. Sci. 2006, 95, 542−551. (36) Farokhzad, O. C.; Jon, S.; Khademhosseini, A.; Tran, T.-N. T.; LaVan, D. A.; Langer, R. Nanoparticle-Aptamer Bioconjugates: A New Approach for Targeting Prostate Cancer Cells. Cancer Res. 2004, 64, 7668−7672. (37) Fuchs, P. C.; Barry, A. L.; Brown, S. D. PMX-622 (Polymyxin BDextran 70) Does Not Alter In Vitro Activities of 11 Antimicrobial Agents. Antimicrob. Agents Chemother. 1998, 42, 2765−2767. (38) Yura, H.; Yoshimura, N.; Hamashima, T.; Akamatsu, K.; Nishikawa, M.; Takakura, Y.; Hashida, M. Synthesis and pharmacokinetics of a novel macromolecular prodrug of Tacrolimus (FK506), FK506−dextran conjugate. J. Controlled Release 1999, 57, 87−99. (39) Liu, Z.; Tabakman, S. M.; Chen, Z.; Dai, H. Preparation of carbon nanotube bioconjugates for biomedical applications. Nat. Protocols 2009, 4, 1372−1381. (40) Dhar, S.; Liu, Z.; Thomale, J.; Dai, H.; Lippard, S. J. Targeted Single-Wall Carbon Nanotube-Mediated Pt(IV) Prodrug Delivery Using Folate as a Homing Device. J. Am. Chem. Soc. 2008, 130, 11467−11476. (41) Liu, Z.; Chen, K.; Davis, C.; Sherlock, S.; Cao, Q.; Chen, X.; Dai, H. Drug Delivery with Carbon Nanotubes for In vivo Cancer Treatment. Cancer Res. 2008, 68, 6652−6660. (42) Wu, W.; Wieckowski, S.; Pastorin, G.; Benincasa, M.; Klumpp, C.; Briand, J.-P.; Gennaro, R.; Prato, M.; Bianco, A. Targeted Delivery of Amphotericin B to Cells by Using Functionalized Carbon Nanotubes. Angew. Chem., Int. Ed. 2005, 44, 6358−6362. (43) Samori, C.; Ali-Boucetta, H.; Sainz, R.; Guo, C.; Toma, F. M.; Fabbro, C.; da Ros, T.; Prato, M.; Kostarelos, K.; Bianco, A. Enhanced anticancer activity of multi-walled carbon nanotube-methotrexate conjugates using cleavable linkers. Chem. Commun. (Cambridge, U. K.) 2010, 46, 1494−1496. (44) Gupta, A. K.; Gupta, M. Synthesis and surface engineering of iron oxide nanoparticles for biomedical applications. Biomaterials 2005, 26, 3995−4021. (45) Laurent, S.; Forge, D.; Port, M.; Roch, A.; Robic, C.; Vander Elst, L.; Muller, R. N. Magnetic Iron Oxide Nanoparticles: Synthesis, Stabilization, Vectorization, Physicochemical Characterizations, and Biological Applications. Chem. Rev. (Washington, DC, U. S.) 2008, 108, 2064−2110. (46) Patra, C. R.; Bhattacharya, R.; Mukhopadhyay, D.; Mukherjee, P. Fabrication of gold nanoparticles for targeted therapy in pancreatic cancer. Adv. Drug Delivery Rev. 2010, 62, 346−361. (47) Mitchell, L. A.; Lauer, F. T.; Burchiel, S. W.; McDonald, J. D. Mechanisms for how inhaled multiwalled carbon nanotubes suppress systemic immune function in mice. Nat. Nanotechnol. 2009, 4, 451− 456. (48) Sood, A.; Salih, S.; Roh, D.; Lacharme Lora, L.; Parry, M.; Hardiman, B.; Keehan, R.; Grummer, R.; Winterhager, E.; Gokhale, P. J.; Andrews, P. W.; Abbott, C.; Forbes, K.; Westwood, M.; Aplin, J. D.; Ingham, E.; Papageorgiou, I.; Berry, M.; Liu, J.; Dick, A. D.; Garland, R. J.; Williams, N.; Singh, R.; Simon, A. K.; Lewis, M.; Ham, J.; Roger, L.; Baird, D. M.; Crompton, L. A.; Caldwell, M. A.; Swalwell, H.; Birch Machin, M.; Lopez Castejon, G.; Randall, A.; Lin, H.; Suleiman, M. S.; Evans, W. H.; Newson, R.; Case, C. P. Signalling of DNA damage and cytokines across cell barriers exposed to nanoparticles depends on barrier thickness. Nat. Nanotechnol. 2011, 6, 824−833. (49) Zolnik, B. S.; Gonzál ez-Fernán dez, Á .; Sadrieh, N.; Dobrovolskaia, M. A. Minireview: Nanoparticles and the Immune System. Endocrinology 2010, 151, 458−465. (50) Alkilany, A. M.; Nagaria, P. K.; Hexel, C. R.; Shaw, T. J.; Murphy, C. J.; Wyatt, M. D. Cellular Uptake and Cytotoxicity of Gold

Nanorods: Molecular Origin of Cytotoxicity and Surface Effects. Small 2009, 5, 701−708. (51) Hong, S.; Leroueil, P. R.; Janus, E. K.; Peters, J. L.; Kober, M.M.; Islam, M. T.; Orr, B. G.; Baker, J. R.; Banaszak Holl, M. M. Interaction of Polycationic Polymers with Supported Lipid Bilayers and Cells: Nanoscale Hole Formation and Enhanced Membrane Permeability. Bioconjugate Chem. 2006, 17, 728−734. (52) Kelly, C. V.; Leroueil, P. R.; Orr, B. G.; Banaszak Holl, M. M.; Andricioaei, I. Poly(amidoamine) Dendrimers on Lipid Bilayers II: Effects of Bilayer Phase and Dendrimer Termination. J. Phys. Chem. B 2008, 112, 9346−9353. (53) Mecke, A.; Majoros, I. J.; Patri, A. K.; Baker, J. R.; Banaszak Holl, M. M.; Orr, B. G. Lipid Bilayer Disruption by Polycationic Polymers: The Roles of Size and Chemical Functional Group. Langmuir 2005, 21, 10348−10354. (54) Thomas, T. P.; Majoros, I. J.; Kotlyar, A.; Kukowska-Latallo, J. F.; Bielinska, A.; Myc, A.; Baker, J. R., Jr. Targeting and Inhibition of Cell Growth by an Engineered Dendritic Nanodevice. J. Med. Chem. 2005, 48, 3729−3735. (55) Zhang, Y.; Bai, Y.; Jia, J.; Gao, N.; Li, Y.; Zhang, R.; Jiang, G.; Yan, B. Perturbation of physiological systems by nanoparticles. Chem. Soc. Rev. 2014, 43, 3762−3809. (56) Mu, Q.; Jiang, G.; Chen, L.; Zhou, H.; Fourches, D.; Tropsha, A.; Yan, B. Chemical Basis of Interactions Between Engineered Nanoparticles and Biological Systems. Chem. Rev. (Washington, DC, U. S.) 2014, 114, 7740−7781. (57) Leroueil, P. R.; Hong, S.; Mecke, A.; Baker, J. R.; Orr, B. G.; Banaszak Holl, M. M. Nanoparticle Interaction with Biological Membranes: Does Nanotechnology Present a Janus Face? Acc. Chem. Res. 2007, 40, 335−342. (58) Elsaesser, A.; Howard, C. V. Toxicology of nanoparticles. Adv. Drug Delivery Rev. 2012, 64, 129−137. (59) Love, S. A.; Maurer-Jones, M. A.; Thompson, J. W.; Lin, Y.-S.; Haynes, C. L. Assessing Nanoparticle Toxicity. Annu. Rev. Anal. Chem. 2012, 5, 181−205. (60) Peynshaert, K.; Manshian, B. B.; Joris, F.; Braeckmans, K.; De Smedt, S. C.; Demeester, J.; Soenen, S. J. Exploiting Intrinsic Nanoparticle Toxicity: The Pros and Cons of Nanoparticle-Induced Autophagy in Biomedical Research. Chem. Rev. (Washington, DC, U. S.) 2014, 114, 7581−7609. (61) Soenen, S. J.; Parak, W. J.; Rejman, J.; Manshian, B. (Intra)Cellular Stability of Inorganic Nanoparticles: Effects on Cytotoxicity, Particle Functionality, and Biomedical Applications. Chem. Rev. (Washington, DC, U. S.) 2015, 115, 2109−2135. (62) Khlebtsov, N.; Dykman, L. Biodistribution and toxicity of engineered gold nanoparticles: a review of in vitro and in vivo studies. Chem. Soc. Rev. 2011, 40, 1647−1671. (63) Li, Y.-F.; Chen, C. Fate and Toxicity of Metallic and MetalContaining Nanoparticles for Biomedical Applications. Small 2011, 7, 2965−2980. (64) Crampton, H.; Hollink, E.; Perez, L. M.; Simanek, E. E. A divergent route towards single-chemical entity triazine dendrimers with opportunities for structural diversity. New J. Chem. 2007, 31, 1283−1290. (65) Ihre, H.; Padilla De Jesús, O. L.; Fréchet, J. M. J. Fast and Convenient Divergent Synthesis of Aliphatic Ester Dendrimers by Anhydride Coupling. J. Am. Chem. Soc. 2001, 123, 5908−5917. (66) Daniel, M.-C.; Astruc, D. Gold Nanoparticles; Assembly, Supramolecular Chemistry, Quantum-Size-Related Properties, and Applications toward Biology, Catalysis, and Nanotechnology. Chem. Rev. (Washington, DC, U. S.) 2004, 104, 293−346. (67) Ghosh, S. K.; Pal, A.; Kundu, S.; Nath, S.; Pal, T. Fluorescence quenching of 1-methylaminopyrene near gold nanoparticles: size regime dependence of the small metallic particles. Chem. Phys. Lett. 2004, 395, 366−372. (68) Hobbs, S. K.; Monsky, W. L.; Yuan, F.; Roberts, W. G.; Griffith, L.; Torchilin, V. P.; Jain, R. K. Regulation of transport pathways in tumor vessels: Role of tumor type and microenvironment. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 4607−4612. 3420

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

(69) Maeda, H.; Matsumura, Y. Tumoritropic and lymphotropic principles of macromolecular drugs. Crit. Rev. Ther. Drug Carrier Syst. 1989, 6, 193−210. (70) Maeda, H. The enhanced permeability and retention (EPR) effect in tumor vasculature: the key role of tumor-selective macromolecular drug targeting. Adv. Enzyme Regul. 2001, 41, 189− 207. (71) Maeda, H.; Fang, J.; Inutsuka, T.; Kitamoto, Y. Vascular permeability enhancement in solid tumor: various factors, mechanisms involved and its implications. Int. Immunopharmacol. 2003, 3, 319− 328. (72) Hilgenbrink, A. R.; Low, P. S. Folate receptor-mediated drug targeting: From therapeutics to diagnostics. J. Pharm. Sci. 2005, 94, 2135−2146. (73) Lee, R. J.; Low, P. S. Folate-mediated tumor cell targeting of liposome-entrapped doxorubicin in vitro. Biochim. Biophys. Acta, Biomembr. 1995, 1233, 134−144. (74) Low, P. S.; Kularatne, S. A. Folate-targeted therapeutic and imaging agents for cancer. Curr. Opin. Chem. Biol. 2009, 13, 256−262. (75) Thomas, T. P.; Choi, S. K.; Li, M.-H.; Kotlyar, A.; Baker, J. R., Jr. Design of Riboflavin-Presenting PAMAM Dendrimers as a New Nanoplatform for Cancer-Targeted Delivery. Bioorg. Med. Chem. Lett. 2010, 20, 5191−5194. (76) Plantinga, A.; Witte, A.; Li, M.-H.; Harmon, A.; Choi, S. K.; Banaszak Holl, M. M.; Orr, B. G.; Baker, J. R., Jr.; Sinniah, K. Bioanalytical Screening of Riboflavin Antagonists for Targeted Drug DeliveryA Thermodynamic and Kinetic Study. ACS Med. Chem. Lett. 2011, 2, 363−367. (77) Montet, X.; Funovics, M.; Montet-Abou, K.; Weissleder, R.; Josephson, L. Multivalent Effects of RGD Peptides Obtained by Nanoparticle Display. J. Med. Chem. 2006, 49, 6087−6093. (78) Shukla, R.; Thomas, T. P.; Peters, J.; Kotlyar, A.; Myc, A.; Baker, J. R., Jr. Tumor angiogenic vasculature targeting with PAMAM dendrimer-RGD conjugates. Chem. Commun. (Cambridge, U. K.) 2005, 5739−5741. (79) Temming, K.; Lacombe, M.; Schaapveld, R. Q. J.; Orfi, L.; Kéri, G.; Poelstra, K.; Molema, G.; Kok, R. J. Rational Design of RGD− Albumin Conjugates for Targeted Delivery of the VEGF-R Kinase Inhibitor PTK787 to Angiogenic Endothelium. ChemMedChem 2006, 1, 1200−1203. (80) Chen, Y.; Foss, C. A.; Byun, Y.; Nimmagadda, S.; Pullambhatla, M.; Fox, J. J.; Castanares, M.; Lupold, S. E.; Babich, J. W.; Mease, R. C.; Pomper, M. G. Radiohalogenated Prostate-Specific Membrane Antigen (PSMA)-Based Ureas as Imaging Agents for Prostate Cancer. J. Med. Chem. 2008, 51, 7933−7943. (81) Shukla, R.; Thomas, T. P.; Desai, A. M.; Kotlyar, A.; Park, S. J.; Baker, J. R., Jr. HER2 specific delivery of methotrexate by dendrimer conjugated anti-HER2 mAb. Nanotechnology 2008, 19, 295102. (82) Thomas, T. P.; Shukla, R.; Kotlyar, A.; Liang, B.; Ye, J. Y.; Norris, T. B.; Baker, J. R., Jr. Dendrimer-Epidermal Growth Factor Conjugate Displays Superagonist Activity. Biomacromolecules 2008, 9, 603−609. (83) Raha, S.; Paunesku, T.; Woloschak, G. Peptide-mediated cancer targeting of nanoconjugates. Wiley Interdiscip. Rev.: Nanomed. Nanobiotechnol. 2010, 3, 269−281. (84) Thomas, T. P.; Shukla, R.; Kotlyar, A.; Kukowska-Latallo, J.; Baker, J. R., Jr. Dendrimer-based tumor cell targeting of fibroblast growth factor-1. Bioorg. Med. Chem. Lett. 2010, 20, 700−703. (85) Vinogradov, S. V.; Batrakova, E. V.; Kabanov, A. V. Nanogels for Oligonucleotide Delivery to the Brain. Bioconjugate Chem. 2003, 15, 50−60. (86) Banquy, X.; Leclair, G.; Rabanel, J.-M.; Argaw, A.; Bouchard, J.F.; Hildgen, P.; Giasson, S. Selectins Ligand Decorated Drug Carriers for Activated Endothelial Cell Targeting. Bioconjugate Chem. 2008, 19, 2030−2039. (87) Bhaskar, V.; Law, D. A.; Ibsen, E.; Breinberg, D.; Cass, K. M.; DuBridge, R. B.; Evangelista, F.; Henshall, S. M.; Hevezi, P.; Miller, J. C.; Pong, M.; Powers, R.; Senter, P.; Stockett, D.; Sutherland, R. L.; von Freeden-Jeffry, U.; Willhite, D.; Murray, R.; Afar, D. E. H.;

Ramakrishnan, V. E-Selectin Up-Regulation Allows for Targeted Drug Delivery in Prostate Cancer. Cancer Res. 2003, 63, 6387−6394. (88) Jubeli, E.; Moine, L.; Vergnaud-Gauduchon, J.; Barratt, G. Eselectin as a target for drug delivery and molecular imaging. J. Controlled Release 2012, 158, 194−206. (89) Qian, Z. M.; Li, H.; Sun, H.; Ho, K. Targeted Drug Delivery via the Transferrin Receptor-Mediated Endocytosis Pathway. Pharmacol. Rev. 2002, 54, 561−587. (90) Hong, S.; Leroueil, P. R.; Majoros, I. J.; Orr, B. G.; Baker, J. R., Jr.; Banaszak Holl, M. M. The Binding Avidity of a Nanoparticle-Based Multivalent Targeted Drug Delivery Platform. Chem. Biol. (Oxford, U. K.) 2007, 14, 107−115. (91) Mammen, M.; Choi, S. K.; Whitesides, G. M. Polyvalent Interactions in Biological Systems: Implications for Design and Use of Multivalent Ligands and Inhibitors. Angew. Chem., Int. Ed. 1998, 37, 2754−2794. (92) Lee, Y. C.; Lee, R. T. Carbohydrate-Protein Interactions: Basis of Glycobiology. Acc. Chem. Res. 1995, 28, 321−327. (93) Kiessling, L. L.; Gestwicki, J. E.; Strong, L. E. Synthetic Multivalent Ligands in the Exploration of Cell-Surface Interactions. Curr. Opin. Chem. Biol. 2000, 4, 696−703. (94) Roy, R. Syntheses and Some Applications of Chemically Defined Multivalent Glycoconjugates. Curr. Opin. Struct. Biol. 1996, 6, 692−702. (95) Dendrimer-Based Nanomedicine; Majoros, I., Baker Jr., J., Eds.; Pan Stanford: Hackensack, NJ, 2008. (96) Choi, Y.; Thomas, T.; Kotlyar, A.; Islam, M. T.; Baker, J. R., Jr. Synthesis and Functional Evaluation of DNA-Assembled Polyamidoamine Dendrimer Clusters for Cancer Cell-Specific Targeting. Chem. Biol. (Oxford, U. K.) 2005, 12, 35−43. (97) Lu, Y.; Low, P. S. Folate-mediated delivery of macromolecular anticancer therapeutic agents. Adv. Drug Delivery Rev. 2002, 54, 675− 693. (98) Thomas, T. P.; Huang, B.; Choi, S. K.; Silpe, J. E.; Kotlyar, A.; Desai, A. M.; Gam, J.; Joice, M.; B, J. R., Jr. Polyvalent PAMAMMethotrexate Dendrimer as a Folate Receptor-Targeted Therapeutic. Mol. Pharmaceutics 2012, 9, 2669−2676. (99) Kamen, B. A.; Capdevila, A. Receptor-mediated folate accumulation is regulated by the cellular folate content. Proc. Natl. Acad. Sci. U. S. A. 1986, 83, 5983−5987. (100) Nandini-Kishore, S. G.; Frazier, W. A. [3H]Methotrexate as a ligand for the folate receptor of Dictyostelium discoideum. Proc. Natl. Acad. Sci. U. S. A. 1981, 78, 7299−7303. (101) Rijnboutt, S.; Jansen, G.; Posthuma, G.; Hynes, J. B.; Schornagel, J. H.; Strous, G. J. Endocytosis of GPI-linked Membrane Folate Receptor-α. J. Cell Biol. 1996, 132, 35−47. (102) Henderson, E. A.; Bavetsias, V.; Theti, D. S.; Wilson, S. C.; Clauss, R.; Jackman, A. L. Targeting the α-folate receptor with cyclopenta[g]quinazoline-based inhibitors of thymidylate synthase. Bioorg. Med. Chem. 2006, 14, 5020−5042. (103) Heijden, J. W. V. D.; Oerlemans, R.; Dijkmans, B. A. C.; Qi, H.; Laken, C. J. V. D.; Lems, W. F.; Jackman, A. L.; Kraan, M. C.; Tak, P. P.; Ratnam, M.; Jansen, G. Folate receptor beta as a potential delivery route for novel folate antagonists to macrophages in the synovial tissue of rheumatoid arthritis patients. Arthritis Rheum. 2009, 60, 12−21. (104) Thomas, T. P.; Goonewardena, S. N.; Majoros, I. J.; Kotlyar, A.; Cao, Z.; Leroueil, P. R.; Baker, J. R. Folate-Targeted Nanoparticles Show Efficacy in the Treatment of Inflammatory Arthritis. Arthritis Rheum. 2011, 63, 2671−2680. (105) Phelps, M. A.; Foraker, A. B.; Gao, W.; Dalton, J. T.; Swaan, P. W. A Novel Rhodamine-Riboflavin Conjugate Probe Exhibits Distinct Fluorescence Resonance Energy Transfer that Enables Riboflavin Trafficking and Subcellular Localization Studies. Mol. Pharmaceutics 2004, 1, 257−266. (106) Zempleni, J. Uptake, localization, and noncarboxylase roles of biotin. Annu. Rev. Nutr. 2005, 25, 175−196. (107) Yellepeddi, V. K.; Kumar, A.; Palakurthi, S. Biotinylated Poly(amido)amine (PAMAM) Dendrimers as Carriers for Drug 3421

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Delivery to Ovarian Cancer Cells In Vitro. Anticancer Res. 2009, 29, 2933−2943. (108) Heo, D. N.; Yang, D. H.; Moon, H.-J.; Lee, J. B.; Bae, M. S.; Lee, S. C.; Lee, W. J.; Sun, I.-C.; Kwon, I. K. Gold nanoparticles surface-functionalized with paclitaxel drug and biotin receptor as theranostic agents for cancer therapy. Biomaterials 2012, 33, 856−866. (109) Wu, G.; Barth, R. F.; Yang, W.; Kawabata, S.; Zhang, L.; GreenChurch, K. Targeted delivery of methotrexate to epidermal growth factor receptor-positive brain tumors by means of cetuximab (IMCC225) dendrimer bioconjugates. Mol. Cancer Ther. 2006, 5, 52−59. (110) Ö zcan, F.; Klein, P.; Lemmon, M. A.; Lax, I.; Schlessinger, J. On the nature of low- and high-affinity EGF receptors on living cells. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 5735−5740. (111) Li, Z.; Zhao, R.; Wu, X.; Sun, Y.; Yao, M.; Li, J.; Xu, Y.; Gu, J. Identification and characterization of a novel peptide ligand of epidermal growth factor receptor for targeted delivery of therapeutics. FASEB J. 2005, 19, 1978−1985. (112) Shukla, R.; Thomas, T. P.; Peters, J. L.; Desai, A. M.; Kukowska-Latallo, J.; Patri, A. K.; Kotlyar, A.; Baker, J. R. HER2 Specific Tumor Targeting with Dendrimer Conjugated Anti-HER2 mAb. Bioconjugate Chem. 2006, 17, 1109−1115. (113) Kim, D.; Yan, Y.; Valencia, C. A.; Liu, R. Heptameric Targeting Ligands against EGFR and HER2 with High Stability and Avidity. PLoS One 2012, 7, e43077. (114) Haugsten, E. M.; Wiedlocha, A.; Olsnes, S.; Wesche, J. Roles of Fibroblast Growth Factor Receptors in Carcinogenesis. Mol. Cancer Res. 2010, 8, 1439−1452. (115) Neufeld, G.; Gospodarowicz, D. The identification and partial characterization of the fibroblast growth factor receptor of baby hamster kidney cells. J. Biol. Chem. 1985, 260, 13860−13868. (116) Burkhart, D. J.; Kalet, B. T.; Coleman, M. P.; Post, G. C.; Koch, T. H. Doxorubicin-formaldehyde conjugates targeting αvβ3 integrin. Mol. Cancer Ther. 2004, 3, 1593−1604. (117) Khalili, P.; Arakelian, A.; Chen, G.; Plunkett, M. L.; Beck, I.; Parry, G. C.; Donate, F.; Shaw, D. E.; Mazar, A. P.; Rabbani, S. A. A non-RGD-based integrin binding peptide (ATN-161) blocks breast cancer growth and metastasis in vivo. Mol. Cancer Ther. 2006, 5, 2271−2280. (118) Haubner, R.; Gratias, R.; Diefenbach, B.; Goodman, S. L.; Jonczyk, A.; Kessler, H. Structural and Functional Aspects of RGDContaining Cyclic Pentapeptides as Highly Potent and Selective Integrin αvβ3 Antagonists. J. Am. Chem. Soc. 1996, 118, 7461−7472. (119) Xie, J.; Chen, K.; Lee, H.-Y.; Xu, C.; Hsu, A. R.; Peng, S.; Chen, X.; Sun, S. Ultrasmall c(RGDyK)-Coated Fe3O4 Nanoparticles and Their Specific Targeting to Integrin αvβ3-Rich Tumor Cells. J. Am. Chem. Soc. 2008, 130, 7542−7543. (120) Xiong, J.-P.; Stehle, T.; Zhang, R.; Joachimiak, A.; Frech, M.; Goodman, S. L.; Arnaout, M. A. Crystal Structure of the Extracellular Segment of Integrin alpha Vbeta 3 in Complex with an Arg-Gly-Asp Ligand. Science 2002, 296, 151−155. (121) Cressman, S.; Sun, Y.; Maxwell, E. J.; Fang, N.; Chen, D. Y.; Cullis, P. Binding and Uptake of RGD-Containing Ligands to Cellular αvβ3 Integrins. Int. J. Pept. Res. Ther. 2009, 15, 49−59. (122) Davis, M. I.; Bennett, M. J.; Thomas, L. M.; Bjorkman, P. J. Crystal structure of prostate-specific membrane antigen, a tumor marker and peptidase. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 5981− 5986. (123) Hrkach, J.; Von Hoff, D.; Ali, M. M.; Andrianova, E.; Auer, J.; Campbell, T.; De Witt, D.; Figa, M.; Figueiredo, M.; Horhota, A.; Low, S.; McDonnell, K.; Peeke, E.; Retnarajan, B.; Sabnis, A.; Schnipper, E.; Song, J. J.; Song, Y. H.; Summa, J.; Tompsett, D.; Troiano, G.; Van Geen Hoven, T.; Wright, J.; LoRusso, P.; Kantoff, P. W.; Bander, N. H.; Sweeney, C.; Farokhzad, O. C.; Langer, R.; Zale, S. Preclinical Development and Clinical Translation of a PSMA-Targeted Docetaxel Nanoparticle with a Differentiated Pharmacological Profile. Sci. Transl. Med. 2012, 4, 128−139. (124) Wu, L. Y.; Do, J. C.; Kazak, M.; Page, H.; Toriyabe, Y.; Anderson, M. O.; Berkman, C. E. Phosphoramidate derivatives of

hydroxysteroids as inhibitors of prostate-specific membrane antigen. Bioorg. Med. Chem. Lett. 2008, 18, 281−284. (125) Thomas, T. P.; Patri, A. K.; Myc, A.; Myaing, M. T.; Ye, J. Y.; Norris, T. B.; Baker, J. R. In Vitro Targeting of Synthesized AntibodyConjugated Dendrimer Nanoparticles. Biomacromolecules 2004, 5, 2269−2274. (126) Liu, K.; Xu, L.; Szalkowski, D.; Li, Z.; Ding, V.; Kwei, G.; Huskey, S.; Moller, D. E.; Heck, J. V.; Zhang, B. B.; Jones, A. B. Discovery of a Potent, Highly Selective, and Orally Efficacious SmallMolecule Activator of the Insulin Receptor. J. Med. Chem. 2000, 43, 3487−3494. (127) Zhang, B.; Salituro, G.; Szalkowski, D.; Li, Z.; Zhang, Y.; Royo, I.; Vilella, D.; Díez, M. T.; Pelaez, F.; Ruby, C.; Kendall, R. L.; Mao, X.; Griffin, P.; Calaycay, J.; Zierath, J. R.; Heck, J. V.; Smith, R. G.; Moller, D. E. Discovery of a Small Molecule Insulin Mimetic with Antidiabetic Activity in Mice. Science 1999, 284, 974−977. (128) Haluska, P.; Carboni, J. M.; Loegering, D. A.; Lee, F. Y.; Wittman, M.; Saulnier, M. G.; Frennesson, D. B.; Kalli, K. R.; Conover, C. A.; Attar, R. M.; Kaufmann, S. H.; Gottardis, M.; Erlichman, C. In vitro and In vivo Antitumor Effects of the Dual Insulin-Like Growth Factor-I/Insulin Receptor Inhibitor, BMS-554417. Cancer Res. 2006, 66, 362−371. (129) Pollak, M. Insulin and insulin-like growth factor signalling in neoplasia. Nat. Rev. Cancer 2008, 8, 915−928. (130) Venkateswaran, V.; Haddad, A. Q.; Fleshner, N. E.; Fan, R.; Sugar, L. M.; Nam, R.; Klotz, L. H.; Pollak, M. Association of DietInduced Hyperinsulinemia With Accelerated Growth of Prostate Cancer (LNCaP) Xenografts. J. Natl. Cancer Inst. 2007, 99, 1793− 1800. (131) Maturo, J.; Kulakowski, E. C. Taurine binding to the purified insulin receptor. Biochem. Pharmacol. (Amsterdam, Neth.) 1988, 37, 3755−3760. (132) Yang, P.-H.; Sun, X.; Chiu, J.-F.; Sun, H.; He, Q.-Y. Transferrin-Mediated Gold Nanoparticle Cellular Uptake. Bioconjugate Chem. 2005, 16, 494−496. (133) Hogemann-Savellano, D.; Bosy, E.; Blondet, C.; Sato, F.; Abe, T.; Josephson, L.; Weissleder, R.; Gaudet, J.; Sgroi, D.; Petersy, P. J.; Basilion, J. P. The Transferrin Receptor: A Potential Molecular Imaging Marker for Human Cancer. Neoplasia 2003, 5, 495−506. (134) Qian, Z. M.; To, Y.; Tang, P. L.; Feng, Y. M. Transferrin receptors on the plasma membrane of cultured rat astrocytes. Exp. Brain Res. 1999, 129, 473−476. (135) Boghaert, E. R.; Sridharan, L.; Armellino, D. C.; Khandke, K. M.; DiJoseph, J. F.; Kunz, A.; Dougher, M. M.; Jiang, F.; Kalyandrug, L. B.; Hamann, P. R.; Frost, P.; Damle, N. K. Antibody-Targeted Chemotherapy with the Calicheamicin Conjugate hu3S193-N-Acetyl γ Calicheamicin Dimethyl Hydrazide Targets Lewisy and Eliminates Lewisy-Positive Human Carcinoma Cells and Xenografts. Clin. Cancer Res. 2004, 10, 4538−4549. (136) Ndungu, J. M.; Lu, Y. J.; Zhu, S.; Yang, C.; Wang, X.; Chen, G.; Shin, D. M.; Snyder, J. P.; Shoji, M.; Sun, A. Targeted Delivery of Paclitaxel to Tumor Cells: Synthesis and in Vitro Evaluation. J. Med. Chem. 2010, 53, 3127−3132. (137) Hu, J.; Xu, T.; Cheng, Y. NMR Insights into Dendrimer-Based Host−Guest Systems. Chem. Rev. (Washington, DC, U. S.) 2012, 112, 3856−3891. (138) Newkome, G. R.; He, E.; Moorefield, C. N. Suprasupermolecules with Novel Properties: Metallodendrimers. Chem. Rev. (Washington, DC, U. S.) 1999, 99, 1689−1746. (139) Zeng, F.; Zimmerman, S. C. Dendrimers in Supramolecular Chemistry: From Molecular Recognition to Self-Assembly. Chem. Rev. (Washington, DC, U. S.) 1997, 97, 1681−1712. (140) Brunsveld, L.; Folmer, B. J. B.; Meijer, E. W.; Sijbesma, R. P. Supramolecular Polymers. Chem. Rev. (Washington, DC, U. S.) 2001, 101, 4071−4098. (141) Al-Jamal, W. T.; Kostarelos, K. Liposomes: From a Clinically Established Drug Delivery System to a Nanoparticle Platform for Theranostic Nanomedicine. Acc. Chem. Res. 2011, 44, 1094−1104. 3422

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

(142) Torchilin, V. P. Recent advances with liposomes as pharmaceutical carriers. Nat. Rev. Drug Discovery 2005, 4, 145−160. (143) Lee, E. S.; Gao, Z.; Bae, Y. H. Recent progress in tumor pH targeting nanotechnology. J. Controlled Release 2008, 132, 164−170. (144) Gullotti, E.; Yeo, Y. Extracellularly Activated Nanocarriers: A New Paradigm of Tumor Targeted Drug Delivery. Mol. Pharmaceutics 2009, 6, 1041−1051. (145) Chen, H. H. W.; Song, I.-S.; Hossain, A.; Choi, M.-K.; Yamane, Y.; Liang, Z. D.; Lu, J.; Wu, L. Y.-H.; Siddik, Z. H.; Klomp, L. W. J.; Savaraj, N.; Kuo, M. T. Elevated Glutathione Levels Confer Cellular Sensitization to Cisplatin Toxicity by Up-Regulation of Copper Transporter hCtr1. Mol. Pharmacol. 2008, 74, 697−704. (146) Doronina, S. O.; Toki, B. E.; Torgov, M. Y.; Mendelsohn, B. A.; Cerveny, C. G.; Chace, D. F.; DeBlanc, R. L.; Gearing, R. P.; Bovee, T. D.; Siegall, C. B.; Francisco, J. A.; Wahl, A. F.; Meyer, D. L.; Senter, P. D. Development of potent monoclonal antibody auristatin conjugates for cancer therapy. Nat. Biotechnol. 2003, 21, 778−784. (147) Georgelin, T.; Bombard, S.; Siaugue, J.-M.; Cabuil, V. Nanoparticle-Mediated Delivery of Bleomycin. Angew. Chem., Int. Ed. 2010, 49, 8897−8901. (148) Harada, M.; Sakakibara, H.; Yano, T.; Suzuki, T.; Okuno, S. Determinants for the drug release from T-0128, camptothecin analogue-carboxymethyl dextran conjugate. J. Controlled Release 2000, 69, 399−412. (149) Vijayalakshmi, N.; Ray, A.; Malugin, A.; Ghandehari, H. Carboxyl-Terminated PAMAM-SN38 Conjugates: Synthesis, Characterization, and in Vitro Evaluation. Bioconjugate Chem. 2010, 21, 1804−1810. (150) Hamann, P. R.; Hinman, L. M.; Beyer, C. F.; Lindh, D.; Upeslacis, J.; Flowers, D. A.; Bernstein, I. An Anti-CD33 Antibody− Calicheamicin Conjugate for Treatment of Acute Myeloid Leukemia. Choice of Linker. Bioconjugate Chem. 2001, 13, 40−46. (151) Bhirde, A. A.; Patel, V.; Gavard, J.; Zhang, G.; Sousa, A. A.; Masedunskas, A.; Leapman, R. D.; Weigert, R.; Gutkind, J. S.; Rusling, J. F. Targeted Killing of Cancer Cells in Vivo and in Vitro with EGFDirected Carbon Nanotube-Based Drug Delivery. ACS Nano 2009, 3, 307−316. (152) Aryal, S.; Hu, C.-M. J.; Zhang, L. Polymer−Cisplatin Conjugate Nanoparticles for Acid-Responsive Drug Delivery. ACS Nano 2010, 4, 251−258. (153) Johansson, E. M. V.; Dubois, J.; Darbre, T.; Reymond, J.-L. Glycopeptide dendrimer colchicine conjugates targeting cancer cells. Bioorg. Med. Chem. 2010, 18, 6589−6597. (154) Ř íhová, B.; Strohalm, J.; Prausová, J.; Kubácǩ ová, K.; Jelínková, M.; Rozprimová, L.; Šírová, M.; Plocová, D.; Etrych, T.; Šubr, V.; Mrkvan, T.; Kovár,̌ M.; Ulbrich, K. Cytostatic and immunomobilizing activities of polymer-bound drugs: experimental and first clinical data. J. Controlled Release 2003, 91, 1−16. (155) de Groot, F. M. H.; Broxterman, H. J.; Adams, H. P. H. M.; van Vliet, A.; Tesser, G. I.; Elderkamp, Y. W.; Schraa, A. J.; Jan Kok, R.; Molema, G.; Pinedo, H. M.; Scheeren, H. W. Design, Synthesis, and Biological Evaluation of a Dual Tumor-specific Motive Containing Integrin-targeted Plasmin-cleavable Doxorubicin Prodrug. Mol. Cancer Ther. 2002, 1, 901−911. (156) Etrych, T.; Mrkvan, T.; Ř íhová, B.; Ulbrich, K. Star-shaped immunoglobulin-containing HPMA-based conjugates with doxorubicin for cancer therapy. J. Controlled Release 2007, 122, 31−38. (157) Choi, S. K.; Thomas, T.; Li, M.; Kotlyar, A.; Desai, A.; Baker, J. R., Jr. Light-Controlled Release of Caged Doxorubicin from Folate Receptor-Targeting PAMAM Dendrimer Nanoconjugate. Chem. Commun. (Cambridge, U. K.) 2010, 46, 2632−2634. (158) Kaminskas, L. M.; Kelly, B. D.; McLeod, V. M.; Sberna, G.; Boyd, B. J.; Owen, D. J.; Porter, C. J. H. Capping Methotrexate αCarboxyl Groups Enhances Systemic Exposure and Retains the Cytotoxicity of Drug Conjugated PEGylated Polylysine Dendrimers. Mol. Pharmaceutics 2011, 8, 338−349. (159) Kaminskas, L. M.; Kelly, B. D.; McLeod, V. M.; Boyd, B. J.; Krippner, G. Y.; Williams, E. D.; Porter, C. J. H. Pharmacokinetics and

Tumor Disposition of PEGylated, Methotrexate Conjugated Poly-llysine Dendrimers. Mol. Pharmaceutics 2009, 6, 1190−1204. (160) Choi, S. K.; Thomas, T. P.; Li, M.-H.; Desai, A.; Kotlyar, A.; Baker, J. R. Photochemical release of methotrexate from folate receptor-targeting PAMAM dendrimer nanoconjugate. Photochem. Photobiol. Sci. 2012, 11, 653−660. (161) Choi, S. K.; Verma, M.; Silpe, J.; Moody, R. E.; Tang, K.; Hanson, J. J.; Baker, J. R., Jr. A photochemical approach for controlled drug release in targeted drug delivery. Bioorg. Med. Chem. 2012, 20, 1281−1290. (162) Manabe, Y.; Tsubota, T.; Haruta, Y.; Kataoka, K.; Okazaki, M.; Haisa, S.; Nakamura, K.; Kimura, I. Production of a monoclonal antibody-mitomycin C conjugate, utilizing dextran T-40, and its biological activity. Biochem. Pharmacol. 1985, 34, 289−291. (163) Widdison, W. C.; Wilhelm, S. D.; Cavanagh, E. E.; Whiteman, K. R.; Leece, B. A.; Kovtun, Y.; Goldmacher, V. S.; Xie, H.; Steeves, R. M.; Lutz, R. J.; Zhao, R.; Wang, L.; Blättler, W. A.; Chari, R. V. J. Semisynthetic Maytansine Analogues for the Targeted Treatment of Cancer. J. Med. Chem. 2006, 49, 4392−4408. (164) Kellogg, B. A.; Garrett, L.; Kovtun, Y.; Lai, K. C.; Leece, B.; Miller, M.; Payne, G.; Steeves, R.; Whiteman, K. R.; Widdison, W.; Xie, H.; Singh, R.; Chari, R. V. J.; Lambert, J. M.; Lutz, R. J. DisulfideLinked Antibody−Maytansinoid Conjugates: Optimization of In Vivo Activity by Varying the Steric Hindrance at Carbon Atoms Adjacent to the Disulfide Linkage. Bioconjugate Chem. 2011, 22, 717−727. (165) Ajaj, K. A.; Biniossek, M. L.; Kratz, F. Development of ProteinBinding Bifunctional Linkers for a New Generation of Dual-Acting Prodrugs. Bioconjugate Chem. 2009, 20, 390−396. (166) Majoros, I. J.; Myc, A.; Thomas, T.; Mehta, C. B.; Baker, J. R. PAMAM Dendrimer-Based Multifunctional Conjugate for Cancer Therapy: Synthesis, Characterization, and Functionality. Biomacromolecules 2006, 7, 572−579. (167) Paciotti, G. F.; Kingston, D. G. I.; Tamarkin, L. Colloidal gold nanoparticles: A novel nanoparticle platform for developing multifunctional tumor-targeted drug delivery vectors. Drug Dev. Res. 2006, 67, 47−54. (168) Zheng, Y.; Fu, F.; Zhang, M.; Shen, M.; Zhu, M.; Shi, X. Multifunctional dendrimers modified with alpha-tocopheryl succinate for targeted cancer therapy. MedChemComm 2014, 5, 879−885. (169) Zhu, J.; Zheng, L.; Wen, S.; Tang, Y.; Shen, M.; Zhang, G.; Shi, X. Targeted cancer theranostics using alpha-tocopheryl succinateconjugated multifunctional dendrimer-entrapped gold nanoparticles. Biomaterials 2014, 35, 7635−7646. (170) Dreaden, E. C.; Mwakwari, S. C.; Sodji, Q. H.; Oyelere, A. K.; El-Sayed, M. A. Tamoxifen−Poly(ethylene glycol)−Thiol Gold Nanoparticle Conjugates: Enhanced Potency and Selective Delivery for Breast Cancer Treatment. Bioconjugate Chem. 2009, 20, 2247− 2253. (171) Gu, H.; Ho, P. L.; Tong, E.; Wang, L.; Xu, B. Presenting Vancomycin on Nanoparticles to Enhance Antimicrobial Activities. Nano Lett. 2003, 3, 1261−1263. (172) Kell, A. J.; Stewart, G.; Ryan, S.; Peytavi, R.; Boissinot, M.; Huletsky, A.; Bergeron, M. G.; Simard, B. Vancomycin-Modified Nanoparticles for Efficient Targeting and Preconcentration of GramPositive and Gram-Negative Bacteria. ACS Nano 2008, 2, 1777−1788. (173) Metallo, S. J.; Kane, R. S.; Holmlin, R. E.; Whitesides, G. M. Using Bifunctional Polymers Presenting Vancomycin and Fluorescein Groups To Direct Anti-Fluorescein Antibodies to Self-Assembled Monolayers Presenting d-Alanine-d-Alanine Groups. J. Am. Chem. Soc. 2003, 125, 4534−4540. (174) Choi, S. K.; Myc, A.; Silpe, J. E.; Sumit, M.; Wong, P. T.; McCarthy, K.; Desai, A. M.; Thomas, T. P.; Kotlyar, A.; Banaszak Holl, M. M.; Orr, B. G.; Baker, J. R. Dendrimer-Based Multivalent Vancomycin Nanoplatform for Targeting the Drug-Resistant Bacterial Surface. ACS Nano 2013, 7, 214−228. (175) Bosnjakovic, A.; Mishra, M. K.; Ren, W.; Kurtoglu, Y. E.; Shi, T.; Fan, D.; Kannan, R. M. Poly(amidoamine) dendrimererythromycin conjugates for drug delivery to macrophages involved 3423

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

in periprosthetic inflammation. Nanomedicine (N. Y., NY, U. S.) 2011, 7, 284−294. (176) Gunaseelan, S.; Debrah, O.; Wan, L.; Leibowitz, M. J.; Rabson, A. B.; Stein, S.; Sinko, P. J. Synthesis of Poly(ethylene glycol)-Based Saquinavir Prodrug Conjugates and Assessment of Release and AntiHIV-1 Bioactivity Using a Novel Protease Inhibition Assay. Bioconjugate Chem. 2004, 15, 1322−1333. (177) Levy, G. A.; Adamson, G.; Phillips, M. J.; Scrocchi, L. A.; Fung, L.; Biessels, P.; Ng, N. F.; Ghanekar, A.; Rowe, A.; Ma, M. X.; Levy, A.; Koscik, C.; He, W.; Gorczynski, R.; Brookes, S.; Woods, C.; McGilvray, I. D.; Bell, D. Targeted delivery of ribavirin improves outcome of murine viral fulminant hepatitis via enhanced anti-viral activity. Hepatology (Hoboken, NJ, U. S.) 2006, 43, 581−591. (178) Di Stefano, G.; Lanza, M.; Busi, C.; Barbieri, L.; Fiume, L. Conjugates of Nucleoside Analogs with Lactosaminated Human Albumin to Selectively Increase the Drug Levels in Liver Blood: Requirements for a Regional Chemotherapy. J. Pharmacol. Exp. Ther. 2002, 301, 638−642. (179) Mahajan, S. D.; Law, W.-C.; Aalinkeel, R.; Reynolds, J.; Nair, B. B.; Yong, K.-T.; Roy, I.; Prasad, P. N.; Schwartz, S. A. In Methods in Enzymology; Nejat, D., Ed.; Academic Press: Waltham, MA, 2012; Vol. 509. (180) Bowman, M.-C.; Ballard, T. E.; Ackerson, C. J.; Feldheim, D. L.; Margolis, D. M.; Melander, C. Inhibition of HIV Fusion with Multivalent Gold Nanoparticles. J. Am. Chem. Soc. 2008, 130, 6896− 6897. (181) Azzi, J.; Tang, L.; Moore, R.; Tong, R.; El Haddad, N.; Akiyoshi, T.; Mfarrej, B.; Yang, S.; Jurewicz, M.; Ichimura, T.; Lindeman, N.; Cheng, J.; Abdi, R. Polylactide-cyclosporin A nanoparticles for targeted immunosuppression. FASEB J. 2010, 24, 3927−3938. (182) Tian, J.; Stella, V. J. Degradation of paclitaxel and related compounds in aqueous solutions III: Degradation under acidic pH conditions and overall kinetics. J. Pharm. Sci. 2010, 99, 1288−1298. (183) Testa, B.; Mayer, J. M. Hydrolysis in Drug and Prodrug Metabolism: Chemistry, Biochemistry, and Enzymology; Wiley-VCH: Weinheim, Germany, 2003. (184) Jencks, W. P.; Carriuolo, J. General Base Catalysis of Ester Hydrolysis. J. Am. Chem. Soc. 1961, 83, 1743−1750. (185) Mortellaro, M. A.; Bleisch, T. J.; Duerr, B. F.; Kang, M. S.; Huang, H.; Czarnik, A. W. Metal Ion-Catalyzed Hydrolysis of Acrylate Esters and Amides by Way of Their Conjugate Addition Adducts. J. Org. Chem. 1995, 60, 7238−7246. (186) Salvi, A.; Carrupt, P.-A.; Mayer, J. M.; Testa, B. Esterase-like Activity of Human Serum Albumin Toward Prodrug Esters of Nicotinic Acid. Drug Metab. Dispos. 1997, 25, 395−398. (187) Fukami, T.; Yokoi, T. The Emerging Role of Human Esterases. Drug Metab. Pharmacokinet. 2012, 27, 466−477. (188) Runquist, E. A.; Havel, R. J. Acid hydrolases in early and late endosome fractions from rat liver. J. Biol. Chem. 1991, 266, 22557− 22563. (189) Rebilly, J.-N.; Colasson, B.; Bistri, O.; Over, D.; Reinaud, O. Biomimetic cavity-based metal complexes. Chem. Soc. Rev. 2015, 44, 467−489. (190) Zhang, B.; Breslow, R. Ester Hydrolysis by a Catalytic Cyclodextrin Dimer Enzyme Mimic with a Metallobipyridyl Linking Group. J. Am. Chem. Soc. 1997, 119, 1676−1681. (191) Bareford, L. M.; Swaan, P. W. Endocytic mechanisms for targeted drug delivery. Adv. Drug Delivery Rev. 2007, 59, 748−758. (192) Chan, P.; Lovrić, J.; Warwicker, J. Subcellular pH and predicted pH-dependent features of proteins. PROTEOMICS 2006, 6, 3494− 3501. (193) Geisow, M. J. Fluorescein conjugates as indicators of subcellular pH: A critical evaluation. Exp. Cell Res. 1984, 150, 29−35. (194) Geisow, M. J.; Evans, W. H. pH in the endosome: Measurements during pinocytosis and receptor-mediated endocytosis. Exp. Cell Res. 1984, 150, 36−46. (195) Llopis, J.; McCaffery, J. M.; Miyawaki, A.; Farquhar, M. G.; Tsien, R. Y. Measurement of cytosolic, mitochondrial, and Golgi pH in

single living cells with green fluorescent proteins. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 6803−6808. (196) Paradise, R. K.; Lauffenburger, D. A.; Van Vliet, K. J. Acidic Extracellular pH Promotes Activation of Integrin αvβ3. PLoS One 2011, 6, e15746. (197) Curran, R. E.; Claxton, C. R. J.; Hutchison, L.; Harradine, P. J.; Martin, I. J.; Littlewood, P. Control and Measurement of Plasma pH in Equilibrium Dialysis: Influence on Drug Plasma Protein Binding. Drug Metab. Dispos. 2011, 39, 551−557. (198) Ma, C.; Subramani, S. Peroxisome matrix and membrane protein biogenesis. IUBMB Life 2009, 61, 713−722. (199) Yang, Y.-h.; Aloysius, H.; Inoyama, D.; Chen, Y.; Hu, L.-q. Enzyme-mediated hydrolytic activation of prodrugs. Acta Pharm. Sin. B 2011, 1, 143−159. (200) Rooseboom, M.; Commandeur, J. N. M.; Vermeulen, N. P. E. Enzyme-Catalyzed Activation of Anticancer Prodrugs. Pharmacol. Rev. 2004, 56, 53−102. (201) Chen, K.; Wang, K.; Kirichian, A. M.; Al Aowad, A. F.; Iyer, L. K.; Adelstein, S. J.; Kassis, A. I. In silico design, synthesis, and biological evaluation of radioiodinated quinazolinone derivatives for alkaline phosphatase-mediated cancer diagnosis and therapy. Mol. Cancer Ther. 2006, 5, 3001−3013. (202) Sanghani, S. P.; Quinney, S. K.; Fredenburg, T. B.; Sun, Z.; Davis, W. I.; Murry, D. J.; Cummings, O. W.; Seitz, D. E.; Bosron, W. F. Carboxylesterases Expressed in Human Colon Tumor Tissue and Their Role in CPT-11 Hydrolysis. Clin. Cancer Res. 2003, 9, 4983− 4991. (203) Yoon, K. J. P.; Krull, E. J.; Morton, C. L.; Bornmann, W. G.; Lee, R. E.; Potter, P. M.; Danks, M. K. Activation of a camptothecin prodrug by specific carboxylesterases as predicted by quantitative structure-activity relationship and molecular docking studies. Mol. Cancer Ther. 2003, 2, 1171−1181. (204) Tak, R. V.; Pal, D.; Gao, H.; Dey, S.; Mitra, A. K. Transport of acyclovir ester prodrugs through rabbit cornea and SIRC-rabbit corneal epithelial cell line. J. Pharm. Sci. 2001, 90, 1505−1515. (205) Tougou, K.; Nakamura, A.; Watanabe, S.; Okuyama, Y.; Morino, A. Paraoxonase Has a Major Role in the Hydrolysis of Prulifloxacin (NM441), a Prodrug of a New Antibacterial Agent. Drug Metab. Dispos. 1998, 26, 355−359. (206) Schiff, P. B.; Fant, J.; Horwitz, S. B. Promotion of microtubule assembly in vitro by taxol. Nature 1979, 277, 665−667. (207) Magnani, M.; Maccari, G.; Andreu, J. M.; Díaz, J. F.; Botta, M. Possible binding site for paclitaxel at microtubule pores. FEBS J. 2009, 276, 2701−2712. (208) Meurer-Grob, P.; Kasparian, J.; Wade, R. H. Microtubule Structure at Improved Resolution. Biochemistry 2001, 40, 8000−8008. (209) Wang, X.; Li, J.; Wang, Y.; Cho, K. J.; Kim, G.; Gjyrezi, A.; Koenig, L.; Giannakakou, P.; Shin, H. J. C.; Tighiouart, M.; Nie, S.; Chen, Z.; Shin, D. M. HFT-T, a Targeting Nanoparticle, Enhances Specific Delivery of Paclitaxel to Folate Receptor-Positive Tumors. ACS Nano 2009, 3, 3165−3174. (210) Ojima, I.; Geng, X.; Wu, X.; Qu, C.; Borella, C. P.; Xie, H.; Wilhelm, S. D.; Leece, B. A.; Bartle, L. M.; Goldmacher, V. S.; Chari, R. V. J. Tumor-Specific Novel Taxoid−Monoclonal Antibody Conjugates. J. Med. Chem. 2002, 45, 5620−5623. (211) Singer, J. W. Paclitaxel poliglumex (XYOTAX, CT-2103): A macromolecular taxane. J. Controlled Release 2005, 109, 120−126. (212) Lim, J.; Chouai, A.; Lo, S.-T.; Liu, W.; Sun, X.; Simanek, E. E. Design, Synthesis, Characterization, and Biological Evaluation of Triazine Dendrimers Bearing Paclitaxel Using Ester and Ester/ Disulfide Linkages. Bioconjugate Chem. 2009, 20, 2154−2161. (213) Yasuda, Y.; Kageyama, T.; Akamine, A.; Shibata, M.; Kominami, E.; Uchiyama, Y.; Yamamoto, K. Characterization of New Fluorogenic Substrates for the Rapid and Sensitive Assay of Cathepsin E and Cathepsin D. J. Biochem. 1999, 125, 1137−1143. (214) Damen, E. W. P.; Wiegerinck, P. H. G.; Braamer, L.; Sperling, D.; de Vos, D.; Scheeren, H. W. Paclitaxel esters of malic acid as prodrugs with improved water solubility. Bioorg. Med. Chem. 2000, 8, 427−432. 3424

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

(215) Zhang, Y.; Thomas, T. P.; Lee, K.-H.; Li, M.; Zong, H.; Desai, A. M.; Kotlyar, A.; Huang, B.; Banaszak Holl, M. M.; Baker, J. R., Jr. Polyvalent saccharide-functionalized generation 3 poly(amidoamine) dendrimer−methotrexate conjugate as a potential anticancer agent. Bioorg. Med. Chem. 2011, 19, 2557−2564. (216) Rosowsky, A.; Forsch, R.; Uren, J.; Wick, M. Methotrexate analogs. 14. Synthesis of new γ-substituted derivatives as dihydrofolate reductase inhibitors and potential anticancer agents. J. Med. Chem. 1981, 24, 1450−1455. (217) Schnell, J. R.; Dyson, H. J.; Wright, P. E. Structure, dynamics, and catalytic function of dihydrofolate. Annu. Rev. Biophys. Biomol. Struct. 2004, 33, 119−140. (218) Williams, J. W.; Morrison, J. F.; Duggleby, R. G. Methotrexate, a high-affinity pseudosubstrate of dihydrofolate reductase. Biochemistry 1979, 18, 2567−2573. (219) Tattersall, M. H. N.; Brown, B.; Frei, E. The reversal of methotrexate toxicity by thymidine with maintenance of antitumour effects. Nature 1975, 253, 198−200. (220) Majoros, I. J.; Thomas, T. P.; Mehta, C. B.; Baker, J. R., Jr. Poly(amidoamine) Dendrimer-Based Multifunctional Engineered Nanodevice for Cancer Therapy. J. Med. Chem. 2005, 48, 5892−5899. (221) Zhang, Y.; Thomas, T. P.; Desai, A.; Zong, H.; Leroueil, P. R.; Majoros, I. J.; Baker, J. R. Targeted Dendrimeric Anticancer Prodrug: A Methotrexate-Folic Acid-Poly(amidoamine) Conjugate and a Novel, Rapid, “One Pot” Synthetic Approach. Bioconjugate Chem. 2010, 21, 489−495. (222) Zong, H.; Thomas, T. P.; Lee, K.-H.; Desai, A. M.; Li, M.-h.; Kotlyar, A.; Zhang, Y.; Leroueil, P. R.; Gam, J. J.; Banaszak Holl, M. M.; Baker, J. R. Bifunctional PAMAM Dendrimer Conjugates of Folic Acid and Methotrexate with Defined Ratio. Biomacromolecules 2012, 13, 982−991. (223) Ward, B. B.; Dunham, T.; Majoros, I. J.; Baker, J. R., Jr. Targeted Dendrimer Chemotherapy in an Animal Model for Head and Neck Squamous Cell Carcinoma. J. Oral Maxillofac. Surg. 2011, 69, 2452−2459. (224) Thomas, T. P.; Kukowska-Latallo, J. R. In Dendrimer-Based Nanomedicine; Majoros, I., Baker, J. R., Jr., Eds.; Pan Stanford: Hackensack, NJ, 2008. (225) Wong, E.; Giandomenico, C. M. Current Status of PlatinumBased Antitumor Drugs. Chem. Rev. (Washington, DC, U. S.) 1999, 99, 2451−2466. (226) Feazell, R. P.; Nakayama-Ratchford, N.; Dai, H.; Lippard, S. J. Soluble Single-Walled Carbon Nanotubes as Longboat Delivery Systems for Platinum(IV) Anticancer Drug Design. J. Am. Chem. Soc. 2007, 129, 8438−8439. (227) Barnes, K. R.; Kutikov, A.; Lippard, S. J. Synthesis, Characterization, and Cytotoxicity of a Series of Estrogen-Tethered Platinum(IV) Complexes. Chem. Biol. (Oxford, U. K.) 2004, 11, 557− 564. (228) Schreiber, S. Chemistry and biology of the immunophilins and their immunosuppressive ligands. Science 1991, 251, 283−287. (229) DeFeo-Jones, D.; Garsky, V. M.; Wong, B. K.; Feng, D.-M.; Bolyar, T.; Haskell, K.; Kiefer, D. M.; Leander, K.; McAvoy, E.; Lumma, P.; Wai, J.; Senderak, E. T.; Motzel, S. L.; Keenan, K.; Zwieten, M. V.; Lin, J. H.; Freidinger, R.; Huff, J.; Oliff, A.; Jones, R. E. A peptide-doxorubicin ’prodrug’ activated by prostate-specific antigen selectively kills prostate tumor cells positive for prostate-specific antigen in vivo. Nat. Med. (N. Y., NY, U. S.) 2000, 6, 1248−1252. (230) Wong, B. K.; DeFeo-Jones, D.; Jones, R. E.; Garsky, V. M.; Feng, D.-M.; Oliff, A.; Chiba, M.; Ellis, J. D.; Lin, J. H. PSA-Specific and Non-PSA-Specific Conversion of a PSA-Targeted Peptide Conjugate of Doxorubicin to Its Active Metabolites. Drug Metab. Dispos. 2001, 29, 313−318. (231) Hervio, L. S.; Coombs, G. S.; Bergstrom, R. C.; Trivedi, K.; Corey, D. R.; Madison, E. L. Negative selectivity and the evolution of protease cascades: the specificity of plasmin for peptide and protein substrates. Chem. Biol. (Oxford, U. K.) 2000, 7, 443−452. (232) de Groot, F. M. H.; van Berkom, L. W. A.; Scheeren, H. W. Synthesis and Biological Evaluation of 2‘-Carbamate-Linked and 2‘-

Carbonate-Linked Prodrugs of Paclitaxel: Selective Activation by the Tumor-Associated Protease Plasmin. J. Med. Chem. 2000, 43, 3093− 3102. (233) Deng, S.-J.; Bickett, D. M.; Mitchell, J. L.; Lambert, M. H.; Blackburn, R. K.; Carter, H. L.; Neugebauer, J.; Pahel, G.; Weiner, M. P.; Moss, M. L. Substrate Specificity of Human Collagenase 3 Assessed Using a Phage-displayed Peptide Library. J. Biol. Chem. 2000, 275, 31422−31427. (234) Silver, D. A.; Pellicer, I.; Fair, W. R.; Heston, W. D.; CordonCardo, C. Prostate-specific membrane antigen expression in normal and malignant human tissues. Clin. Cancer Res. 1997, 3, 81−85. (235) Page-McCaw, A.; Ewald, A. J.; Werb, Z. Matrix metalloproteinases and the regulation of tissue remodelling. Nat. Rev. Mol. Cell Biol. 2007, 8, 221−233. (236) Roy, R.; Yang, J.; Moses, M. A. Matrix Metalloproteinases As Novel Biomarker s and Potential Therapeutic Targets in Human Cancer. J. Clin. Oncol. 2009, 27, 5287−5297. (237) Hauff, S. J.; Raju, S. C.; Orosco, R. K.; Gross, A. M.; DiazPerez, J. A.; Savariar, E.; Nashi, N.; Hasselman, J.; Whitney, M.; Myers, J. N.; Lippman, S. M.; Tsien, R. Y.; Ideker, T.; Nguyen, Q. T. MatrixMetalloproteinases in Head and Neck Carcinoma−Cancer Genome Atlas Analysis and Fluorescence Imaging in Mice. Otolaryngol.Head Neck Surg. 2014, 151, 612−618. (238) Savariar, E. N.; Felsen, C. N.; Nashi, N.; Jiang, T.; Ellies, L. G.; Steinbach, P.; Tsien, R. Y.; Nguyen, Q. T. Real-time In Vivo Molecular Detection of Primary Tumors and Metastases with Ratiometric Activatable Cell-Penetrating Peptides. Cancer Res. 2013, 73, 855−864. (239) Olson, E. S.; Aguilera, T. A.; Jiang, T.; Ellies, L. G.; Nguyen, Q. T.; Wong, E. H.; Gross, L. A.; Tsien, R. Y. In vivo characterization of activatable cell penetrating peptides for targeting protease activity in cancer. Integr. Biol. 2009, 1, 382−393. (240) Olson, E. S.; Jiang, T.; Aguilera, T. A.; Nguyen, Q. T.; Ellies, L. G.; Scadeng, M.; Tsien, R. Y. Activatable cell penetrating peptides linked to nanoparticles as dual probes for in vivo fluorescence and MR imaging of proteases. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 4311− 4316. (241) Jiang, T.; Olson, E. S.; Nguyen, Q. T.; Roy, M.; Jennings, P. A.; Tsien, R. Y. Tumor imaging by means of proteolytic activation of cellpenetrating peptides. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 17867− 17872. (242) Whitney, M.; Savariar, E. N.; Friedman, B.; Levin, R. A.; Crisp, J. L.; Glasgow, H. L.; Lefkowitz, R.; Adams, S. R.; Steinbach, P.; Nashi, N.; Nguyen, Q. T.; Tsien, R. Y. Ratiometric Activatable CellPenetrating Peptides Provide Rapid In Vivo Readout of Thrombin Activation. Angew. Chem., Int. Ed. 2013, 52, 325−330. (243) Felsen, C. N.; Savariar, E. N.; Whitney, M.; Tsien, R. Y. Detection and monitoring of localized matrix metalloproteinase upregulation in a murine model of asthma. Am. J. Physiol.: Lung Cell. Mol. Physiol. 2014, 306, L764−L774. (244) Olson, E. S.; Whitney, M. A.; Friedman, B.; Aguilera, T. A.; Crisp, J. L.; Baik, F. M.; Jiang, T.; Baird, S. M.; Tsimikas, S.; Tsien, R. Y.; Nguyen, Q. T. In vivo fluorescence imaging of atherosclerotic plaques with activatable cell-penetrating peptides targeting thrombin activity. Integr. Biol. 2012, 4, 595−605. (245) Shen, W.-C.; Ryser, H. J. P. Cis-aconityl spacer between daunomycin and macromolecular carriers: A model of pH-sensitive linkage releasing drug from a lysosomotropic conjugate. Biochem. Biophys. Res. Commun. 1981, 102, 1048−1054. (246) Butler, P. J.; Harris, J. I.; Hartley, B. S.; Leberman, R. Reversible blocking of peptide amino groups by maleic anhydride. Biochem. J. 1967, 103, 78b−79p. (247) Habeeb, A. F. S. A.; Atassi, M. Z. Enzymic and immunochemical properties of lysozyme. Evaluation of several amino group reversible blocking reagents. Biochemistry 1970, 9, 4939−4944. (248) Akinboye, E. S.; Rosen, M. D.; Denmeade, S. R.; Kwabi-Addo, B.; Bakare, O. Design, Synthesis, and Evaluation of pH-Dependent Hydrolyzable Emetine Analogues as Treatment for Prostate Cancer. J. Med. Chem. 2012, 55, 7450−7459. 3425

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

(249) Chau, Y.; Padera, R. F.; Dang, N. M.; Langer, R. Antitumor efficacy of a novel polymer−peptide−drug conjugate in human tumor xenograft models. Int. J. Cancer 2006, 118, 1519−1526. (250) Chau, Y.; Tan, F. E.; Langer, R. Synthesis and Characterization of Dextran−Peptide−Methotrexate Conjugates for Tumor Targeting via Mediation by Matrix Metalloproteinase II and Matrix Metalloproteinase IX. Bioconjugate Chem. 2004, 15, 931−941. (251) Huang, B.; Desai, A.; Zong, H.; Tang, S.; Leroueil, P.; Baker, J. R., Jr. Copper-free click conjugation of methotrexate to a PAMAM dendrimer platform. Tetrahedron Lett. 2011, 52, 1411−1414. (252) Thomas, T. P.; Joice, M.; Sumit, M.; Silpe, J. E.; Kotlyar, A.; Bharathi, S.; Kukowska-Latallo, J.; Baker, J. R.; Choi, S. K. Design and In Vitro Validation of Multivalent Dendrimer Methotrexates as a Folate-targeting Anticancer Therapeutic. Curr. Pharm. Des. 2013, 19, 6594−6605. (253) Han, J.; Lim, S.-J.; Lee, M.-K.; Kim, C.-K. Altered Pharmacokinetics and Liver Targetability of Methotrexate by Conjugation with Lactosylated Albumins. Drug Delivery 2001, 8, 125−134. (254) Ryser, H. J.-P.; Shen, W.-C. Conjugation of methotrexate to poly(L-lysine) increases drug transport and overcomes drug resistance in cultured cells. Proc. Natl. Acad. Sci. U. S. A. 1978, 75, 3867−3870. (255) Endo, N.; Takeda, Y.; Kishida, K.; Kato, Y.; Saito, M.; Umemoto, N.; Hara, T. Target-selective cytotoxicity of methotrexate conjugated with monoclonal anti-MM46 antibody. Cancer Immunol. Immunother. 1987, 25, 1−6. (256) Umemoto, N.; Kato, Y.; Endo, N.; Takeda, Y.; Hara, T. Preparation and in vitro cytotoxicity of a methotrexate-anti-MM46 monoclonal antibody conjugate via an oligopeptide spacer. Int. J. Cancer 1989, 43, 677−684. (257) Sengupta, S.; Eavarone, D.; Capila, I.; Zhao, G.; Watson, N.; Kiziltepe, T.; Sasisekharan, R. Temporal targeting of tumour cells and neovasculature with a nanoscale delivery system. Nature 2005, 436, 568−572. (258) Zhang, H.; Li, F.; Yi, J.; Gu, C.; Fan, L.; Qiao, Y.; Tao, Y.; Cheng, C.; Wu, H. Folate-decorated maleilated pullulan−doxorubicin conjugate for active tumor-targeted drug delivery. Eur. J. Pharm. Sci. 2011, 42, 517−526. (259) Son, Y. J.; Jang, J.-S.; Cho, Y. W.; Chung, H.; Park, R.-W.; Kwon, I. C.; Kim, I.-S.; Park, J. Y.; Seo, S. B.; Park, C. R.; Jeong, S. Y. Biodistribution and anti-tumor efficacy of doxorubicin loaded glycolchitosan nanoaggregates by EPR effect. J. Controlled Release 2003, 91, 135−145. (260) Yoo, H. S.; Lee, E. A.; Park, T. G. Doxorubicin-conjugated biodegradable polymeric micelles having acid-cleavable linkages. J. Controlled Release 2002, 82, 17−27. (261) Basu, S.; Harfouche, R.; Soni, S.; Chimote, G.; Mashelkar, R. A.; Sengupta, S. Nanoparticle-mediated targeting of MAPK signaling predisposes tumor to chemotherapy. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 7957−7961. (262) Fosang, A. J.; Last, K.; Fujii, Y.; Seiki, M.; Okada, Y. Membrane-type 1 MMP (MMP-14) cleaves at three sites in the aggrecan interglobular domain. FEBS Lett. 1998, 430, 186−190. (263) Ajayan, P. M. Nanotubes from Carbon. Chem. Rev. (Washington, DC, U. S.) 1999, 99, 1787−1800. (264) Tasis, D.; Tagmatarchis, N.; Bianco, A.; Prato, M. Chemistry of Carbon Nanotubes. Chem. Rev. (Washington, DC, U. S.) 2006, 106, 1105−1136. (265) Green, M. J.; Gough, A. K. S.; Devlin, J.; Smith, J.; Astin, P.; Taylor, D.; Emery, P. Serum MMP-3 and MMP-1 and progression of joint damage in early rheumatoid arthritis. Rheumatology (Oxford, U. K.) 2003, 42, 83−88. (266) Schmalfeldt, B.; Prechtel, D.; Härting, K.; Späthe, K.; Rutke, S.; Konik, E.; Fridman, R.; Berger, U.; Schmitt, M.; Kuhn, W.; Lengyel, E. Increased Expression of Matrix Metalloproteinases (MMP)-2, MMP-9, and the Urokinase-Type Plasminogen Activator Is Associated with Progression from Benign to Advanced Ovarian Cancer. Clin. Cancer Res. 2001, 7, 2396−2404.

(267) Incorvaia, L.; Badalamenti, G.; Rini, G.; Arcara, C.; Fricano, S.; Sferrazza, C.; Di Trapani, D.; Gebbia, N.; Leto, G. MMP-2, MMP-9 and Activin A Blood Levels in Patients with Breast Cancer or Prostate Cancer Metastatic to the Bone. Anticancer Res. 2007, 27, 1519−1525. (268) Li, M.-H.; Choi, S. K.; Thomas, T. P.; Desai, A.; Lee, K.-H.; Kotlyar, A.; Banaszak Holl, M. M.; Baker, J. R., Jr. Dendrimer-Based Multivalent Methotrexates As Dual Acting Nanoconjugates for Cancer Cell Targeting. Eur. J. Med. Chem. 2012, 47, 560−572. (269) Neuzil, J. Vitamin E succinate and cancer treatment: a vitamin E prototype for selective antitumour activity. Br. J. Cancer 2003, 89, 1822−1826. (270) Hu, X.; Wang, R.; Yue, J.; Liu, S.; Xie, Z.; Jing, X. Targeting and anti-tumor effect of folic acid-labeled polymer-Doxorubicin conjugates with pH-sensitive hydrazone linker. J. Mater. Chem. 2012, 22, 13303−13310. (271) Padilla De Jesus, O. L.; Ihre, H. R.; Gagne, L.; Frechet, J. M. J.; Szoka, F. C. Polyester Dendritic Systems for Drug Delivery Applications: In Vitro and In Vivo Evaluation. Bioconjugate Chem. 2002, 13, 453−461. (272) Bae, Y.; Jang, W.-D.; Nishiyama, N.; Fukushima, S.; Kataoka, K. Multifunctional polymeric micelles with folate-mediated cancer cell targeting and pH-triggered drug releasing properties for active intracellular drug delivery. Mol. BioSyst. 2005, 1, 242−250. (273) Kojima, C.; Suehiro, T.; Watanabe, K.; Ogawa, M.; Fukuhara, A.; Nishisaka, E.; Harada, A.; Kono, K.; Inui, T.; Magata, Y. Doxorubicin-conjugated dendrimer/collagen hybrid gels for metastasis-associated drug delivery systems. Acta Biomater. 2013, 9, 5673− 5680. (274) Bae, Y.; Nishiyama, N.; Fukushima, S.; Koyama, H.; Yasuhiro, M.; Kataoka, K. Preparation and Biological Characterization of Polymeric Micelle Drug Carriers with Intracellular pH-Triggered Drug Release Property: Tumor Permeability, Controlled Subcellular Drug Distribution, and Enhanced in Vivo Antitumor Efficacy. Bioconjugate Chem. 2005, 16, 122−130. (275) Wahl, A. F.; Donaldson, K. L.; Mixan, B. J.; Trail, P. A.; Siegall, C. B. Selective tumor sensitization to taxanes with the mAb-drug conjugate cBR96-doxorubicin. Int. J. Cancer 2001, 93, 590−600. (276) Rodrigues, P. C. A.; Scheuermann, K.; Stockmar, C.; Maier, G.; Fiebig, H. H.; Unger, C.; Mülhaupt, R.; Kratz, F. Synthesis and In vitro efficacy of acid-Sensitive poly(ethylene glycol) paclitaxel conjugates. Bioorg. Med. Chem. Lett. 2003, 13, 355−360. (277) Binauld, S.; Scarano, W.; Stenzel, M. H. pH-Triggered Release of Platinum Drugs Conjugated to Micelles via an Acid-Cleavable Linker. Macromolecules 2012, 45, 6989−6999. (278) Zhou, L.; Cheng, R.; Tao, H.; Ma, S.; Guo, W.; Meng, F.; Liu, H.; Liu, Z.; Zhong, Z. Endosomal pH-Activatable Poly(ethylene oxide)-graft-Doxorubicin Prodrugs: Synthesis, Drug Release, and Biodistribution in Tumor-Bearing Mice. Biomacromolecules 2011, 12, 1460−1467. (279) Prabaharan, M.; Grailer, J. J.; Pilla, S.; Steeber, D. A.; Gong, S. Amphiphilic multi-arm-block copolymer conjugated with doxorubicin via pH-sensitive hydrazone bond for tumor-targeted drug delivery. Biomaterials 2009, 30, 5757−5766. (280) Hrubý, M.; Koňaḱ , Č .; Ulbrich, K. Polymeric micellar pHsensitive drug delivery system for doxorubicin. J. Controlled Release 2005, 103, 137−148. (281) Xiao, Y.; Hong, H.; Matson, V. Z.; Javadi, A.; Xu, W.; Yang, Y.; Zhang, Y.; Engle, J. W.; Nickles, R. J.; Cai, W.; Steeber, D. A.; Gong, S. Gold Nanorods Conjugated with Doxorubicin and cRGD for Combined Anticancer Drug Delivery and PET Imaging. Theranostics 2012, 2, 757−768. (282) Wang, F.; Wang, Y.-C.; Dou, S.; Xiong, M.-H.; Sun, T.-M.; Wang, J. Doxorubicin-Tethered Responsive Gold Nanoparticles Facilitate Intracellular Drug Delivery for Overcoming Multidrug Resistance in Cancer Cells. ACS Nano 2011, 5, 3679−3692. (283) Willner, D.; Trail, P. A.; Hofstead, S. J.; King, H. D.; Lasch, S. J.; Braslawsky, G. R.; Greenfield, R. S.; Kaneko, T.; Firestone, R. A. (6Maleimidocaproyl)hydrazone of doxorubicin. A new derivative for the 3426

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

preparation of immunoconjugates of doxorubicin. Bioconjugate Chem. 1993, 4, 521−527. (284) Kalia, J.; Raines, R. T. Hydrolytic Stability of Hydrazones and Oximes. Angew. Chem., Int. Ed. 2008, 47, 7523−7526. (285) Yang, J.; Chen, H.; Vlahov, I. R.; Cheng, J.-X.; Low, P. S. Evaluation of disulfide reduction during receptor-mediated endocytosis by using FRET imaging. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 13872−13877. (286) Paulsen, C. E.; Carroll, K. S. Cysteine-Mediated Redox Signaling: Chemistry, Biology, and Tools for Discovery. Chem. Rev. (Washington, DC, U. S.) 2013, 113, 4633−4679. (287) Lee, M. H.; Yang, Z.; Lim, C. W.; Lee, Y. H.; Dongbang, S.; Kang, C.; Kim, J. S. Disulfide-Cleavage-Triggered Chemosensors and Their Biological Applications. Chem. Rev. (Washington, DC, U. S.) 2013, 113, 5071−5109. (288) Kuśmierek, K.; Chwatko, G.; Głowacki, R.; Bald, E. Determination of endogenous thiols and thiol drugs in urine by HPLC with ultraviolet detection. J. Chromatogr. B: Anal. Technol. Biomed. Life Sci. 2009, 877, 3300−3308. (289) Green, R. M.; Graham, M.; O’Donovan, M. R.; Chipman, J. K.; Hodges, N. J. Subcellular compartmentalization of glutathione: Correlations with parameters of oxidative stress related to genotoxicity. Mutagenesis 2006, 21, 383−390. (290) Reddy, J.; Westrick, E.; Vlahov, I.; Howard, S.; Santhapuram, H.; Leamon, C. Folate receptor specific anti-tumor activity of folate− mitomycin conjugates. Cancer Chemother. Pharmacol. 2006, 58, 229− 236. (291) Trail, P. A.; Willner, D.; Lasch, S. J.; Henderson, A. J.; Greenfield, R. S.; King, D.; Zoeckler, M. E.; Braslawsky, G. R. Antigenspecific Activity of Carcinoma-reactive BR64-Doxorubicin Conjugates Evaluated in Vitro and in Human Tumor Xenograft Models. Cancer Res. 1992, 52, 5693−5700. (292) Vrudhula, V. M.; MacMaster, J. F.; Li, Z.; Kerr, D. E.; Senter, P. D. Reductively activated disulfide prodrugs of paclitaxel. Bioorg. Med. Chem. Lett. 2002, 12, 3591−3594. (293) Yang, Z.; Lee, J. H.; Jeon, H. M.; Han, J. H.; Park, N.; He, Y.; Lee, H.; Hong, K. S.; Kang, C.; Kim, J. S. Folate-Based Near-Infrared Fluorescent Theranostic Gemcitabine Delivery. J. Am. Chem. Soc. 2013, 135, 11657−11662. (294) Hinman, L. M.; Hamann, P. R.; Wallace, R.; Menendez, A. T.; Durr, F. E.; Upeslacis, J. Preparation and Characterization of Monoclonal Antibody Conjugates of the Calicheamicins: A Novel and Potent Family of Antitumor Antibiotics. Cancer Res. 1993, 53, 3336−3342. (295) Wang, X.; Cai, X.; Hu, J.; Shao, N.; Wang, F.; Zhang, Q.; Xiao, J.; Cheng, Y. Glutathione-Triggered “Off−On” Release of Anticancer Drugs from Dendrimer-Encapsulated Gold Nanoparticles. J. Am. Chem. Soc. 2013, 135, 9805−9810. (296) Elkins, K.; Zheng, B.; Go, M.; Slaga, D.; Du, C.; Scales, S. J.; Yu, S.-F.; McBride, J.; de Tute, R.; Rawstron, A.; Jack, A. S.; Ebens, A.; Polson, A. G. FcRL5 as a Target of Antibody−Drug Conjugates for the Treatment of Multiple Myeloma. Mol. Cancer Ther. 2012, 11, 2222− 2232. (297) Bailey, K. M.; Wojtkowiak, J. W.; Hashim, A. I.; Gillies, R. J. In Advances in Biochemical Pharmacology; Keiran, S. M. S., Ed.; Academic Press: Waltham, MA, 2012; Vol. 65. (298) Wilson, W. R.; Hay, M. P. Targeting hypoxia in cancer therapy. Nat. Rev. Cancer 2011, 11, 393−410. (299) Naughton, D. P. Drug targeting to hypoxic tissue using selfinactivating bioreductive delivery systems. Adv. Drug Delivery Rev. 2001, 53, 229−233. (300) Brown, J. M.; Wilson, W. R. Exploiting tumour hypoxia in cancer treatment. Nat. Rev. Cancer 2004, 4, 437−447. (301) Iyer, V. N.; Szybalski, W. Mitomycins and Porfiromycin: Chemical Mechanism of Activation and Cross-linking of DNA. Science 1964, 145, 55−58. (302) Phillips, R. M.; Hendriks, H. R.; Peters, G. J.; et al. EO9 (Apaziquone): from the clinic to the laboratory and back again. Br. J. Pharmacol. 2013, 168, 11−18.

(303) Hu, J.; Handisides, D. R.; Van Valckenborgh, E.; De Raeve, H.; Menu, E.; Vande Broek, I.; Liu, Q.; Sun, J. D.; Van Camp, B.; Hart, C. P.; Vanderkerken, K. Targeting the multiple myeloma hypoxic niche with TH-302, a hypoxia-activated prodrug. Blood 2010, 116, 1524− 1527. (304) Gu, Y.; Patterson, A. V.; Atwell, G. J.; Chernikova, S. B.; Brown, J. M.; Thompson, L. H.; Wilson, W. R. Roles of DNA repair and reductase activity in the cytotoxicity of the hypoxia-activated dinitrobenzamide mustard PR-104A. Mol. Cancer Ther. 2009, 8, 1714− 1723. (305) Yan, C.; Kepa, J. K.; Siegel, D.; Stratford, I. J.; Ross, D. Dissecting the Role of Multiple Reductases in Bioactivation and Cytotoxicity of the Antitumor Agent 2,5-Diaziridinyl-3-(hydroxymethyl)-6-methyl-1,4-benzoquinone (RH1). Mol. Pharmacol. 2008, 74, 1657−1665. (306) Wang, B.; Liu, S.; Borchardt, R. T. Development of a Novel Redox-Sensitive Protecting Group for Amines Which Utilizes a Facilitated Lactonization Reaction. J. Org. Chem. 1995, 60, 539−543. (307) Huang, B.; Tang, S.; Desai, A.; Cheng, X.-m.; Kotlyar, A.; Spek, A. V. D.; Thomas, T. P.; Baker, J. R., Jr. Human plasma-mediated hypoxic activation of indolequinone-based naloxone pro-drugs. Bioorg. Med. Chem. Lett. 2009, 19, 5016−5020. (308) Parveen, I.; Naughton, D. P.; Whish, W. J. D.; Threadgill, M. D. 2-Nitroimidazol-5-ylmethyl as a potential bioreductively activated prodrug system: reductively triggered release of the parp inhibitor 5bromoisoquinolinone. Bioorg. Med. Chem. Lett. 1999, 9, 2031−2036. (309) Mahmud, N. P.; Garrett, S. W.; Threadgill, M. D. The 5nitrofuran-2-ylmethylidene group as a potential bioreductively activated prodrug system for diol-containing drugs. Anti-Cancer Drug Des. 1998, 13, 655−662. (310) Ferrer, S.; Naughton, D. P.; Threadgill, M. D. Studies on the reductively triggered release of heterocyclic and steroid drugs from 5nitrothien-2-ylmethyl prodrugs. Tetrahedron 2003, 59, 3437−3444. (311) Cairns, R. A.; Papandreou, I.; Sutphin, P. D.; Denko, N. C. Metabolic targeting of hypoxia and HIF1 in solid tumors can enhance cytotoxic chemotherapy. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 9445−9450. (312) Hunter, F. W.; Wang, J.; Patel, R.; Hsu, H.-L.; Hickey, A. J. R.; Hay, M. P.; Wilson, W. R. Homologous recombination repairdependent cytotoxicity of the benzotriazine di-N-oxide CEN-209: Comparison with other hypoxia-activated prodrugs. Biochem. Pharmacol. (Amsterdam, Neth.) 2012, 83, 574−585. (313) Levine, M. N.; Raines, R. T. Trimethyl lock: a trigger for molecular release in chemistry, biology, and pharmacology. Chem. Sci. 2012, 3, 2412−2420. (314) Shan, D.; Zheng, A.; Ballard, E. C.; Wang, W.; Borchardt, R. T.; Wang, B. A Facilitated Cyclic Ether Formation and Its Potential Application in Solid-Phase Peptide and Organic Synthesis. Chem. Pharm. Bull. 2000, 48, 238−244. (315) Ong, W.; Yang, Y.; Cruciano, A. C.; McCarley, R. L. RedoxTriggered Contents Release from Liposomes. J. Am. Chem. Soc. 2008, 130, 14739−14744. (316) van Haaften, R. I. M.; Evelo, C. T. A.; Haenen, G. R. M. M.; Bast, A. No reduction of α-tocopherol quinone by glutathione in rat liver microsomes. Biochem. Pharmacol. (Amsterdam, Neth.) 2001, 61, 715−719. (317) Borchardt, R. T.; Cohen, L. A. Stereopopulation control. II. Rate enhancement of intramolecular nucleophilic displacement. J. Am. Chem. Soc. 1972, 94, 9166−9174. (318) Tedeschi, G.; Chen, S.; Massey, V. DT-diaphorase: Redox Potential, Steady-State, and Rapid Reaction Studies. J. Biol. Chem. 1995, 270, 1198−1204. (319) Damen, E. W. P.; Nevalainen, T. J.; van den Bergh, T. J. M.; de Groot, F. M. H.; Scheeren, H. W. Synthesis of novel paclitaxel prodrugs designed for bioreductive activation in hypoxic tumour tissue. Bioorg. Med. Chem. 2002, 10, 71−77. (320) Cogan, P. S.; Koch, T. H. Studies of Targeting and Intracellular Trafficking of an Anti-Androgen Doxorubicin-Formaldehyde Con3427

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

jugate in PC-3 Prostate Cancer Cells Bearing Androgen Receptor-GFP Chimera. J. Med. Chem. 2004, 47, 5690−5699. (321) Wong, A. D.; DeWit, M. A.; Gillies, E. R. Amplified release through the stimulus triggered degradation of self-immolative oligomers, dendrimers, and linear polymers. Adv. Drug Delivery Rev. 2012, 64, 1031−1045. (322) Houba, P. H. J.; Boven, E.; van der Meulen-Muileman, I. H.; Leenders, R. G. G.; Scheeren, J. W.; Pinedo, H. M.; Haisma, H. J. A novel doxorubicin-glucuronide prodrug DOX-GA3 for tumourselective chemotherapy: distribution and efficacy in experimental human ovarian cancer. Br. J. Cancer 2001, 84, 550−557. (323) Schuster, H. J.; Krewer, B.; von Hof, J. M.; Schmuck, K.; Schuberth, I.; Alves, F.; Tietze, L. F. Synthesis of the first spacer containing prodrug of a duocarmycin analogue and determination of its biological activity. Org. Biomol. Chem. 2010, 8, 1833−1842. (324) Greenwald, R. B.; Choe, Y. H.; Conover, C. D.; Shum, K.; Wu, D.; Royzen, M. Drug Delivery Systems Based on Trimethyl Lock Lactonization: Poly(ethylene glycol) Prodrugs of Amino-Containing Compounds. J. Med. Chem. 2000, 43, 475−487. (325) Rodrigues, M. L.; Presta, L. G.; Kotts, C. E.; Wirth, C.; Mordenti, J.; Osaka, G.; Wong, W. L. T.; Nuijens, A.; Blackburn, B.; Carter, P. Development of a humanized disulfide-stabilized antip185(HER2) Fv-β-lactamase fusion protein for activation of a cephalosporin doxorubicin prodrug. Cancer Res. 1995, 55, 63−70. (326) Rodrigues, M. L.; Carter, P.; Wirth, C.; Mullins, S.; Lee, A.; Blackburn, B. K. Synthesis and β-lactamase-mediated activation of a cephalosporin-taxol prodrug. Chem. Biol. (Oxford, U. K.) 1995, 2, 223−227. (327) Chen, S.; Zhao, X.; Chen, J.; Chen, J.; Kuznetsova, L.; Wong, S. S.; Ojima, I. Mechanism-Based Tumor-Targeting Drug Delivery System. Validation of Efficient Vitamin Receptor-Mediated Endocytosis and Drug Release. Bioconjugate Chem. 2010, 21, 979−987. (328) Amir, R. J.; Pessah, N.; Shamis, M.; Shabat, D. Self-Immolative Dendrimers. Angew. Chem., Int. Ed. 2003, 42, 4494−4499. (329) Shamis, M.; Lode, H. N.; Shabat, D. Bioactivation of SelfImmolative Dendritic Prodrugs by Catalytic Antibody 38C2. J. Am. Chem. Soc. 2004, 126, 1726−1731. (330) Nicewarner Peña, S. R.; Raina, S.; Goodrich, G. P.; Fedoroff, N. V.; Keating, C. D. Hybridization and Enzymatic Extension of Au Nanoparticle-Bound Oligonucleotides. J. Am. Chem. Soc. 2002, 124, 7314−7323. (331) Walker, J. W.; McCray, J. A.; Hess, G. P. Photolabile protecting groups for an acetylcholine receptor ligand. Synthesis and photochemistry of a new class of o-nitrobenzyl derivatives and their effects on receptor function. Biochemistry 1986, 25, 1799−1805. (332) Mayer, G.; Heckel, A. Biologically Active Molecules with a Light Switch. Angew. Chem., Int. Ed. 2006, 45, 4900−4921. (333) Furuta, T.; Wang, S. S. H.; Dantzker, J. L.; Dore, T. M.; Bybee, W. J.; Callaway, E. M.; Denk, W.; Tsien, R. Y. Brominated 7hydroxycoumarin-4-ylmethyls: Photolabile protecting groups with biologically useful cross-sections for two photon photolysis. Proc. Natl. Acad. Sci. U. S. A. 1999, 96, 1193−1200. (334) Du, H.; Boyd, M. K. The 9-xanthenylmethyl group: a novel photocleavable protecting group for amines. Tetrahedron Lett. 2001, 42, 6645−6647. (335) Goard, M.; Aakalu, G.; Fedoryak, O. D.; Quinonez, C.; St. Julien, J.; Poteet, S. J.; Schuman, E. M.; Dore, T. M. Light-Mediated Inhibition of Protein Synthesis. Chem. Biol. (Oxford, U. K.) 2005, 12, 685−693. (336) Lemke, E. A.; Summerer, D.; Geierstanger, B. H.; Brittain, S. M.; Schultz, P. G. Control of protein phosphorylation with a genetically encoded photocaged amino acid. Nat. Chem. Biol. 2007, 3, 769−772. (337) Momotake, A.; Lindegger, N.; Niggli, E.; Barsotti, R. J.; EllisDavies, G. C. R. The nitrodibenzofuran chromophore: a new caging group for ultra-efficient photolysis in living cells. Nat. Methods 2006, 3, 35−40.

(338) Shi, Y.; Koh, J. T. Light-Activated Transcription and Repression by Using Photocaged SERMs. ChemBioChem 2004, 5, 788−796. (339) Link, K. H.; Shi, Y.; Koh, J. T. Light Activated Recombination. J. Am. Chem. Soc. 2005, 127, 13088−13089. (340) Neveu, P.; Aujard, I.; Benbrahim, C.; Le Saux, T.; Allemand, J.F.; Vriz, S.; Bensimon, D.; Jullien, L. A Caged Retinoic Acid for Oneand Two-Photon Excitation in Zebrafish Embryos. Angew. Chem., Int. Ed. 2008, 47, 3744−3746. (341) Agasti, S. S.; Chompoosor, A.; You, C.-C.; Ghosh, P.; Kim, C. K.; Rotello, V. M. Photoregulated Release of Caged Anticancer Drugs from Gold Nanoparticles. J. Am. Chem. Soc. 2009, 131, 5728−5729. (342) Sauers, D. J.; Temburni, M. K.; Biggins, J. B.; Ceo, L. M.; Galileo, D. S.; Koh, J. T. Light-Activated Gene Expression Directs Segregation of Co-cultured Cells in Vitro. ACS Chem. Biol. 2010, 5, 313−320. (343) Inlay, M. A.; Choe, V.; Bharathi, S.; Fernhoff, N. B.; Baker, J. R.; Weissman, I. L.; Choi, S. K. Synthesis of a photocaged tamoxifen for light-dependent activation of Cre-ER recombinase-driven gene modification. Chem. Commun. (Cambridge, U. K.) 2013, 49, 4971− 4973. (344) Lin, Q.; Huang, Q.; Li, C.; Bao, C.; Liu, Z.; Li, F.; Zhu, L. Anticancer Drug Release from a Mesoporous Silica Based Nanophotocage Regulated by Either a One- or Two-Photon Process. J. Am. Chem. Soc. 2010, 132, 10645−10647. (345) Lin, W.; Peng, D.; Wang, B.; Long, L.; Guo, C.; Yuan, J. A Model for Light-Triggered Porphyrin Anticancer Prodrugs Based on an o-Nitrobenzyl Photolabile Group. Eur. J. Org. Chem. 2008, 2008, 793−796. (346) Noguchi, M.; Skwarczynski, M.; Prakash, H.; Hirota, S.; Kimura, T.; Hayashi, Y.; Kiso, Y. Development of novel water-soluble photocleavable protective group and its application for design of photoresponsive paclitaxel prodrugs. Bioorg. Med. Chem. 2008, 16, 5389−5397. (347) Skwarczynski, M.; Noguchi, M.; Hirota, S.; Sohma, Y.; Kimura, T.; Hayashi, Y.; Kiso, Y. Development of first photoresponsive prodrug of paclitaxel. Bioorg. Med. Chem. Lett. 2006, 16, 4492−4496. (348) Hu, X.; Tian, J.; Liu, T.; Zhang, G.; Liu, S. Photo-Triggered Release of Caged Camptothecin Prodrugs from Dually Responsive Shell Cross-Linked Micelles. Macromolecules 2013, 46, 6243−6256. (349) Sinha, D. K.; Neveu, P.; Gagey, N.; Aujard, I.; Saux, T. L.; Rampon, C.; Gauron, C.; Kawakami, K.; Leucht, C.; Bally-Cuif, L.; Volovitch, M.; Bensimon, D.; Jullien, L.; Vriz, S. Photoactivation of the CreERT2 Recombinase for Conditional Site-Specific Recombination with High Spatiotemporal Resolution. Zebrafish 2010, 7, 199−204. (350) Aujard, I.; Benbrahim, C.; Gouget, M.; Ruel, O.; Baudin, J.-B.; Neveu, P.; Jullien, L. o-Nitrobenzyl Photolabile Protecting Groups with Red-Shifted Absorption: Syntheses and Uncaging Cross-Sections for One- and Two-Photon Excitation. Chem.Eur. J. 2006, 12, 6865− 6879. (351) Li, Y. M.; Shi, J.; Cai, R.; Chen, X.; Luo, Z. F.; Guo, Q. X. New quinoline-based caging groups synthesized for photo-regulation of aptamer activity. J. Photochem. Photobiol., A 2010, 211, 129−134. (352) Qvit, N.; Monderer-Rothkoff, G.; Ido, A.; Shalev, D. E.; Amster-Choder, O.; Gilon, C. Development of bifunctional photoactivatable benzophenone probes and their application to glycoside substrates. Pept. Sci. 2008, 90, 526−536. (353) Wylie, R. G.; Shoichet, M. S. Two-photon micropatterning of amines within an agarose hydrogel. J. Mater. Chem. 2008, 18, 2716− 2721. (354) Bao, C.; Fan, G.; Lin, Q.; Li, B.; Cheng, S.; Huang, Q.; Zhu, L. Styryl Conjugated Coumarin Caged Alcohol: Efficient Photorelease by Either One-Photon Long Wavelength or Two-Photon NIR Excitation. Org. Lett. 2011, 14, 572−575. (355) Chien, Y.-H.; Chou, Y.-L.; Wang, S.-W.; Hung, S.-T.; Liau, M.C.; Chao, Y.-J.; Su, C.-H.; Yeh, C.-S. Near-Infrared Light Photocontrolled Targeting, Bioimaging, and Chemotherapy with Caged Upconversion Nanoparticles in Vitro and in Vivo. ACS Nano 2013, 7, 8516−8528. 3428

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

(356) Faal, T.; Wong, P.; Tang, S.; Coulter, A.; Chen, Y.; Tu, C. H.; Baker, J. R.; Choi, S. K.; Inlay, M. A. 4-Hydroxytamoxifen probes for light-dependent spatiotemporal control of Cre-ER mediated reporter gene expression. Mol. BioSyst. 2015, 11, 783−790. (357) Gorka, A. P.; Nani, R. R.; Zhu, J.; Mackem, S.; Schnermann, M. J. A Near-IR Uncaging Strategy Based on Cyanine Photochemistry. J. Am. Chem. Soc. 2014, 136, 14153−14159. (358) Billington, A. P.; Walstrom, K. M.; Ramesh, D.; Guzikowski, A. P.; Carpenter, B. K.; Hess, G. P. Synthesis and photochemistry of photolabile N-glycine derivatives and effects of one on the glycine receptor. Biochemistry 1992, 31, 5500−5507. (359) Gee, K. R.; Niu, L.; Schaper, K.; Jayaraman, V.; Hess, G. P. Synthesis and Photochemistry of a Photolabile Precursor of N-Methyld-Aspartate (NMDA) That Is Photolyzed in the Microsecond Time Region and Is Suitable for Chemical Kinetic Investigations of the NMDA Receptor. Biochemistry 1999, 38, 3140−3147. (360) Haase, M.; Schäfer, H. Upconverting Nanoparticles. Angew. Chem., Int. Ed. 2011, 50, 5808−5829. (361) Yang, Y.; Shao, Q.; Deng, R.; Wang, C.; Teng, X.; Cheng, K.; Cheng, Z.; Huang, L.; Liu, Z.; Liu, X.; Xing, B. In Vitro and In Vivo Uncaging and Bioluminescence Imaging by Using Photocaged Upconversion Nanoparticles. Angew. Chem., Int. Ed. 2012, 51, 3125− 3129. (362) Zou, W.; Visser, C.; Maduro, J. A.; Pshenichnikov, M. S.; Hummelen, J. C. Broadband dye-sensitized upconversion of nearinfrared light. Nat. Photonics 2012, 6, 560−564. (363) Anderson, R. R.; Parrish, J. A. The Optics of Human Skin. J. Invest. Dermatol. 1981, 77, 13−19. (364) Poon, L.; Zandberg, W.; Hsiao, D.; Erno, Z.; Sen, D.; Gates, B. D.; Branda, N. R. Photothermal Release of Single-Stranded DNA from the Surface of Gold Nanoparticles Through Controlled Denaturating and Au−S Bond Breaking. ACS Nano 2010, 4, 6395−6403. (365) Riedinger, A.; Guardia, P.; Curcio, A.; Garcia, M. A.; Cingolani, R.; Manna, L.; Pellegrino, T. Subnanometer Local Temperature Probing and Remotely Controlled Drug Release Based on AzoFunctionalized Iron Oxide Nanoparticles. Nano Lett. 2013, 13, 2399− 2406. (366) Jain, A.; Gupta, Y.; Jain, S. K. Azo Chemistry and Its Potential for Colonic Delivery. Crit. Rev. Ther. Drug Carrier Syst. 2006, 23, 349− 400. (367) Chourasia, M. K.; Jain, S. K. Pharmaceutical approaches to colon targeted drug delivery systems. J. Pharm. Pharm. Sci. 2003, 6, 33−66. (368) Klotz, U. The Pharmacological Profile and Clinical Use of Mesalazine (5-Aminosalicylic Acid). Arzneimittelforschung 2012, 62, 53−58. (369) Lakatos, P. L.; Lakatos, L. Once daily 5-aminosalicylic acid for the treatment of ulcerative colitis; are we there yet? Pharmacol. Res. 2008, 58, 190−195. (370) Peppercorn, M. A. Sulfasalazine Pharmacology, Clinical Use, Toxicity, and Related New Drug Development. Ann. Int. Med. 1984, 101, 377−386. (371) Brown, J. P.; McGarraugh, G. V.; Parkinson, T. M.; Wingard, R. E.; Onderdonk, A. B. A polymeric drug for treatment of inflammatory bowel disease. J. Med. Chem. 1983, 26, 1300−1307. (372) Kopečková, P.; Rathi, R.; Takada, S.; Ř íhová, B.; Berenson, M. M.; Kopeček, J. Bioadhesive N-(2-hydroxypropyl) methacrylamide copolymers for colon-specific drug delivery. J. Controlled Release 1994, 28, 211−222. (373) Kopečková, P.; Kopeček, J. Release of 5-aminosalicylic acid from bioadhesive N-(2-hydroxypropyl)methacrylamide copolymers by azoreductases in vitro. Makromol. Chem. 1990, 191, 2037−2045. (374) Sakuma, S.; Lu, Z.-R.; Kopečk ová, P.; Kopeče k, J. Biorecognizable HPMA copolymer−drug conjugates for colon-specific delivery of 9-aminocamptothecin. J. Controlled Release 2001, 75, 365− 379. (375) Gao, Y.; Chen, Y.; Ji, X.; He, X.; Yin, Q.; Zhang, Z.; Shi, J.; Li, Y. Controlled Intracellular Release of Doxorubicin in Multidrug-

Resistant Cancer Cells by Tuning the Shell-Pore Sizes of Mesoporous Silica Nanoparticles. ACS Nano 2011, 5, 9788−9798. (376) Mura, S.; Nicolas, J.; Couvreur, P. Stimuli-responsive nanocarriers for drug delivery. Nat. Mater. 2013, 12, 991−1003. (377) Champion, J.; Walker, A.; Mitragotri, S. Role of Particle Size in Phagocytosis of Polymeric Microspheres. Pharm. Res. 2008, 25, 1815− 1821. (378) Champion, J. A.; Katare, Y. K.; Mitragotri, S. Particle shape: A new design parameter for micro- and nanoscale drug delivery carriers. J. Controlled Release 2007, 121, 3−9. (379) Patri, A. K.; Kukowska-Latallo, J. F.; Baker, J. R., Jr. Targeted drug delivery with dendrimers: Comparison of the release kinetics of covalently conjugated drug and non-covalent drug inclusion complex. Adv. Drug Delivery Rev. 2005, 57, 2203−2214. (380) Fu, F.; Wu, Y.; Zhu, J.; Wen, S.; Shen, M.; Shi, X. Multifunctional Lactobionic Acid-Modified Dendrimers for Targeted Drug Delivery to Liver Cancer Cells: Investigating the Role Played by PEG Spacer. ACS Appl. Mater. Interfaces 2014, 6, 16416−16425. (381) Pacheco, P.; White, D.; Sulchek, T. Effects of Microparticle Size and Fc Density on Macrophage Phagocytosis. PLoS One 2013, 8, e60989. (382) Berchane, N. S.; Carson, K. H.; Rice-Ficht, A. C.; Andrews, M. J. Effect of mean diameter and polydispersity of PLG microspheres on drug release: Experiment and theory. Int. J. Pharm. (Amsterdam, Neth.) 2007, 337, 118−126. (383) Champion, J. A.; Mitragotri, S. Role of target geometry in phagocytosis. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 4930−4934. (384) Zhang, Y.; Tekobo, S.; Tu, Y.; Zhou, Q.; Jin, X.; Dergunov, S. A.; Pinkhassik, E.; Yan, B. Permission to Enter Cell by Shape: Nanodisk vs Nanosphere. ACS Appl. Mater. Interfaces 2012, 4, 4099− 4105. (385) Zhang, Y.; Wu, L.; Jiang, C.; Yan, B. Reprogramming Cellular Signaling Machinery Using Surface-Modified Carbon Nanotubes. Chem. Res. Toxicol. 2015, 28, 296−305. (386) Zhang, M.; Guo, R.; Kéri, M.; Bányai, I.; Zheng, Y.; Cao, M.; Cao, X.; Shi, X. Impact of Dendrimer Surface Functional Groups on the Release of Doxorubicin from Dendrimer Carriers. J. Phys. Chem. B 2014, 118, 1696−1706. (387) Choi, S. K.; Leroueil, P.; Li, M.-H.; Desai, A.; Zong, H.; Van Der Spek, A. F. L.; Baker, J. R., Jr. Specificity and Negative Cooperativity in Dendrimer−Oxime Drug Complexation. Macromolecules 2011, 44, 4026−4029. (388) Choi, S. K.; Thomas, T. P.; Leroueil, P. R.; Kotlyar, A.; Van Der Spek, A. F. L.; Baker, J. R. Specific and Cooperative Interactions between Oximes and PAMAM Dendrimers as Demonstrated by 1H NMR Study. J. Phys. Chem. B 2012, 116, 10387−10397. (389) Hu, J.; Cheng, Y.; Wu, Q.; Zhao, L.; Xu, T. Host-Guest Chemistry of Dendrimer-Drug Complexes. 2. Effects of Molecular Properties of Guests and Surface Functionalities of Dendrimers. J. Phys. Chem. B 2009, 113, 10650−10659. (390) Shi, X.; Macgregor, R. B. The Effect of Charge on the Volume Change of DNA Binding with Intercalator DAPP. J. Phys. Chem. B 2007, 111, 3321−3324. (391) Wang, H.; Shao, N.; Qiao, S.; Cheng, Y. Host−Guest Chemistry of Dendrimer−Cyclodextrin Conjugates: Selective Encapsulations of Guests within Dendrimer or Cyclodextrin Cavities Revealed by NOE NMR Techniques. J. Phys. Chem. B 2012, 116, 11217−11224. (392) Yang, K.; Weng, L.; Cheng, Y.; Zhang, H.; Zhang, J.; Wu, Q.; Xu, T. Host−Guest Chemistry of Dendrimer−Drug Complexes. 6. Fully Acetylated Dendrimers as Biocompatible Drug Vehicles Using Dexamethasone 21-Phosphate as a Model Drug. J. Phys. Chem. B 2011, 115, 2185−2195. (393) Jiang, Y.-Y.; Tang, G.-T.; Zhang, L.-H.; Kong, S.-Y.; Zhu, S.-J.; Pei, Y.-Y. PEGylated PAMAM dendrimers as a potential drug delivery carrier: in vitro and in vivo comparative evaluation of covalently conjugated drug and noncovalent drug inclusion complex. J. Drug Targeting 2010, 18, 389−403. 3429

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

targeted intracellular drug and gene delivery. J. Controlled Release 2011, 152, 2−12. (414) Chen, W.; Du, J. Ultrasound and pH Dually Responsive Polymer Vesicles for Anticancer Drug Delivery. Sci. Rep. 2013, 3, 2162. (415) Rapoport, N. Y.; Kennedy, A. M.; Shea, J. E.; Scaife, C. L.; Nam, K.-H. Controlled and targeted tumor chemotherapy by ultrasound-activated nanoemulsions/microbubbles. J. Controlled Release 2009, 138, 268−276. (416) Louguet, S.; Rousseau, B.; Epherre, R.; Guidolin, N.; Goglio, G.; Mornet, S.; Duguet, E.; Lecommandoux, S.; Schatz, C. Thermoresponsive polymer brush-functionalized magnetic Manganite nanoparticles for remotely triggered drug release. Polym. Chem. 2012, 3, 1408−1417. (417) Zhang, L.; Li, Y.; Yu, J. C. Chemical modification of inorganic nanostructures for targeted and controlled drug delivery in cancer treatment. J. Mater. Chem. B 2014, 2, 452−470. (418) Kim, D.-H.; Guo, Y.; Zhang, Z.; Procissi, D.; Nicolai, J.; Omary, R. A.; Larson, A. C. Temperature-Sensitive Magnetic Drug Carriers for Concurrent Gemcitabine Chemohyperthermia. Adv. Healthcare Mater. 2014, 3, 714−724. (419) Chen, K.-J.; Liang, H.-F.; Chen, H.-L.; Wang, Y.; Cheng, P.-Y.; Liu, H.-L.; Xia, Y.; Sung, H.-W. A Thermoresponsive BubbleGenerating Liposomal System for Triggering Localized Extracellular Drug Delivery. ACS Nano 2013, 7, 438−446. (420) Al-Ahmady, Z. S.; Al-Jamal, W. T.; Bossche, J. V.; Bui, T. T.; Drake, A. F.; Mason, A. J.; Kostarelos, K. Lipid−Peptide Vesicle Nanoscale Hybrids for Triggered Drug Release by Mild Hyperthermia in Vitro and in Vivo. ACS Nano 2012, 6, 9335−9346. (421) Bayer, A. M.; Alam, S.; Mattern-Schain, S. I.; Best, M. D. Triggered Liposomal Release through a Synthetic Phosphatidylcholine Analogue Bearing a Photocleavable Moiety Embedded within the sn-2 Acyl Chain. Chem.Eur. J. 2014, 20, 3350−3357. (422) Mitran, R.-A.; Berger, D.; Băjenaru, L.; Năstase, S.; Andronescu, C.; Matei, C. Azobenzene functionalized mesoporous AlMCM-41-type support for drug release applications. Cent. Eur. J. Chem. 2014, 12, 1−8. (423) Kwon, S.; Singh, R. K.; Perez, R. A.; Abou Neel, E. A.; Kim, H.W.; Chrzanowski, W. Silica-based mesoporous nanoparticles for controlled drug delivery. J. Tissue Eng. 2013, 4. (424) Tarn, D.; Ashley, C. E.; Xue, M.; Carnes, E. C.; Zink, J. I.; Brinker, C. J. Mesoporous Silica Nanoparticle Nanocarriers: Biofunctionality and Biocompatibility. Acc. Chem. Res. 2013, 46, 792−801. (425) Yang, P.; Gai, S.; Lin, J. Functionalized mesoporous silica materials for controlled drug delivery. Chem. Soc. Rev. 2012, 41, 3679− 3698. (426) Slowing, I. I.; Vivero-Escoto, J. L.; Wu, C.-W.; Lin, V. S. Y. Mesoporous silica nanoparticles as controlled release drug delivery and gene transfection carriers. Adv. Drug Delivery Rev. 2008, 60, 1278− 1288. (427) Tang, F.; Li, L.; Chen, D. Mesoporous Silica Nanoparticles: Synthesis, Biocompatibility and Drug Delivery. Adv. Mater. (Weinheim, Ger.) 2012, 24, 1504−1534. (428) Vivero-Escoto, J. L.; Slowing, I. I.; Trewyn, B. G.; Lin, V. S. Y. Mesoporous Silica Nanoparticles for Intracellular Controlled Drug Delivery. Small 2010, 6, 1952−1967. (429) Wu, C.; Wang, Z.; Zhi, Z.; Jiang, T.; Zhang, J.; Wang, S. Development of biodegradable porous starch foam for improving oral delivery of poorly water soluble drugs. Int. J. Pharm. (Amsterdam, Neth.) 2011, 403, 162−169. (430) Berkland, C.; King, M.; Cox, A.; Kim, K.; Pack, D. W. Precise control of PLG microsphere size provides enhanced control of drug release rate. J. Controlled Release 2002, 82, 137−147. (431) Horcajada, P.; Rámila, A.; Pérez-Pariente, J.; Vallet-Regí, M. Influence of pore size of MCM-41 matrices on drug delivery rate. Microporous Mesoporous Mater. 2004, 68, 105−109. (432) Han, L.; Zhao, J.; Zhang, X.; Cao, W.; Hu, X.; Zou, G.; Duan, X.; Liang, X.-J. Enhanced siRNA Delivery and Silencing Gold−

(394) Wong, P. T.; Tang, K.; Coulter, A.; Tang, S.; Baker, J. R.; Choi, S. K. Multivalent Dendrimer Vectors with DNA Intercalation Motifs for Gene Delivery. Biomacromolecules 2014, 15, 4134−4145. (395) Wang, Y.; Guo, R.; Cao, X.; Shen, M.; Shi, X. Encapsulation of 2-methoxyestradiol within multifunctional poly(amidoamine) dendrimers for targeted cancer therapy. Biomaterials 2011, 32, 3322− 3329. (396) Wang, Y.; Cao, X.; Guo, R.; Shen, M.; Zhang, M.; Zhu, M.; Shi, X. Targeted delivery of doxorubicin into cancer cells using a folic aciddendrimer conjugate. Polym. Chem. 2011, 2, 1754−1760. (397) Zhang, M.; Guo, R.; Wang, Y.; Cao, X.; Shen, M.; Shi, X. Multifunctional dendrimer/combretastatin A4 inclusion complexes enable in vitro targeted cancer therapy. Int. J. Nanomed. 2011, 6, 2337−2349. (398) Fant, K.; Esbjörner, E. K.; Jenkins, A.; Grossel, M. C.; Lincoln, P.; Nordén, B. Effects of PEGylation and Acetylation of PAMAM Dendrimers on DNA Binding, Cytotoxicity and in Vitro Transfection Efficiency. Mol. Pharmaceutics 2010, 7, 1734−1746. (399) He, H.; Li, Y.; Jia, X.-R.; Du, J.; Ying, X.; Lu, W.-L.; Lou, J.-N.; Wei, Y. PEGylated Poly(amidoamine) dendrimer-based dual-targeting carrier for treating brain tumors. Biomaterials 2011, 32, 478−487. (400) Zhang, Y.; Kohler, N.; Zhang, M. Surface modification of superparamagnetic magnetite nanoparticles and their intracellular uptake. Biomaterials 2002, 23, 1553−1561. (401) Shi, X.; Lee, I.; Chen, X.; Shen, M.; Xiao, S.; Zhu, M.; Baker, J. R.; Wang, S. H. Influence of dendrimer surface charge on the bioactivity of 2-methoxyestradiol complexed with dendrimers. Soft Matter 2010, 6, 2539−2545. (402) Zhu, J.; Shi, X. Dendrimer-based nanodevices for targeted drug delivery applications. J. Mater. Chem. B 2013, 1, 4199−4211. (403) Yassin, M. A.; Appelhans, D.; Mendes, R. G.; Rümmeli, M. H.; Voit, B. pH-Dependent Release of Doxorubicin from Fast PhotoCross-Linkable Polymersomes Based on Benzophenone Units. Chem.Eur. J. 2012, 18, 12227−12231. (404) Lee, S.-M.; Lee, O.-S.; O’Halloran, T. V.; Schatz, G. C.; Nguyen, S. T. Triggered Release of Pharmacophores from [Ni(HAsO3)]-Loaded Polymer-Caged Nanobin Enhances Pro-apoptotic Activity: A Combined Experimental and Theoretical Study. ACS Nano 2011, 5, 3961−3969. (405) Liu, R.; Li, D.; He, B.; Xu, X.; Sheng, M.; Lai, Y.; Wang, G.; Gu, Z. Anti-tumor drug delivery of pH-sensitive poly(ethylene glycol)poly(L-histidine-)-poly(L-lactide) nanoparticles. J. Controlled Release 2011, 152, 49−56. (406) Ju, C.; Mo, R.; Xue, J.; Zhang, L.; Zhao, Z.; Xue, L.; Ping, Q.; Zhang, C. Sequential Intra-Intercellular Nanoparticle Delivery System for Deep Tumor Penetration. Angew. Chem., Int. Ed. 2014, 6253− 6258. (407) Zhang, A.; Zhang, Z.; Shi, F.; Xiao, C.; Ding, J.; Zhuang, X.; He, C.; Chen, L.; Chen, X. Redox-Sensitive Shell-Crosslinked Polypeptide-block-Polysaccharide Micelles for Efficient Intracellular Anticancer Drug Delivery. Macromol. Biosci. 2013, 13, 1249−1258. (408) Kommareddy, S.; Amiji, M. Poly(ethylene glycol)-modified thiolated gelatin nanoparticles for glutathione-responsive intracellular DNA delivery. Nanomedicine (N. Y., NY, U. S.) 2007, 3, 32−42. (409) Qiao, Z.-Y.; Zhang, R.; Du, F.-S.; Liang, D.-H.; Li, Z.-C. Multiresponsive nanogels containing motifs of ortho ester, oligo(ethylene glycol) and disulfide linkage as carriers of hydrophobic anti-cancer drugs. J. Controlled Release 2011, 152, 57−66. (410) Kim, T.-i.; Rothmund, T.; Kissel, T.; Kim, S. W. Bioreducible polymers with cell penetrating and endosome buffering functionality for gene delivery systems. J. Controlled Release 2011, 152, 110−119. (411) Jackson, A. W.; Fulton, D. A. Making polymeric nanoparticles stimuli-responsive with dynamic covalent bonds. Polym. Chem. 2013, 4, 31−45. (412) Jiao, Y.; Sun, Y.; Chang, B.; Lu, D.; Yang, W. Redox- and Temperature-Controlled Drug Release from Hollow Mesoporous Silica Nanoparticles. Chem.Eur. J. 2013, 19, 15410−15420. (413) Cheng, R.; Feng, F.; Meng, F.; Deng, C.; Feijen, J.; Zhong, Z. Glutathione-responsive nano-vehicles as a promising platform for 3430

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

Triggered Self-Assembly of a Genetically Encoded Polymer. ACS Nano 2010, 4, 2217−2227. (452) Dayton, P. A.; Zhao, S.; Bloch, S. H.; Schumann, P.; Penrose, K.; Matsunaga, T. O.; Zutshi, R.; Doinikov, A.; Ferrara, K. W. Application of ultrasound to selectively localize nanodroplets for targeted imaging and therapy. Mol. Imaging 2006, 5, 160−174. (453) Bhattacharyya, S.; Wang, H.; Ducheyne, P. Polymer-coated mesoporous silica nanoparticles for the controlled release of macromolecules. Acta Biomater. 2012, 8, 3429−3435. (454) Liu, J.; Bu, W.; Pan, L.; Shi, J. NIR-Triggered Anticancer Drug Delivery by Upconverting Nanoparticles with Integrated AzobenzeneModified Mesoporous Silica. Angew. Chem., Int. Ed. 2013, 52, 4375− 4379. (455) Li, M.; Tang, Z.; Sun, H.; Ding, J.; Song, W.; Chen, X. pH and reduction dual-responsive nanogel cross-linked by quaternization reaction for enhanced cellular internalization and intracellular drug delivery. Polym. Chem. 2013, 4, 1199−1207. (456) Xu, X.; Flores, J. D.; McCormick, C. L. Reversible Imine Shell Cross-Linked Micelles from Aqueous RAFT-Synthesized Thermoresponsive Triblock Copolymers as Potential Nanocarriers for “pHTriggered” Drug Release. Macromolecules 2011, 44, 1327−1334. (457) He, L.; Huang, Y.; Zhu, H.; Pang, G.; Zheng, W.; Wong, Y.-S.; Chen, T. Cancer-Targeted Monodisperse Mesoporous Silica Nanoparticles as Carrier of Ruthenium Polypyridyl Complexes to Enhance Theranostic Effects. Adv. Funct. Mater. 2014, 24, 2754−2763. (458) Chiang, W.-H.; Huang, W.-C.; Shen, M.-Y.; Wang, C.-H.; Huang, Y.-F.; Lin, S.-C.; Chern, C.-S.; Chiu, H.-C. Dual-Layered Nanogel-Coated Hollow Lipid/Polypeptide Conjugate Assemblies for Potential pH-Triggered Intracellular Drug Release. PLoS One 2014, 9, e92268. (459) Chen, X.; Cheng, X.; Soeriyadi, A. H.; Sagnella, S. M.; Lu, X.; Scott, J. A.; Lowe, S. B.; Kavallaris, M.; Gooding, J. J. Stimuliresponsive functionalized mesoporous silica nanoparticles for drug release in response to various biological stimuli. Biomater. Sci. 2014, 2, 121−130. (460) Chen, X.; Chen, L.; Yao, X.; Zhang, Z.; He, C.; Zhang, J.; Chen, X. Dual responsive supramolecular nanogels for intracellular drug delivery. Chem. Commun. (Cambridge, U. K.) 2014, 50, 3789− 3791. (461) Wang, K.; Luo, G.-F.; Liu, Y.; Li, C.; Cheng, S.-X.; Zhuo, R.-X.; Zhang, X.-Z. Redox-sensitive shell cross-linked PEG-polypeptide hybrid micelles for controlled drug release. Polym. Chem. 2012, 3, 1084−1090. (462) Yu, S.; Ding, J.; He, C.; Cao, Y.; Xu, W.; Chen, X. Disulfide Cross-Linked Polyurethane Micelles as a Reduction-Triggered Drug Delivery System for Cancer Therapy. Adv. Healthcare Mater. 2014, 3, 752−760. (463) Tao, Y.; Han, J.; Ye, C.; Thomas, T.; Dou, H. Reductionresponsive gold-nanoparticle-conjugated Pluronic micelles: an effective anti-cancer drug delivery system. J. Mater. Chem. 2012, 22, 18864− 18871. (464) Gui, R.; Wan, A.; Liu, X.; Jin, H. Intracellular fluorescent thermometry and photothermal-triggered drug release developed from gold nanoclusters and doxorubicin dual-loaded liposomes. Chem. Commun. (Cambridge, U. K.) 2014, 50, 1546−1548. (465) Wang, G.; Tong, X.; Zhao, Y. Preparation of AzobenzeneContaining Amphiphilic Diblock Copolymers for Light-Responsive Micellar Aggregates. Macromolecules 2004, 37, 8911−8917. (466) Kurapati, R.; Raichur, A. M. Near-infrared light-responsive graphene oxide composite multilayer capsules: a novel route for remote controlled drug delivery. Chem. Commun. (Cambridge, U. K.) 2013, 49, 734−736. (467) López-Noriega, A.; Hastings, C. L.; Ozbakir, B.; O’Donnell, K. E.; O’Brien, F. J.; Storm, G.; Hennink, W. E.; Duffy, G. P.; RuizHernández, E. Hyperthermia-Induced Drug Delivery from Thermosensitive Liposomes Encapsulated in an Injectable Hydrogel for Local Chemotherapy. Adv. Healthcare Mater. 2014, 3, 854−859. (468) Bilalis, P.; Efthimiadou, E. K.; Chatzipavlidis, A.; Boukos, N.; Kordas, G. C. Multi-responsive polymeric microcontainers for

Chitosan Nanosystem with Surface Charge-Reversal Polymer Assembly and Good Biocompatibility. ACS Nano 2012, 6, 7340−7351. (433) Ruiz-Hernández, E.; Baeza, A.; Vallet-Regí, M. Smart Drug Delivery through DNA/Magnetic Nanoparticle Gates. ACS Nano 2011, 5, 1259−1266. (434) Dunn, A. E.; Dunn, D. J.; Macmillan, A.; Whan, R.; StaitGardner, T.; Price, W. S.; Lim, M.; Boyer, C. Spatial and temporal control of drug release through pH and alternating magnetic field induced breakage of Schiff base bonds. Polym. Chem. 2014, 5, 3311− 3315. (435) Satarkar, N. S.; Hilt, J. Z. Magnetic hydrogel nanocomposites for remote controlled pulsatile drug release. J. Controlled Release 2008, 130, 246−251. (436) Thomas, C. R.; Ferris, D. P.; Lee, J.-H.; Choi, E.; Cho, M. H.; Kim, E. S.; Stoddart, J. F.; Shin, J.-S.; Cheon, J.; Zink, J. I. Noninvasive Remote-Controlled Release of Drug Molecules in Vitro Using Magnetic Actuation of Mechanized Nanoparticles. J. Am. Chem. Soc. 2010, 132, 10623−10625. (437) Satarkar, N. S.; Zhang, W.; Eitel, R. E.; Hilt, J. Z. Magnetic hydrogel nanocomposites as remote controlled microfluidic valves. Lab Chip 2009, 9, 1773−1779. (438) Rydzek, G.; Schaaf, P.; Voegel, J.-C.; Jierry, L.; Boulmedais, F. Strategies for covalently reticulated polymer multilayers. Soft Matter 2012, 8, 9738−9755. (439) Wong, S.; Shim, M. S.; Kwon, Y. J. Synthetically designed peptide-based biomaterials with stimuli-responsive and membraneactive properties for biomedical applications. J. Mater. Chem. B 2014, 2, 595−615. (440) Lai, T. C.; Cho, H.; Kwon, G. S. Reversibly core cross-linked polymeric micelles with pH- and reduction-sensitivities: effects of cross-linking degree on particle stability, drug release kinetics, and antitumor efficacy. Polym. Chem. 2014, 5, 1650−1661. (441) Koo, A. N.; Lee, H. J.; Kim, S. E.; Chang, J. H.; Park, C.; Kim, C.; Park, J. H.; Lee, S. C. Disulfide-cross-linked PEG-poly(amino acid)s copolymer micelles for glutathione-mediated intracellular drug delivery. Chem. Commun. (Cambridge, U. K.) 2008, 6570−6572. (442) Gohy, J.-F.; Zhao, Y. Photo-responsive block copolymer micelles: design and behavior. Chem. Soc. Rev. 2013, 42, 7117−7129. (443) Hawkins, M. J.; Soon-Shiong, P.; Desai, N. Protein nanoparticles as drug carriers in clinical medicine. Adv. Drug Delivery Rev. 2008, 60, 876−885. (444) Gradishar, W. J. Albumin-bound paclitaxel: a next-generation taxane. Expert Opin. Pharmacother. 2006, 7, 1041−1053. (445) Papahadjopoulos, D.; Allen, T. M.; Gabizon, A.; Mayhew, E.; Matthay, K.; Huang, S. K.; Lee, K. D.; Woodle, M. C.; Lasic, D. D.; Redemann, C. Sterically stabilized liposomes: improvements in pharmacokinetics and antitumor therapeutic efficacy. Proc. Natl. Acad. Sci. U. S. A. 1991, 88, 11460−11464. (446) Lee, E. S.; Na, K.; Bae, Y. H. Doxorubicin loaded pH-sensitive polymeric micelles for reversal of resistant MCF-7 tumor. J. Controlled Release 2005, 103, 405−418. (447) Nakanishi, T.; Fukushima, S.; Okamoto, K.; Suzuki, M.; Matsumura, Y.; Yokoyama, M.; Okano, T.; Sakurai, Y.; Kataoka, K. Development of the polymer micelle carrier system for doxorubicin. J. Controlled Release 2001, 74, 295−302. (448) Kim, D.; Lee, E. S.; Oh, K. T.; Gao, Z. G.; Bae, Y. H. Doxorubicin-Loaded Polymeric Micelle Overcomes Multidrug Resistance of Cancer by Double-Targeting Folate Receptor and Early Endosomal pH. Small 2008, 4, 2043−2050. (449) Nelson, J. L.; Roeder, B. L.; Carmen, J. C.; Roloff, F.; Pitt, W. G. Ultrasonically Activated Chemotherapeutic Drug Delivery in a Rat Model. Cancer Res. 2002, 62, 7280−7283. (450) Liu, W.; Dreher, M. R.; Furgeson, D. Y.; Peixoto, K. V.; Yuan, H.; Zalutsky, M. R.; Chilkoti, A. Tumor accumulation, degradation and pharmacokinetics of elastin-like polypeptides in nude mice. J. Controlled Release 2006, 116, 170−178. (451) Simnick, A. J.; Valencia, C. A.; Liu, R.; Chilkoti, A. Morphing Low-Affinity Ligands into High-Avidity Nanoparticles by Thermally 3431

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Chemical Reviews

Review

potential biomedical applications: synthesis and functionality evaluation. Polym. Int. 2012, 61, 888−894. (469) Lee, S.-F.; Zhu, X.-M.; Wang, Y.-X. J.; Xuan, S.-H.; You, Q.; Chan, W.-H.; Wong, C.-H.; Wang, F.; Yu, J. C.; Cheng, C. H. K.; Leung, K. C.-F. Ultrasound, pH, and Magnetically Responsive CrownEther-Coated Core/Shell Nanoparticles as Drug Encapsulation and Release Systems. ACS Appl. Mater. Interfaces 2013, 5, 1566−1574. (470) Li, C.; Yu, D.-F.; Newman, R. A.; Cabral, F.; Stephens, L. C.; Hunter, N.; Milas, L.; Wallace, S. Complete Regression of Wellestablished Tumors Using a Novel Water-soluble Poly(l-Glutamic Acid)-Paclitaxel Conjugate. Cancer Res. 1998, 58, 2404−2409. (471) Li, C.; Wallace, S. Polymer-drug conjugates: Recent development in clinical oncology. Adv. Drug Delivery Rev. 2008, 60, 886−898. (472) Macromolecular Anticancer Therapeutics; Reddy, L. H., Couvreur, P., Eds.; Humana Press: New York, 2010. (473) Meerum Terwogt, J. M.; ten Bokkel Huinink, W. W.; Schellens, J. H.; Schot, M.; Mandjes, I. A.; Zurlo, M. G.; Rocchetti, M.; Rosing, H.; Koopman, F. J.; Beijnen, J. H. Phase I clinical and pharmacokinetic study of PNU166945, a novel water-soluble polymerconjugated prodrug of paclitaxel. Anti-Cancer Drugs 2001, 12, 315− 323. (474) Duncan, R. Development of HPMA copolymer−anticancer conjugates: Clinical experience and lessons learnt. Adv. Drug Delivery Rev. 2009, 61, 1131−1148. (475) Canal, F.; Sanchis, J.; Vicent, M. J. Polymer−drug conjugates as nano-sized medicines. Curr. Opin. Biotechnol. 2011, 22, 894−900. (476) Yurkovetskiy, A. V.; Fram, R. J. XMT-1001, a novel polymeric camptothecin pro-drug in clinical development for patients with advanced cancer. Adv. Drug Delivery Rev. 2009, 61, 1193−1202. (477) Schoemaker, N. E.; van Kesteren, C.; Rosing, H.; Jansen, S.; Swart, M.; Lieverst, J.; Fraier, D.; Breda, M.; Pellizzoni, C.; Spinelli, R.; Porro, M. G.; Beijnen, J. H.; Schellens, J. H. M.; ten Bokkel Huinink, W. W. A phase I and pharmacokinetic study of MAG-CPT, a watersoluble polymer conjugate of camptothecin. Br. J. Cancer 2002, 87, 608−614. (478) Young, C.; Schluep, T.; Hwang, J.; Eliasof, S. CRLX101 (formerly IT-101)A Novel Nanopharmaceutical of Camptothecin in Clinical Development. Curr. Bioact. Compd. 2011, 7, 8−14. (479) Veltkamp, S. A.; Witteveen, E. O.; Capriati, A.; Crea, A.; Animati, F.; Voogel-Fuchs, M.; van den Heuvel, I. J. G. M.; Beijnen, J. H.; Voest, E. E.; Schellens, J. H. M. Clinical and Pharmacologic Study of the Novel Prodrug Delimotecan (MEN 4901/T-0128) in Patients with Solid Tumors. Clin. Cancer Res. 2008, 14, 7535−7544. (480) Seymour, L. W.; Ferry, D. R.; Anderson, D.; Hesslewood, S.; Julyan, P. J.; Poyner, R.; Doran, J.; Young, A. M.; Burtles, S.; Kerr, D. J.; et al. Hepatic Drug Targeting: Phase I Evaluation of PolymerBound Doxorubicin. J. Clin. Oncol. 2002, 20, 1668−1676. (481) Vasey, P. A.; Kaye, S. B.; Morrison, R.; Twelves, C.; Wilson, P.; Duncan, R.; Thomson, A. H.; Murray, L. S.; Hilditch, T. E.; Murray, T.; Burtles, S.; Fraier, D.; Frigerio, E.; Cassidy, J. Phase I Clinical and Pharmacokinetic Study of PK1 [N-(2-Hydroxypropyl)methacrylamide Copolymer Doxorubicin]: First Member of a New Class of Chemotherapeutic AgentsDrug-Polymer Conjugates. Clin. Cancer Res. 1999, 5, 83−94. (482) Yokoyama, M.; Okano, T.; Sakurai, Y.; Ekimoto, H.; Shibazaki, C.; Kataoka, K. Toxicity and Antitumor Activity against Solid Tumors of Micelle-forming Polymeric Anticancer Drug and Its Extremely Long Circulation in Blood. Cancer Res. 1991, 51, 3229−3236. (483) Kwon, G.; Suwa, S.; Yokoyama, M.; Okano, T.; Sakurai, Y.; Kataoka, K. Enhanced tumor accumulation and prolonged circulation times of micelle-forming poly (ethylene oxide-aspartate) block copolymer-adriamycin conjugates. J. Controlled Release 1994, 29, 17− 23. (484) Kataoka, K.; Matsumoto, T.; Yokoyama, M.; Okano, T.; Sakurai, Y.; Fukushima, S.; Okamoto, K.; Kwon, G. S. Doxorubicinloaded poly(ethylene glycol)−poly(β-benzyl-l-aspartate) copolymer micelles: their pharmaceutical characteristics and biological significance. J. Controlled Release 2000, 64, 143−153.

(485) Gianasi, E.; Buckley, R. G.; Latigo, J.; Wasil, M.; Duncan, R. HPMA Copolymers Platinates Containing Dicarboxylato Ligands. Preparation, Characterisation and In Vitro and In Vivo Evaluation. J. Drug Targeting 2002, 10, 549−556. (486) Duncan, R. The dawning era of polymer therapeutics. Nat. Rev. Drug Discovery 2003, 2, 347−360. (487) Bolling, C.; Graefe, T.; Lübbing, C.; Jankevicius, F.; Uktveris, S.; Cesas, A.; Meyer-Moldenhauer, W. H.; Starkmann, H.; Weigel, M.; Burk, K.; Hanauske, A. R. Phase II study of MTX-HSA in combination with Cisplatin as first line treatment in patients with advanced or metastatic transitional cell carcinoma. Invest. New Drugs 2006, 24, 521−527. (488) Wosikowski, K.; Biedermann, E.; Rattel, B.; Breiter, N.; Jank, P.; Löser, R.; Jansen, G.; Peters, G. J. In Vitro and in Vivo Antitumor Activity of Methotrexate Conjugated to Human Serum Albumin in Human Cancer Cells. Clin. Cancer Res. 2003, 9, 1917−1926. (489) Yurkovetskiy, A.; Cabral, C. B.; Meyers, L. L.; Fram, R. J.; Lowinger, T. B. Cancer Res. 2010, 70 (Suppl. 8). Proceedings of AACR 101st Annual Meeting, Washington, DC, April 19−21, 2010. (490) Vicent, M. J.; Ringsdorf, H.; Duncan, R. Polymer therapeutics: Clinical applications and challenges for development. Adv. Drug Delivery Rev. 2009, 61, 1117−1120. (491) Hlavacek, W. S.; Posner, R. G.; Perelson, A. S. Steric Effects on Multivalent Ligand-Receptor Binding: Exclusion of Ligand Sites by Bound Cell Surface Receptors. Biophys. J. 1999, 76, 3031−3043. (492) Witte, A. B.; Timmer, C. M.; Gam, J. J.; Choi, S. K.; Banaszak Holl, M. M.; Orr, B. G.; Baker, J. R.; Sinniah, K. Biophysical Characterization of a Riboflavin-conjugated Dendrimer Platform for Targeted Drug Delivery. Biomacromolecules 2012, 13, 507−516. (493) Silpe, J. E.; Sumit, M.; Thomas, T. P.; Huang, B.; Kotlyar, A.; van Dongen, M. A.; Banaszak Holl, M. M.; Orr, B. G.; Choi, S. K. Avidity Modulation of Folate-Targeted Multivalent Dendrimers for Evaluating Biophysical Models of Cancer Targeting Nanoparticles. ACS Chem. Biol. 2013, 8, 2063−2071. (494) Longmire, M.; Choyke, P. L.; Kobayashi, H. Clearance properties of nano-sized particles and molecules as imaging agents: considerations and caveats. Nanomedicine (London, U. K.) 2008, 3, 703−717. (495) Choi, H. S.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J. P.; Itty Ipe, B.; Bawendi, M. G.; Frangioni, J. V. Renal clearance of quantum dots. Nat. Biotechnol. 2007, 25, 1165−1170.

3432

DOI: 10.1021/cr5004634 Chem. Rev. 2015, 115, 3388−3432

Mechanisms of drug release in nanotherapeutic delivery systems.

Mechanisms of drug release in nanotherapeutic delivery systems. - PDF Download Free
8MB Sizes 3 Downloads 10 Views