acetal-linked paclitaxel prodrug micellar nanoparticles as a versatile and potent platform for...

9
Acetal-Linked Paclitaxel Prodrug Micellar Nanoparticles as a Versatile and Potent Platform for Cancer Therapy Yudan Gu, Yinan Zhong, Fenghua Meng,* Ru Cheng, Chao Deng, and Zhiyuan Zhong* Biomedical Polymers Laboratory and Jiangsu Key Laboratory of Advanced Functional Polymer Design and Application, Department of Polymer Science and Engineering, College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou 215123, P. R. China ABSTRACT: Endosomal pH-activatable paclitaxel (PTX) prodrug micellar nanoparticles were designed and prepared by conjugating PTX onto water-soluble poly(ethylene glycol)- b-poly(acrylic acid) (PEG-PAA) block copolymers via an acid- labile acetal bond to the PAA block and investigated for potent growth inhibition of human cancer cells in vitro. PTX was readily conjugated to PEG-PAA with high drug contents of 21.6, 27.0, and 42.8 wt % (denoted as PTX prodrugs 1, 2, and 3, respectively) using ethyl glycol vinyl ether (EGVE) as a linker. The resulting PTX conjugates had dened molecular weights and self-assembled in phosphate buer (PB, pH 7.4, 10 mM) into monodisperse micellar nanoparticles with average sizes of 158.3180.3 nm depending on PTX contents. The in vitro release studies showed that drug release from PTX prodrug nanoparticles was highly pH-dependent, in which ca. 86.9%, 66.4% and 29.0% of PTX was released from PTX prodrug 3 at 37 °C in 48 h at pH 5.0, 6.0, and pH 7.4, respectively. MTT assays showed that these pH-sensitive PTX prodrug nanoparticles exhibited high antitumor eect to KB and HeLa cells (IC 50 = 0.18 and 0.9 μg PTX equiv/mL, respectively) as well as PTX-resistant A549 cells. Notably, folate-decorated PTX prodrug micellar nanoparticles based on PTX prodrug 3 and 20 wt % folate-poly(ethylene glycol)-b-poly(D,L-lactide) (FA-PEG-PLA) displayed apparent targetability to folate receptor-overexpressing KB cells with IC 50 over 12 times lower than nontargeting PTX prodrug 3 under otherwise the same conditions. Furthermore, PTX prodrug nanoparticles could also load doxorubicin (DOX) to simultaneously release PTX and DOX under mildly acidic pH. These acetal- linked PTX prodrug micellar nanoparticles have appeared as a highly versatile and potent platform for cancer therapy. INTRODUCTION The lack of suitable vehicles is a major challenge for systemic administration of potent and yet poorly water-soluble anticancer drugs including paclitaxel (PTX) and doxorubicin (DOX). 14 The current clinical PTX formulations based on Cremophor EL and ethanol (Taxol) though improved solubility of PTX are associated with hypersensitivity reactions. 5 In the past decades, various biocompatible nanosystems such as polymeric prodrugs, 6 liposomes, 7 polymeric micelles, 8,9 polymersomes, 10 and poly- meric nanoparticles 11,12 have been developed for safer as well as better-controlled delivery of PTX. The in vivo studies have shown that nanosystems exhibit improved drug tolerance, better drug bioavailability, and apparent tumor-targetability via enhanced permeability and retention (EPR) eect (passive targeting). 13,14 In particular, polymeric prodrugs and micellar nanoparticles have received the most attention. For example, PTX prodrugs have been prepared by conjugating PTX to water- soluble polymers such as poly(ethylene glycol) (PEG), poly(N- (2-hydroxypropyl)-methacrylamide) (PHPMA), poly(L-gluta- mic acid) (PGlu) and hyaluronic acid via a cleavable peptide, ester or phosphonate bond. 1517 It should be noted that PHPMA-PTX (PNU166945) 18 and PGlu-PTX (CT-2103, Xyotax) 19,20 have progressed to phase I and III clinical trials, respectively. However, polymeric prodrugs are often plagued by complex synthesis, low drug conjugation, small size, as well as inferior drug activity due to slow release of drug at the site of action and/or chemical alternation of released drug. In comparison, coreshell micelles with sizes of about 20200 nm load PTX via physical encapsulation with generally higher drug loading contents. 21,22 Notably, micellar PTX formulations, Genexol-PM 23 and NK105, 24 have been approved for treating breast, lung, and ovarian cancers in South Korea and for phase III clinical trials in patients with metastatic or recurrent breast cancers, respectively. The self-assembled micellar nanoparticles, nevertheless, suer poor in vivo stability, premature drug release following injection, and signicant systemic side eects. 4 To combine advantageous features of polymeric prodrugs and micellar nanoparticles, prodrug micellar nanoparticles have been conceived based on amphiphilic block copolymer-drug con- jugates that contain stimuli-sensitive degradable hydrazone, ester or disulde bonds. 2530 These prodrug micellar nanoparticles have shown improved systemic stability, longer circulation time, and better tumor targetability. For instance, Jing et al. reported that P(LGG-PTX)-PEG-P(LGG-PTX) prodrug micelles by conjugating PTX via an ester bond did not show initial burst drug release and kept the in vitro antitumor activity of PTX Received: April 30, 2013 Revised: June 18, 2013 Published: June 18, 2013 Article pubs.acs.org/Biomac © 2013 American Chemical Society 2772 dx.doi.org/10.1021/bm400615n | Biomacromolecules 2013, 14, 27722780

Upload: zhiyuan

Post on 09-Dec-2016

213 views

Category:

Documents


0 download

TRANSCRIPT

Acetal-Linked Paclitaxel Prodrug Micellar Nanoparticles as a Versatileand Potent Platform for Cancer TherapyYudan Gu, Yinan Zhong, Fenghua Meng,* Ru Cheng, Chao Deng, and Zhiyuan Zhong*

Biomedical Polymers Laboratory and Jiangsu Key Laboratory of Advanced Functional Polymer Design and Application, Departmentof Polymer Science and Engineering, College of Chemistry, Chemical Engineering and Materials Science, Soochow University,Suzhou 215123, P. R. China

ABSTRACT: Endosomal pH-activatable paclitaxel (PTX)prodrug micellar nanoparticles were designed and preparedby conjugating PTX onto water-soluble poly(ethylene glycol)-b-poly(acrylic acid) (PEG-PAA) block copolymers via an acid-labile acetal bond to the PAA block and investigated for potentgrowth inhibition of human cancer cells in vitro. PTX wasreadily conjugated to PEG-PAA with high drug contents of21.6, 27.0, and 42.8 wt % (denoted as PTX prodrugs 1, 2, and3, respectively) using ethyl glycol vinyl ether (EGVE) as alinker. The resulting PTX conjugates had defined molecularweights and self-assembled in phosphate buffer (PB, pH 7.4, 10 mM) into monodisperse micellar nanoparticles with average sizesof 158.3−180.3 nm depending on PTX contents. The in vitro release studies showed that drug release from PTX prodrugnanoparticles was highly pH-dependent, in which ca. 86.9%, 66.4% and 29.0% of PTX was released from PTX prodrug 3 at 37 °Cin 48 h at pH 5.0, 6.0, and pH 7.4, respectively. MTT assays showed that these pH-sensitive PTX prodrug nanoparticles exhibitedhigh antitumor effect to KB and HeLa cells (IC50 = 0.18 and 0.9 μg PTX equiv/mL, respectively) as well as PTX-resistant A549cells. Notably, folate-decorated PTX prodrug micellar nanoparticles based on PTX prodrug 3 and 20 wt % folate-poly(ethyleneglycol)-b-poly(D,L-lactide) (FA-PEG-PLA) displayed apparent targetability to folate receptor-overexpressing KB cells with IC50over 12 times lower than nontargeting PTX prodrug 3 under otherwise the same conditions. Furthermore, PTX prodrugnanoparticles could also load doxorubicin (DOX) to simultaneously release PTX and DOX under mildly acidic pH. These acetal-linked PTX prodrug micellar nanoparticles have appeared as a highly versatile and potent platform for cancer therapy.

■ INTRODUCTIONThe lack of suitable vehicles is a major challenge for systemicadministration of potent and yet poorly water-soluble anticancerdrugs including paclitaxel (PTX) and doxorubicin (DOX).1−4

The current clinical PTX formulations based on Cremophor ELand ethanol (Taxol) though improved solubility of PTX areassociated with hypersensitivity reactions.5 In the past decades,various biocompatible nanosystems such as polymeric prodrugs,6

liposomes,7 polymeric micelles,8,9 polymersomes,10 and poly-meric nanoparticles11,12 have been developed for safer as well asbetter-controlled delivery of PTX. The in vivo studies haveshown that nanosystems exhibit improved drug tolerance, betterdrug bioavailability, and apparent tumor-targetability viaenhanced permeability and retention (EPR) effect (passivetargeting).13,14 In particular, polymeric prodrugs and micellarnanoparticles have received the most attention. For example,PTX prodrugs have been prepared by conjugating PTX to water-soluble polymers such as poly(ethylene glycol) (PEG), poly(N-(2-hydroxypropyl)-methacrylamide) (PHPMA), poly(L-gluta-mic acid) (PGlu) and hyaluronic acid via a cleavable peptide,ester or phosphonate bond.15−17 It should be noted thatPHPMA-PTX (PNU166945)18 and PGlu-PTX (CT-2103,Xyotax)19,20 have progressed to phase I and III clinical trials,respectively. However, polymeric prodrugs are often plagued bycomplex synthesis, low drug conjugation, small size, as well as

inferior drug activity due to slow release of drug at the site ofaction and/or chemical alternation of released drug. Incomparison, core−shell micelles with sizes of about 20−200nm load PTX via physical encapsulation with generally higherdrug loading contents.21,22 Notably, micellar PTX formulations,Genexol-PM23 and NK105,24 have been approved for treatingbreast, lung, and ovarian cancers in South Korea and for phase IIIclinical trials in patients with metastatic or recurrent breastcancers, respectively. The self-assembled micellar nanoparticles,nevertheless, suffer poor in vivo stability, premature drug releasefollowing injection, and significant systemic side effects.4

To combine advantageous features of polymeric prodrugs andmicellar nanoparticles, prodrug micellar nanoparticles have beenconceived based on amphiphilic block copolymer-drug con-jugates that contain stimuli-sensitive degradable hydrazone, esteror disulfide bonds.25−30 These prodrug micellar nanoparticleshave shown improved systemic stability, longer circulation time,and better tumor targetability. For instance, Jing et al. reportedthat P(LGG-PTX)-PEG-P(LGG-PTX) prodrug micelles byconjugating PTX via an ester bond did not show initial burstdrug release and kept the in vitro antitumor activity of PTX

Received: April 30, 2013Revised: June 18, 2013Published: June 18, 2013

Article

pubs.acs.org/Biomac

© 2013 American Chemical Society 2772 dx.doi.org/10.1021/bm400615n | Biomacromolecules 2013, 14, 2772−2780

against RBG-6 cells.31 Ulbrich et al. reported that PHPMA-hydrazide-LEV-PTX prodrug micelles with levulic acid-modifiedPTX linked to PHPMA via hydrazone bonds showed pH-dependent PTX release and better antitumor efficacy in the 4T1model of mammary carcinoma than the parent free drugs.32 Itshould be noted, nevertheless, that the in vitro antitumor activityof PTX prodrug micellar nanoparticles is often reduced asreleased drugs are structurally altered and drug release insidetumor cells is further hindered by hydrophobic nature of micellarcore.In this Article, we report on acetal-linked PTX prodrug

micellar nanoparticles that display superior drug content,endosomal pH-triggered rapid drug release and potent in vitroantitumor activity to drug-sensitive as well as resistant cancercells (Scheme 1). These PTX prodrug micellar nanoparticleswere designed and developed by conjugating PTX onto water-soluble poly(ethylene glycol)-b-poly(acrylic acid) (PEG-PAA)block copolymers via an acid-labile acetal bond to PAA blockusing ethyl glycol vinyl ether as a linker. In the past years, variouspH-sensitive nanoparticles have been developed based on acetal-containing amphiphilic copolymers for loading and triggeredrelease of poorly water-soluble anticancer drugs including DOXand PTX.33−36 These acetal-based delivery systems have severaladvantages such as fast degradation under endo/lysosomal pHconditions, absence of acidic degradation products, and versatilepreparation and applications. To the best of our knowledge,acetal-linked PTX prodrug micellar nanoparticles that releasePTX in its pristine nature have not yet been reported. Here,preparation of acetal-linked PTX prodrugs and micellar nano-particles, pH-triggered drug release, and in vitro antitumoractivity of PTX prodrug nanoparticles to KB and HeLa cells aswell as PTX-resistant A549 cells, receptor-mediated delivery offolate-decorated PTX prodrug nanoparticles, and pH-sensitiverelease of PTX and physically loaded DOX were investigated.

■ EXPERIMENTAL SECTIONMaterials. Acrylic acid (AA, 99%, Alfa Aesar) was purified by passing

through a basic alumina column. S-1-dodecyl-S-(R,R-dimethyl-R-aceticacid)trithiocarbonate (DMP),37 PEG-DMP (Mn PEG = 5.0 kg/mol)macro-reversible addition/fragmentation chain transfer (RAFT)agent,38 and folate-poly(ethylene glycol)-b-poly(D,L-lactide) (FA-PEG-PLA)39 were synthesized according to previous reports. FA-PEG-PLA was prepared from PEG-PLA block copolymer with anMn of6.0−4.5 kg/mol and anMw/Mn of 1.20. UV measurements showed that

FA functionality was close to 100%. 2,2′-Azobisisobutyronitrile (AIBN,98%, J&K) was recrystallized twice from hexane and methanol,respectively. N,N-Dimethylformamide (DMF) dried by MgSO4 wasdistilled under reduced pressure, and 1,4-dioxane was dried by refluxingover sodium wire under an argon atmosphere and distilled prior to use.Doxorubicin hydrochloride (DOX·HCl) and PTX were obtained fromBeijing Zhongshuo Pharmaceutical Technology Development Co. Ltd.Dicyclohexyl carbodiimide (DCC, 98%, Alfa Aesar), N-hydroxysucci-nimide (NHS, 98%, Alfa Aesar), methoxy PEG (Mw 550, Alfa Aesar),ethylene glycol vinyl ether (EGVE, 99%, TCI), 4-dimethylamiopryidine(DMAP, 99%, Alfa Aesar), and p-toluenesulfonic acid monohydrate (p-TSA, 97.5% Acros) were used as received. Molecular sieves (4 Å) werepurchased from Sinopharm Chemical Reagent Co., Ltd. (SCRC).Human nasopharyngeal epidermal carcinoma cell line (KB cell), humancervical cell line (HeLa cell) and human breast adenocarcinoma cell line(MCF-7 cell) were obtained from American Type Culture Collection(ATCC). Paclitaxel resistant human lung carcinoma cell line (A549/PTXR) was obtained from Shanghai Bioleaf Biotech Co. Ltd.

Characterization. 1HNMR spectra were recorded on a Unity Inova400 spectrometer operating at 400 MHz using deuterium oxide (D2O)or deuterated chloroform (CDCl3) as a solvent. The chemical shiftswere calibrated against residual solvent signals. The molecular weightsand polydispersity index (PDI) of the polymers were determined byWaters 1515 gel permeation chromatograph (GPC) instrumentequipped with two linear PLgel columns (500 Å and Mixed-C)following a guard column and a differential refractive-index detector.The measurements were performed using DMF as the eluent at a flowrate of 0.8 mL/min at 30 °C and a series of narrow polystyrene standardsfor the calibration of the columns. The size of nanoparticles wasdetermined using dynamic light scattering (DLS). The measurementswere carried out at 25 °C using Zetasizer Nano-ZS (MalvernInstruments) equipped with a 633 nmHe−Ne laser using backscatteringdetection. Transmission electron microscopy (TEM) was performedusing a Tecnai G220 TEM operated at an accelerating voltage of 200 kV.The samples were prepared by dropping 10 μL of 0.1 mg/mLnanoparticle solution on the copper grid. The amount of PTX wasdetermined by high-performance liquid chromatography (HPLC;Agilent 1260) with UV detection at 227 nm using a mixture ofacetonitrile and water at 1/1 (v/v) as a mobile phase.

Synthesis of PEG-PAA Diblock Polymer. PEG-PAA copolymerwas prepared by RAFT polymerization of AA using PEG-DMP (Mn PEG= 5.0 kg/mol) as a macro-RAFT agent. Under a nitrogen atmosphere,AA (163.0 μL, 2.375 mmol), PEG-DMP (0.5 g, 0.095 mmol), AIBN(3.12 mg, 0.019 mmol), and anhydrous dioxane (6 mL) were added intoa 25 mL Schlenk flask. After blowing the mixture with nitrogen for 30min, the flask was sealed and placed in an oil bath thermostatted at 70°C. The polymerization proceeded under stirring for 48 h. The resultingpolymer was isolated by precipitation in cold diethyl ether, filtration, and

Scheme 1. Illustration of Acetal-Linked PTX Prodrug Micellar Nanoparticles Based on PEG-PAA Block Copolymera

a(i) Self-assembly of acetal-linked PTX prodrugs in water to form nano-sized micellar nanoparticles. (ii) Endosomal pH triggered acetal cleavage andrelease of pristine PTX.

Biomacromolecules Article

dx.doi.org/10.1021/bm400615n | Biomacromolecules 2013, 14, 2772−27802773

drying in vacuo. Yield: 85%. 1H NMR (400 MHz, D2O): PEG (−CH2−CH2−O−: δ 3.63; CH3−O−: δ 3.38), PAA (−CH2−CH−COO−): δ2.37; −CH2−CH−COO−: δ 1.6−1.9). Mn (

1H NMR) = 6.4 kg/mol.Synthesis of Vinyl Ether-Functionalized PEG-PAA. Under a

nitrogen flow, to a 50 mL round-bottom flask equipped with a magneticstirrer were added PEG-PAA (300mg, equv. 1.4 mmol COOH), DMAP(85.3 mg, 0.7 mmol), DCC (863.4 mg, 4.2 mmol), and anhydrousdioxane (10 mL). The flask was sealed, and the reaction was allowed toproceed under stirring for 10 h at 25 °C. Thereafter, ethylene glycol vinylether (EGVE, 625 μL, 6.98 mmol) was charged and allowed to react for24 h at 25 °C in dark. The opaque solution was filtered, and the filtratewas precipitated in cold diethyl ether. The product was collected byfiltration and drying in vacuo in dark. Yield: 71%. 1H NMR (400 MHz,CDCl3): PEG (−CH2−CH2−O−: δ 3.63; CH3−O−: δ 3.38), PAA(−CH2−CH−COO−): δ 2.37; −CH2−CH−COO−: δ 1.6−1.9),EGVE units (−COO−CH2−CH2−O−: δ 4.27; −COO−CH2−CH2−O− 3.87; −O−CHCH2: δ 4.05, 4.18 and 6.50).Mn (

1H NMR) = 7.3kg/mol, Mn (GPC) = 9.0 kg/mol, PDI (GPC) = 1.03.Synthesis of PTX Prodrugs. In a typical example, to a 25 mL

Schlenk flask equipped with a magnetic stirrer were added under anitrogen flow vinyl ether-functionalized PEG-PAA (100.0 mg, 164.8μmol of EGVE units), PTX (47.0 mg, 55.0 μmol), p-TSA (0.32 mg, 1.65μmol), 4 Å molecular sieve (1 g) and anhydrous DMF (10 mL). Theflask was sealed and the reaction was carried out under stirring at 50 °C.After 4 days, the resulting mixture was filtered to remove the molecularsieve and the filtrate was dialyzed against 250 mL DMF (molecularweight cutoff (MWCO) 3500) for 24 h with 5 times change of DMF toremove the unreacted PTX. The final product was obtained by dialysisagainst 500 mLDI water with 4 times change of water and lyophilizationfor 48 h. Yield: 79%. 1HNMR (400MHz, CDCl3): PEG block (−CH2−CH2−O−: δ 3.63; CH3−O−: δ 3.38), PAA (−CH2−CH−COO−): δ2.37; −CH2−CH−COO−: δ 1.6−1.9), EGVE unit (−COO−CH2−CH2−O−: δ 4.27; −COO−CH2−CH2−O− 3.87; −O−CHCH2: δ4.05, 4.18 and 6.50), and PTX (δ 1.1−1.2, δ 2.2−2.4, δ 3.5−3.8, δ 7.3−8.4), acetal bond (−O−CH(CH3)−O−: δ 4.37; −O−CH(CH3)−O−:δ 1.36). Mn (

1H NMR) = 9.3 kg/mol, Mn (GPC) = 17.0 kg/mol, PDI(GPC) =1.42.PTX contents in the prodrug were determined by HPLC and 1H

NMR measurements. PTX content was calculated according to thefollowing formula:

= ×

PTX content (wt %)

(weight of PTX/weight of prodrug) 100%

Preparation of PTX Prodrug Micellar Nanoparticles. PTXprodrug micellar nanoparticles with or without FA-PEG-PLA (20 wt %)were prepared via direct hydration method as described in our previousreport.40 Typically, PTX prodrug (1.0 mg) and PEG 550 (15 μL) werecharged into a 4 mL tube. The mixture was heated to 95 °C, stirred at 95°C for 20 min, and cooled to room temperature. Then, 10, 20, 70, and1900 μL of phosphate buffer (PB; pH 7.4, 10 mM) was added in thatorder with 20 min ultrasonication (200 W, 50 Hz) after each addition.After dialysis (MWCO 3500) against PB (pH 7.4, 10 mM),homogeneous prodrug nanoparticle dispersions were obtained. Thesize and size distribution were determined by DLS measurements.The critical aggregation concentration (CAC) was determined by

using pyrene as a fluorescence probe.41 The concentration of polymerwas varied from 1.0 × 10−5 to 0.1 mg/mL, and the concentration ofpyrene was fixed at 0.6 μM. The fluorescence spectra were recordedusing a FLS920 fluorescence spectrometer with the excitationwavelength of 330 nm. The emission fluorescence at 372 and 383 nmwas monitored. The CAC was estimated as the cross-point whenextrapolating the intensity ratio I372/I383 at low and high concentrationregions.In Vitro PTX Release from PTX Prodrug Nanoparticles. The

release profiles of PTX from prodrug nanoparticles were studied using adialysis tube (MWCO 12 000) under shaking (200 rpm) at 37 °C inthree different media, i.e., pH 5.0 (acetate buffer, 10 mM), pH 6.0(acetate buffer, 10 mM), and pH 7.4 (PB, 10 mM). Typically, 0.5 mL ofprodrug nanoparticle dispersions (0.5 mg/mL) was dialyzed against 25

mL of release media. At desired time intervals, 8 mL of release media wastaken out and replenished with an equal volume of fresh media. Theamount of PTX released was determined by HPLC measurements. Thecumulative PTX release ratio was calculated according to the followingformula:

= × =

=∑ +

× ≥−

⎨⎪⎪

⎩⎪⎪

EV Cm

i

EV C V C

mi

100% ( 1)

100% ( 2)

i

ii i

r0

drug

re 1

11 0

drug

where Er is the cumulative released ratio of PTX (%), Ve and V0 are thevolumes of the exchanged medium and the total medium, respectively(in mL), Ci is the PTX concentration of the release medium taken at theith time (in mg/mL), and mdrug is the total mass of PTX in thenanoparticles (in mg).

The release experiments were conducted in triplicate, and the resultspresented are the average data with standard deviation. The change ofprodrug nanoparticle sizes following release of PTX in time wasmonitored by DLS.

MTT Assays of PTX Prodrug Nanoparticles. The antitumoractivity of PTX prodrug nanoparticles was evaluated in humannasopharyngeal epidermal carcinoma cells (KB), human cervicalcarcinoma cells (HeLa), as well as paclitaxel resistant human lungcarcinoma A549 cells (A549/PTXR) by MTT assays. The cells wereplated in a 96-well plate (5 × 103 cells/well) using RPMI-1640 mediumsupplemented with 10% fetal bovine serum (FBS), 1% L-glutamine, andthe antibiotics penicillin (100 IU mL−1) and streptomycin (100 μgmL−1) for 24 h. The medium was aspirated and replaced by 90 μL offresh medium supplemented with 10% FBS. Ten microliters of prodrugnanoparticles in PB (10 mM, pH 7.4) was added to yield final drugconcentrations of 5 × 10−5∼20 μg PTX equiv/mL. The cells werecultured at 37 °C in an atmosphere containing 5% CO2 for 48 h. Then10 μL of 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazoliumbromide(MTT) solution (5 mg/mL) was added. The cells were incubated foranother 4 h. The medium was aspirated, the MTT-formazan generatedby live cells was dissolved in 150 μL of dimethyl sulfoxide (DMSO), andthe absorbance at a wavelength of 492 nm of each well was measuredusing a microplate reader. PTX formulations prepared by dissolvingPTX in Cremophor EL/ethanol (1:1 v/v) with the concentrationsranging from 5 × 10−5 to 20 μg/mL were used as controls. The relativecell viability (%) was determined by comparing the absorbance at 492nm with control wells containing only cell culture medium. Data arepresented as average ± SD (n = 4).

The cytotoxicity of vinyl ether functionalized PEG-PAA prepolymerwas determined in a similar way at polymer concentrations ranging from0.05 to 1.0 mg/mL. Here, acetaldehyde, which is a hydrolysis byproductof PTX prodrug nanoparticles, was added at a concentration of 32.5 μM(corresponding to a maximal amount of acetaldehyde generated from1.0 mg/mL of PTX prodrug 3 nanoparticles).

Folate-Decorated PTX Prodrug Nanoparticles Targeting toKB Cells. Folate-decorated PTX prodrug micellar nanoparticles wereprepared from PTX prodrug 3 and 20 wt % FA-PEG-PLA. Theantitumor activity of folate-decorated PTX prodrug nanoparticles tofolate receptor-overexpressing KB cells was evaluated by MTT assays infolate free or folate containing RPMI-1640 media. MCF-7 cells (lowfolate receptor-expressing) were used as a negative control. The cellswere plated in a 96-well plate (5 × 103 cells/well) for 24 h. The mediumwas aspirated and replaced by 90 μL of fresh medium. Ten microliters offolate-decorated PTX prodrug nanoparticles, prodrug nanoparticles orfree PTX in PB (10 mM, pH 7.4) was added to yield final drugconcentrations of 5 × 10−5 ∼ 20 μg PTX equiv/mL. The cells werecultured in folic acid free (−FA) or folic acid containing (+FA) media at37 °C in an atmosphere containing 5% CO2 for 3 h. Thereafter, themedium was aspirated and replaced by 100 μL of fresh medium and thecells were cultured for another 45 h. Then, 10 μL of MTT solution (5mg/mL) was added. The cells were incubated for additional 4 h. Themedium was aspirated, the MTT-formazan generated by live cells wasdissolved in 150 μL of DMSO, and the absorbance at a wavelength of

Biomacromolecules Article

dx.doi.org/10.1021/bm400615n | Biomacromolecules 2013, 14, 2772−27802774

492 nm of each well was measured using a microplate reader. Therelative cell viability (%) was determined by comparing the absorbanceat 492 nm with control wells containing only cell culture medium. Dataare presented as average ± SD (n = 4).Loading of DOX and pH-Triggered Co-release of PTX and

DOX.DOX could be physically loaded into PTX prodrug nanoparticles.In a typical example, PTX prodrug (1.0 mg) and PEG 550 (15 μL) werecharged into a 4 mL centrifuge tube. The mixture was heated to 95 °C,stirred at 95 °C for 20 min, and cooled to room temperature. Tenmicroliters of PB (pH 7.4, 10 mM) and 22 μL of DOX solution inDMSO (5.0 mg/mL) were added, and the mixture was stirred at 60 °Cfor 20min. Then, 20, 70, and 1900 μL of PB (pH 7.4, 10mM)was addedin that order with 20 min ultrasonication (200 W, 50 Hz) after eachaddition. After dialysis (MWCO 3500) against PB, homogeneous DOX-loaded prodrug nanoparticle dispersions were obtained. The wholeprocedure was performed in the dark. For determination of drug loadingcontent (DLC), DOX-loaded nanoparticles were lyophilized, dissolvedin DMSO, and analyzed with fluorescence spectroscopy (FLS 920,excitation at 480 nm). The calibration curve was obtained with DOX/DMSO solutions with knownDOX concentrations. DLCwas calculatedaccording to the following formula: DLC (wt %) = (weight of loadedDOX/total weight of prodrug and loaded DOX) × 100%.The co-release profiles of PTX and DOX from DOX-loaded prodrug

nanoparticles were studied using a dialysis method (MWCO 12000)under shaking (200 rpm) at 37 °C in three different media, i.e., pH 5.0(acetate buffer, 10 mM), pH 6.0 (acetate buffer, 10 mM), and pH 7.4(PB, 10 mM). Typically, 0.5 mL of DOX-loaded prodrug nanoparticledispersions (0.5 mg/mL) was dialyzed against 25 mL of release media.At desired time intervals, 8 mL of release media was taken out andreplenished with an equal volume of fresh media. The amounts of PTXand DOX released were determined by HPLC and fluorescencemeasurements, respectively. The release experiments were conducted intriplicate and the results presented are the average data with standarddeviation.

■ RESULTS AND DISCUSSIONSynthesis of Acetal-Linked PTX Prodrugs Based on

PEG-PAA Block Copolymers. PTX was conjugated ontowater-soluble PEG-PAA block copolymers via an acid-labileacetal bond to PAA block using EGVE as a linker (Scheme 2).PEG-PAA diblock copolymers were prepared by controlled

RAFT polymerization of AA using PEG-DMP (Mn = 5.0 kg/mol) as a macro-RAFT agent and AIBN as the radical source indioxane at 70 °C for 48 h. 1H NMR displayed clearly peakscharacteristic of PEG (δ 3.63) and PAA blocks (δ 1.6−1.9 and2.37). The degree of polymerization (DP) of PAA block wasdetermined, by comparing integrals of peaks at δ 2.37 and 3.63, tobe 20, which was close to our designed DP of 25. Using PEG-DMP under similar conditions, we have previously synthesizedthermo-sensitive PEG-PAA-PNIPAM triblock copolymers(PNIPAM = poly(N-isopropylacrylamide)), based on whichreduction-sensitive reversibly cross-linked polymersomes weredeveloped for efficient intracellular protein release.38 EGVE wascoupled to PEG-PAA using DCC and DMAP in dioxane for 48 hto yield vinyl ether-functionalized PEG-PAA. 1H NMR spectrumdisplayed new signals at δ 6.47/4.18/4.05 and δ 4.25/3.87besides peaks of PEG-PAA, attributable to the vinyl andmethylene protons of EGVE moieties, respectively (Figure1A). The number of vinyl ether substitution per polymer chainwas estimated to be 12 by comparing the intensities of signals at δ6.47 (vinyl proton of EGVE moieties) with δ 3.63.PTX was readily conjugated to vinyl ether-functionalized

PEG-PAA copolymers via an acetal bond through a “click”-typeconjugate reaction between 2′-hydroxyl groups of PTX andpendent vinyl ethers. Frey and Wurm et al. reported that vinylether side chains of PEG reacted efficiently with alcohols.42 1HNMR showed besides peaks of vinyl ether-functionalized PEG-PAA also signals due to PTX at δ 1.1−1.2, 2.2−2.4, 3.5−3.8, and7.3−8.4 (Figure 1B). Furthermore, a peak assignable to acetalmethine proton was detected at δ 4.37, supporting conjugation ofPTX to PEG-PAA via an acetal bond. PTX content wasdetermined by HPLC measurements. The results showed thatPTX prodrugs with 21.6, 27.0, and 42.8 wt % PTX (denoted asprodrugs 1, 2, and 3) were obtained at varying PTX-to-polymermolar ratios of 4/1, 6/1, and 12/1, respectively (Table 1). 1HNMR analysis by comparing integrals of signals at δ 7.3−8.4 with3.63 indicated similar PTX contents to those determined withHPLC (Table 1). Notably, these PTX prodrugs have significantly

Scheme 2. Synthesis of Acetal-Linked PTX Prodrugs Based on PEG-PAA Block Copolymersa

aConditions: (i) DCC, DMAP, dioxane, 25 °C, 48 h; (ii) p-TSA, 4 Å molecule sieve, DMF, 50 °C, 96 h.

Biomacromolecules Article

dx.doi.org/10.1021/bm400615n | Biomacromolecules 2013, 14, 2772−27802775

higher drug contents than most prodrugs including PHPMA-peptide-PTX prodrugs (5 wt % PTX, PNU166945) used inphase I clinical trials18 and PHPMA-hydrazide-LEV-PTX (7.6−16.3 wt % PTX).32 GPC revealed that all PTX prodrugs had aunimodal distribution with moderate Mw/Mn of 1.25−1.42(Table 1). Moreover, Mn of PTX prodrugs increased in parallel(from 17.3 to 22.1 kg/mol) with increasing PTX contents.

Formation of PTX Prodrug Micelles and pH-TriggeredDrug Release. PTX prodrugs formed nanosized micellarnanoparticles in PB (pH 7.4, 10 mM) (Figure 2A). The average

hydrodynamic sizes increased from 158.3 to 180.3 nm withincreasing PTX contents from 21.6 to 42.8 wt % (Table 2). TEM

micrograph revealed that these prodrug nanoparticles had aspherical morphology and particle sizes close to thosedetermined by DLS (Figure 2B). Notably, these PTX prodrugshad low CACs of 3.16−3.47 mg/L (Table 2), and were stableduring storage for several months at 4 °C.The in vitro drug release from PTX prodrug nanoparticles was

studied at 37 °C under three different conditions, i.e., pH 5.0, 6.0,and 7.4. The results showed that PTX release was significantlyaccelerated under mildly acidic environments (Figure 3). Forexample, ca. 86.9%, 66.4% and 29.0% of PTX was released fromprodrug 3 nanoparticles in 48 h at pH 5.0, 6.0, and pH 7.4,respectively. Moreover, all three prodrug nanoparticles exhibited

Figure 1. 1H NMR spectra (400 MHz, CDCl3) of vinyl ether-functionalized PEG-PAA block copolymer (A) and PTX prodrug 1 (B).

Table 1. Synthesis of PTXProdrugs Based on PEG-PAABlockCopolymers

Mn (kg/mol) PTX (wt %)

PTXprodrug

PTX/polymermolar feed ratio

1HNMRa GPCb

Mw/Mnb

(PDI) HPLCc1H

NMRd

1 4 9.3 17.3 1.42 21.6 22.32 6 10.0 19.5 1.25 27.0 27.63 12 13.2 22.1 1.30 42.8 43.4

aDetermined by 1H NMR. bDetermined by GPC measurements(eluent: DMF, flow rate: 0.8 mL/min, polystyrene standards).cDetermined by HPLC measurements. dCalculated by 1H NMR viacomparing integrals of signals at δ 7.3−8.4 with 3.63.

Figure 2. Size distribution profiles of PTX prodrug nanoparticlesdetermined by DLS (A) and of PTX prodrug 1 nanoparticlesdetermined by TEM (B).

Table 2. Formation of PTX Prodrug Nanoparticles

PTX prodrug nanoparticle size (nm)a PDIa CACb (mg/L)

prodrug 1 158.3 ± 11.5 0.17 ± 0.01 3.16prodrug 2 167.1 ± 16.7 0.18 ± 0.03 3.27prodrug 3 180.3 ± 11.4 0.14 ± 0.05 3.47

aDetermined by DLS at a concentration of 0.5 mg/mL in PB at 25 °C.bCACs determined by fluorescence microscopy using pyrene as aprobe.

Figure 3. pH-dependent drug release from acetal-linked PTX prodrugnanoparticles at 37 °C. The initial prodrug nanoparticle concentrationwas 0.5 mg/mL.

Biomacromolecules Article

dx.doi.org/10.1021/bm400615n | Biomacromolecules 2013, 14, 2772−27802776

similar PTX release profiles under otherwise the sameconditions, in which slightly faster drug release was observedfor PTX prodrug nanoparticles with higher PTX contents. Itshould be noted that the drug release rates from these acetal-linked PTX prodrug nanoparticles were much faster than thoseof ester or peptide-linked PTX prodrug micelles.31,43−45 Asexpected, no burst release was observed in all cases owing tocovalent linking of PTX to PEG-PAA. The release of PTX fromPTX prodrug nanoparticles at pH 5.0 resulted in marked increaseof particle sizes from 160 to 1600 nm and decline of scatteringintensities in 48 h (Figure 4). 1H NMR studies showed that

signals assignable to PTX moieties as well as acetal methineproton (δ 4.37) were diminishing following acidic treatment of

PTX prodrug nanoparticles. In contrast, little size change wasobserved for PTX prodrug nanoparticles in PB (10 mM)containing 10% serum or in PBS (10 mM, 150 mMNaCl) at pH7.4 and 37 °C over a period of 4 days, in accordance with slowrelease of PTX at physiological pH. These results demonstratethat acetal-linked PTX prodrug nanoparticles, while stable atphysiological pH, effectively release PTX under a mildly acidicenvironment.

In Vitro Anti-Tumor Activity of PTX Prodrug MicellarNanoparticles to Drug-Sensitive and Resistant CancerCells. The antitumor activity of PTX prodrug nanoparticles wasinvestigated using MTT assays. Figure 5 showed that PTXprodrug nanoparticles induced pronounced antitumor effects toboth KB and HeLa cancer cells. Notably, PTX prodrugnanoparticles displayed similar drug efficacy to free PTX inHeLa cells at a concentration of 0.1 μg PTX equiv/mL and lower(Figure 5A). At drug concentrations higher than 0.1 μg PTXequiv/mL, antitumor activity of PTX prodrug nanoparticlesbecame somewhat lower than free PTX. The maximal halfinhibitory concentrations (IC50) were determined to be 0.9, 1.6,and 0.3 μg of PTX equiv/mL for prodrug 1 nanoparticles,prodrug 3 nanoparticles, and free PTX, respectively. KB cellswere muchmore sensitive to PTX prodrug nanoparticles and freePTX than HeLa cells, in which low IC50 values of 0.47, 0.18, and0.03 μg PTX equiv/mL were obtained for prodrug 1 nano-particles, prodrug 3 nanoparticles, and free PTX, respectively(Figure 5B). The higher antitumor activity of prodrug 3nanoparticles as compared to prodrug 1 nanoparticles mightbe due to their slightly faster drug release at endosomal pHcondition (Figure 3). It is remarkable that these acetal-linkedPTX prodrug nanoparticles retain high antitumor efficacy asprodrugs, and prodrug nanoparticles usually exhibit one to twomagnitude lower antitumor efficacy as compared to the parent

Figure 4. Change of average size and scattering intensity of PTXprodrug 1 nanoparticles in time at pH 5.0 and 37 °Cmonitored by DLSat a concentration of 0.5 mg/mL. The kcps is the mean count rate in kilocounts per second.

Figure 5.MTT assays of PTX prodrug nanoparticles (A−C) and vinyl ether-functionalized PEG-PAA copolymer (D). (A) HeLa cells (B) KB cells, and(C) PTX-resistant A549/PTXR cells. The cells were incubated with PTX prodrug nanoparticles for 48 h at drug concentrations varying from 5× 10−5 to40 μg PTX equiv/mL. Free PTX was used as a control. (D) The cytotoxic effect of vinyl ether-functionalized PEG-PAA copolymer containing 32.5 μMacetaldehyde (maximal possible concentration of acetal degradation product resulting from 1.0 mg/mL PTX prodrug 3 nanoparticles) to KB, HeLa, andMCF-7 cells. The cells were incubated with polymers for 48 h. Data are presented as the average ± standard deviation (n = 4).

Biomacromolecules Article

dx.doi.org/10.1021/bm400615n | Biomacromolecules 2013, 14, 2772−27802777

free drugs.6,25,46 For example, Yang et al. reported that poly(L-γ-glutamyl-glutamine)-docetaxel prodrug micelles exhibited ap-proximately 22-fold higher IC50 than free docetaxel against NCI-H460 cells.47

In the following, we studied the antitumor activity of acetal-linked PTX prodrug nanoparticles to PTX-resistant A549 cells(A549/PTXR). As expected, little cell death was observed forA549/PTXR cells following 48 h incubation with free PTX, evenat a high PTX dosage of 20 μg/mL (Figure 5C). The IC50 of freePTX toward A549/PTXR cells was determined to be 175.8 μg/mL, which was over 800 times higher than that observed fordrug-sensitive A549 cells (IC50 = 0.2 μg/mL), further confirmingthat A549/PTXR cells are PTX-resistant (Figure 5C). Bycontrast, pH-sensitive PTX prodrug 3 nanoparticles causedsignificant death of A549/PTXR cells under otherwise identicalconditions, in which cell viability was reduced to ca. 50.3% at adosage of 10 μg PTX equiv/mL (Figure 5C). The IC50 wasdetermined to be 10.9 μg PTX equiv/mL. It is remarkable thatthese acetal-linked PTX prodrug nanoparticles can effectivelyovercomemultidrug resistance, which is probably due to effectivecellular uptake via the endocytosis pathway as well as fast PTXrelease in the mildly acidic endo/lysosomal compartments. Incomparison, vinyl ether-functionalized PEG-PAA prepolymerwas practically nontoxic to KB, HeLa, and MCF-7 cells (85−100% cell viabilities) up to a tested concentration of 1.0 mg/mL(Figure 5D). Notably, 32.5 μM acetaldehyde, which is themaximal possible concentration of acetal degradation productresulting from 1.0 mg/mL PTX prodrug 3 nanoparticles, wasadded during cell culture. These results indicate that vinyl ether-functionalized PEG-PAA has good biocompatibility and issuitable for drug delivery.Targetability and Antitumor Activity of Folate-Deco-

rated PTX Prodrug Micelles. The present pH-sensitive PTXprodrug micelles can be used as a versatile platform for efficientchemotherapy. For instance, tumor-targeting PTX prodrugnanoparticles could be developed to further enhance drugspecificity and efficacy. To verify our concept, folate (FA)-decorated PTX prodrug nanoparticles were prepared bycombining 80 wt % PTX prodrug 3 with 20 wt % FA-PEG-PLA. The PEG in FA-PEG-PLA was designed longer than that inPEG-PAA (6.0 versus 5.0 kg/mol), to facilitate exposure of FAligands on nanoparticle surfaces for optimal targeting.39 Theresulting nanoparticles had average sizes of ca. 180 nmwith a PDIof 0.18. Here, folate receptor-overexpressing KB cells and MCF-7 cells (negative control) were incubated for 3 h with prodrugnanoparticles or free PTX at a dosage of 5 μg PTX/mL, mediawas removed and replenished with fresh culture media, and cellswere further cultured for another 45 h. Notably, MTT assaysdemonstrated that antitumor activity of FA-decorated prodrug 3nanoparticles was significantly more potent against KB cells thannontargeting PTX prodrug 3 nanoparticles (34.6% versus 61.8%cell viability) in folate free medium (Figure 6A). In fact, FA-decorated PTX prodrug 3 nanoparticles had antitumor efficacyclose to free PTX in KB cells. The inhibition experiments showedthat the presence of folic acids in culture media (+FA) resulted inlargely reduced antitumor efficacy of FA-decorated PTX prodrug3 nanoparticles to a comparable level to nontargeting prodrug 3nanoparticles, indicating that FA-decorated PTX prodrug 3nanoparticles delivered PTX into KB cells via a receptor-mediated mechanism. Moreover, both FA-decorated andnontargeting PTX prodrug 3 nanoparticles exhibited similarlylow inhibition of MCF-7 cells (ca. 71.1% cell viability). The plotof cell viabilities against drug concentrations showed that FA-

decorated PTX prodrug 3 nanoparticles had an IC50 of 1.42 μgPTX equiv/mL (1.69 μM), which was about 12 times lower thanthat of nontargeting prodrug 3 nanoparticles (17.16 μg PTXequiv/mL, 20.5 μM) (Figure 6B). These results confirm that FA-decorated PTX prodrug 3 nanoparticles had apparent target-ability and high antitumor activity to folate-overexpressingcancer cells.

pH-Triggered Co-release of PTX and Physically LoadedDOX. These acetal-linked PTX prodrug nanoparticles can alsobe used as a pH-sensitive nanocarrier for controlled release ofother anticancer drugs, providing an elegant approach forcombination cancer therapy. In this study, DOX was physicallyloaded into PTX prodrug 3 nanoparticles with loading contentsof 2.3, 3.8, and 7.2 wt % at theoretical drug loading contents of 5,10, and 20 wt %, respectively. The in vitro release results showedthat both PTX and DOX were released in a pH dependentmanner (Figure 7). For instance, 85.3, 63.4, and 24.7% of PTXand 82.8, 57.1 and 32.0% of DOXwere released in 48 h at pH 5.0,6.0 and pH 7.4, respectively. We previously reported pH-triggered release of both PTX and DOX·HCl from pH-sensitivedegradable polymersomes.36 Discher et al. reported that co-delivery of PTX and DOX by polymersomes resulted insynergistic treatment effect.48 These acetal-linked PTX prodrugnanoparticles present a versatile platform for pH-triggered co-release of PTX and various potent drugs.

Figure 6. Antitumor activity of FA-decorated PTX prodrug 3nanoparticles. Nontargeting PTX prodrug 3 nanoparticles and freePTX were used as controls. The cells were incubated with prodrugnanoparticles or free PTX for 3 h, media was removed and replenishedwith fresh culture media, and cells were further cultured for another 45 h.(A) KB cells and MCF-7 cells at a PTX dosage of 5 μg/mL in folic acidfree (−FA) or folic acid containing (+FA) media; and (B) KB cells atdrug concentrations varying from 5 × 10−5 to 20 μg PTX equiv/mL.Data are presented as the average ± standard deviation (n = 4)(Student’s t test, **p < 0.01, ***p < 0.001).

Biomacromolecules Article

dx.doi.org/10.1021/bm400615n | Biomacromolecules 2013, 14, 2772−27802778

■ CONCLUSIONS

We have demonstrated that acetal-linked PTX prodrug micellarnanoparticles based on PEG-PAA block copolymer possessexcellent in vitro antitumor activity to drug-sensitive as well asdrug-resistant cancer cells. Notably, these PTX prodrugnanoparticles have several unique features: (i) They containremarkable drug contents (up to 42.8 wt % PTX), which aresignificantly higher than macromolecular prodrugs as well asmicelles. (ii) They, though robust under physiological pHenvironments, show accelerated drug release under endosomalpH conditions. Importantly, because PTX is linked to PEG-PAAvia an acetal bond, released PTX is in a pristine state and retainspotent therapeutic activity. (iii) They exhibit potent antitumoractivity to multidrug-resistant (MDR) cells likely due to effectivecellular uptake via the endocytosis pathway as well as fastintracellular release of PTX. (iv) Tumor-targeting PTX prodrugmicellar nanoparticles can be formulated to achieve specific andefficient cancer chemotherapy. (v) They can also be used as apH-sensitive nanocarrier for controlled release of otheranticancer drugs to achieve synergistic treatment effect. (vi)They can readily be prepared with controlled structures andmolecular weights from PEG-PAA block copolymer. Theseacetal-linked PTX prodrug micellar nanoparticles have appearedas a highly versatile and potent platform for targeted cancerchemotherapy.

■ AUTHOR INFORMATION

Corresponding Author*Tel/Fax: +86-512-65880098; E-mail: [email protected](F.M.); [email protected] (Z.Z.).

NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTS

This work is financially supported by research grants from theNational Natural Science Foundation of China (NSFC51003070, 51103093, 51173126, 51273137 and 51273139),the National Science Fund for Distinguished Young Scholars(NSFC 51225302), and a Project Funded by the PriorityAcademic Program Development (PAPD) of Jiangsu HigherEducation Institutions.

■ REFERENCES(1) Langer, R. Nature 1998, 392, 5−10.(2) Nicolas, J.; Mura, S.; Brambilla, D.; Mackiewicz, N.; Couvreur, P.Chem. Soc. Rev. 2013, 42, 1147−1235.(3) Haag, R.; Kratz, F. Angew. Chem., Int. Ed. 2006, 45, 1198−1215.

(4) Deng, C.; Jiang, Y. J.; Cheng, R.; Meng, F. H.; Zhong, Z. Y. NanoToday 2012, 7, 467−480.(5) Gelderblom, H.; Verweij, J.; Nooter, K.; Sparreboom, A. Eur. J.Cancer 2001, 37, 1590−1598.(6) Duncan, R. Nat. Rev. Cancer 2006, 6, 688−701.(7) Torchilin, V. P. Nat. Rev. Drug Discovery 2005, 4, 145−160.(8) Kataoka, K.; Harada, A.; Nagasaki, Y. Adv. Drug Delivery Rev. 2001,47, 113−131.(9) Fleige, E.; Quadir, M. A.; Haag, R. Adv. Drug Delivery Rev. 2012, 64,866−884.(10) Meng, F. H.; Zhong, Z. Y.; Feijen, J. Biomacromolecules 2009, 10,197−209.(11) Davis, M. E.; Chen, Z.; Shin, D. M.Nat. Rev. Drug Discovery 2008,7, 771−782.(12) Cheng, R.; Meng, F. H.; Deng, C.; Klok, H. -A.; Zhong, Z. Y.Biomaterials 2013, 34, 3647−3657.(13) Cabral, H.; Matsumoto, Y.; Mizuno, K.; Chen, Q.; Murakami, M.;Kimura, M.; Terada, Y.; Kano, M. R.; Miyazono, K.; Uesaka, M.;Nishiyama, N.; Kataoka, K. Nat. Nanotechnol. 2011, 6, 815−823.(14) Lammers, T.; Subr, V.; Ulbrich, K.; Peschke, P.; Huber, P. E.;Hennink, W. E.; Storm, G. Biomaterials 2009, 30, 3466−3475.(15) Rautio, J.; Kumpulainen, H.; Heimbach, T.; Oliyai, R.; Oh, D.;Jarvinen, T.; Savolainen, J. Nat. Rev. Drug Discovery 2008, 7, 255−270.(16) Leonelli, F.; La Bella, A.; Migneco, L.; Bettolo, R.Molecules 2008,13, 360−378.(17) Li, C.; Yu, D. F.; Newman, R. A.; Cabral, F.; Stephens, L. C.;Hunter, N.; Milas, L.; Wallace, S. Cancer Res. 1998, 58, 2404−2409.(18) Meerum Terwogt, J. M.; ten Bokkel Huinink, W. W.; Schellens, J.H. M.; Schot, M.; Mandjes, I. A. M.; Zurlo, M. G.; Rocchetti, M.; Rosing,H.; Koopman, F. J.; Beijnen, J. H. Anti-Cancer Drugs 2001, 12, 315−323.(19) O’Brien, M. E. R.; Socinski, M. A.; Popovich, A. Y.; Bondarenko, I.N.; Tomova, A.; Bilynsky, B. T.; Hotko, Y. S.; Ganul, V. L.; Kostinsky, I.Y.; Eisenfeld, A. J.; Sandalic, L.; Oldham, F. B.; Bandstra, B.; Sandler, A.B.; Singer, J. W. J. Thorac. Oncol. 2008, 3, 728−734.(20) Paz-Ares, L.; Ross, H.; O’Brien, M.; Riviere, A.; Gatzemeier, U.;Von Pawel, J.; Kaukel, E.; Freitag, L.; Digel, W.; Bischoff, H.; Garcia-Campelo, R.; Iannotti, N.; Reiterer, P.; Bover, I.; Prendiville, J.;Eisenfeld, A. J.; Oldham, F. B.; Bandstra, B.; Singer, J. W.; Bonomi, P. Br.J. Cancer 2008, 98, 1608−1613.(21) Matsumura, Y.; Kataoka, K. Cancer Sci. 2009, 100, 572−579.(22) Oerlemans, C.; Bult, W.; Bos, M.; Storm, G.; Nijsen, J. F.;Hennink, W. Pharm. Res. 2010, 27, 2569−2589.(23) Kim, T. Y.; Kim, D. W.; Chung, J. Y.; Shin, S. G.; Kim, S. C.; Heo,D. S.; Kim, N. K.; Bang, Y. J. Clin. Cancer Res. 2004, 10, 3708−3716.(24) http://c l in ica l t r ia l s feeds .org/cl in ica l - t r ia l s/show/NCT01644890.(25) Hu, X. L.; Jing, X. B. Expert Opin. Drug Delivery 2009, 6, 1079−1090.(26) Chen, W. L.; Shi, Y. L.; Feng, H.; Du, M.; Zhang, J. Z.; Hu, J. H.;Yang, D. J. Phys. Chem. B 2012, 116, 9231−9237.(27) Alani, A. W.; Bae, Y.; Rao, D. A.; Kwon, G. S. Biomaterials 2010,31, 1765−1772.

Figure 7. pH-dependent release of PTX (A) and DOX (B) from DOX-loaded PTX prodrug 3 nanoparticles at 37 °C. DOX-loaded PTX prodrug 3nanoparticle concentration was 0.5 mg/mL and DOX loading content was 3.8 wt %.

Biomacromolecules Article

dx.doi.org/10.1021/bm400615n | Biomacromolecules 2013, 14, 2772−27802779

(28) Xiao, H. H.; Song, H. Q.; Yang, Q.; Cai, H. D.; Qi, R. G.; Yan, L. S.;Liu, S.; Zheng, Y. H.; Huang, Y. B.; Liu, T. J.; Jing, X. B. Biomaterials2012, 33, 6507−6519.(29) Zhang, X. F.; Li, Y. X.; Chen, X. S.; Wang, X. H.; Xu, X. Y.; Liang,Q. Z.; Hu, J. L.; Jing, X. B. Biomaterials 2005, 26, 2121−2128.(30) Talelli, M.; Iman, M.; Varkouhi, A. K.; Rijcken, C. J. F.; Schiffelers,R. M.; Etrych, T.; Ulbrich, K.; van Nostrum, C. F.; Lammers, T.; Storm,G.; Hennink, W. E. Biomaterials 2010, 31, 7797−7804.(31) Xie, Z. G.; Guan, H. L.; Chen, X. S.; Lu, C. H.; Chen, L.; Hu, X. L.;Shi, Q.; Jing, X. B. J. Controlled Release 2007, 117, 210−216.(32) Etrych, T.; Sírova, M.; Starovoytova, L.; Ríhova, B.; Ulbrich, K.Mol. Pharmaceutics 2010, 7, 1015−1026.(33)Wei, H.; Zhuo, R. X.; Zhang, X. Z. Prog. Polym. Sci. 2012, 38, 503−535.(34) Gillies, E. R.; Jonsson, T. B.; Frechet, J. M. J. J. Am. Chem. Soc.2004, 126, 11936−11943.(35) Gillies, E. R.; Frechet, J. M. J. Bioconjugate Chem 2005, 16, 361−368.(36) Chen, W.; Meng, F. H.; Cheng, R.; Zhong, Z. Y. J. ControlledRelease 2010, 142, 40−46.(37) Lai, J. T.; Filla, D.; Shea, R.Macromolecules 2002, 35, 6754−6756.(38) Xu, H. F.; Meng, F. H.; Zhong, Z. Y. J. Mater. Chem. 2009, 19,4183−4190.(39) Xiong, J.; Meng, F. H.; Wang, C.; Cheng, R.; Liu, Z.; Zhong, Z. Y.J. Mater. Chem. 2011, 21, 5786−5794.(40)Wang,W.; Sun, H. L.; Meng, F. H.; Ma, S. B.; Liu, H. Y.; Zhong, Z.Y. Soft Matter 2012, 8, 3949−3956.(41) Yang, R.; Meng, F. H.; Ma, S. B.; Huang, F. S.; Liu, H. Y.; Zhong,Z. Y. Biomacromolecules 2011, 12, 3047−3055.(42) Mangold, C.; Dingels, C.; Obermeier, B.; Frey, H.; Wurm, F.Macromolecules 2011, 44, 6326−6334.(43) Khandare, J. J.; Jayant, S.; Singh, A.; Chandna, P.;Wang, Y.; Vorsa,N.; Minko, T. Bioconjugate Chem. 2006, 17, 1464−1472.(44) Liang, L.; Lin, S. W.; Dai, W. B.; Lu, J. K.; Yang, T. Y.; Xiang, Y.;Zhang, Y.; Li, R. T.; Zhang, Q. J. Controlled Release 2012, 160, 618−629.(45)Miller, K.; Erez, R.; Segal, E.; Shabat, D.; Satchi Fainaro, R. Angew.Chem., Int. Ed. 2009, 48, 2949−2954.(46) Fox, M. E.; Guillaudeu, S.; Frechet, J. M. J.; Jerger, K.; Macaraeg,N.; Szoka, F. C. Mol. Pharmaceutics 2009, 6, 1562−1572.(47) Yang, D. B.; Van, S.; Shu, Y. Y.; Liu, X. Q.; Ge, Y. F.; Jiang, X. G.;Jin, Y.; Yu, L. Int. J. Nanomed. 2012, 7, 581−589.(48) Ahmed, F.; Pakunlu, R. I.; Brannan, A.; Bates, F.; Minko, T.;Discher, D. E. J. Controlled Release 2006, 116, 150−158.

Biomacromolecules Article

dx.doi.org/10.1021/bm400615n | Biomacromolecules 2013, 14, 2772−27802780