Real-Time Drug Release Analysis of Enzyme and pH Responsive

Aug 3, 2015 - Graduate student gets prison sentence for poisoning. A former chemistry PhD candidate at Queen's University in Canada who confessed t...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/JPCB

Real-Time Drug Release Analysis of Enzyme and pH Responsive Polysaccharide Nanovesicles Poothayil Subash Pramod,§ Nilesh Umakant Deshpande, and Manickam Jayakannan* Department of Chemistry, Indian Institute of Science Education and Research (IISER)-Pune, Dr. Homi Bhabha Road, Pune 411008, Maharashtra, India

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

S Supporting Information *

ABSTRACT: The accurate estimation of drug release kinetics of polymeric vehicles is an indispensable prerequisite for the developments of successful drug carriers for cancer therapy. The present investigation reports the development of timeresolved fluorescence spectroscopic approach for the real-time release kinetics of fluorophore loaded polysaccharide vesicles that are potential vectors in cancer treatment. The polysaccharide vesicles were custom designed with appropriate enzyme and pH responsiveness and loaded with water-soluble biocompatible fluorophore Rhodamine B (Rh-B). The semipermeable membrane dialysis method along with steady state absorbance spectroscopic technique was found to be inaccurate for the estimation of drug release. Time correlated single photon counting (TCSPC) technique was found to exhibit significant difference in excited state decay profiles and fluorescent lifetime of Rh-B in the free and polymer bound states. This enabled the establishment of real-time drug release protocols by TCSPC method for polysaccharide vesicles that are responsible to pH and enzyme with respect to intracellular compartments. Real-time analysis predicted the release kinetics 20−25% higher accuracy when compared to the dialysis method under in vitro conditions. Moreover, the ability of enzyme to cleave the polysaccharide vesicles was further validated by docking studies. The positioning of the molecules in active site of enzyme and the binding energy data were generated using AUTODOCK program to study the rupture of polysaccharide vesicles. This new TCSPC technique could be very useful for studying the drug release pattern of synthetic polymer vesicles loaded with Rh-B fluorophore.



dialysis for the in vitro drug release kinetics.16 Recent studies revealed that this approach had limitations due to the process by which the estimation of drug release analysis was performed.17 Typically, the dialysis approach involved two steps: (i) the rupture of the polymer membrane and transportation of the loaded cargoes (or drugs) to the liquid medium inside the dialysis tube and (ii) transportation of the drug molecules from the dialysis tube to the outer reservoir by crossing the barrier of the dialysis membrane. These processes are rather different from the actual forces experienced by the polymer vesicles when they are administered in the cell in vitro or intravenous injection in vivo. As a consequence, most often the drug release kinetics profiles obtained from the dialysis method were found to be mismatched with the cellular cleavage of polymer nanoscaffolds.17 Reliable methods to calculate the in vitro drug release kinetics are necessary for establishing a good correlation between drug release capabilities and clinical effectiveness of formulation of synthetic polymer vesicles. Real-time analysis of drug release kinetics by in situ methods are very good protocols

INTRODUCTION Tailor made polymeric structures have received tremendous attention in the recent past for controlled drug delivery in cancer treatment.1,2 Self-organized polymeric structures such as micelles,3 vesicles,4 and nanoparticles5,6 were found to undergo passive selectivity in cancer tissues through enhanced permeability and retention (EPR) effect.7,8 Among all these assemblies, polymer vesicular assemblies are unique nanoscaffolds since they can be utilized for loading and delivering the combination of both hydrophilic and hydrophobic drugs.9 Unlike the liposmes, the synthetic polymer vesicular drug carriers were found to be more stable under physiological conditions and in systemic circulation that are essential for the in vivo administration of drugs.10 Cancer tissue or cancer related pathological inflammation is found to be overexpressed by enzyme, such as glutathione reductase11 and esterase.12,13 Enzyme responsive vesicles (or polymersomes) were developed to exploit the cancer tissue environment for selective and targeted delivery to cancer cells.14,15 Nevertheless, there has been no effort taken until now to study the real-time drug release kinetics of enzyme responsive polymer vesicles to unlock the mechanism of enzyme action on the polymer backbone. Absorption spectroscopy analysis is routinely employed in combination with semipermeable membrane © 2015 American Chemical Society

Received: June 17, 2015 Revised: July 25, 2015 Published: August 3, 2015 10511

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

The Journal of Physical Chemistry B

Figure 1. (a) Chemical structure of enzyme responsive dextran vesicles and their cellular entry. (b) Drug/dye loaded vesicles and their cleavage in the presence of esterase enzyme. (c) TCSPC method for real-time release profiles for free and polymer bound drug/dye molecules.

(DOX) and water insoluble drug camptothecin (CPT) in their core and layer, respectively. CPT loaded vesicles showed excellent drug pharmacophore retaining capability and also two times better killing of cancer cells compared to the free drug.25 DOX loaded polymer vesicles showed 90% cell killing in breast cancer MCF-7 cells compared to the DOX alone (only 30% effect). When both drugs were loaded together in the same vesicle, that is, dual drug loaded, vesicles showed synergistic effect in MCF-7 and colon cancer cells (DLD-1) as a result of combination therapy.27 The cellular uptake studies for fluoroprobe dye Rhodamine B (Rh-B) loaded polymer vesicles revealed that both DOX and Rh-B loaded vesicles showed almost identical cell penetrating ability, perinuclear accumulation, as well as responsiveness to biological stimuli like enzymes.27 Enzyme and pH dual responsive dextran vesicles for delivery of DOX into breast cancer cells were also developed.28 Hence, developing a new real-time drug analysis method for these responsive and cell penetrating polysaccharide vesicles would facilitate the development of diagnostic fluoroprobes for the detection of cancer diseases. The present investigation is emphasized to develop real-time drug release methodologies for polysaccharide vesicles that are responsive to biological enzymes and/or acidic pH using timeresolved fluorescence spectroscopy. The pitfalls of dialysis technique for exact drug release patterns were successfully overcome by time-resolved fluorescent (TCSPC) method (see Figure 1b). DOX and Rh-B loaded vesicles were subjects for the above purpose, and it was found that the Rh-B loaded vesicular system is an excellent candidate for real-time drug

to overcome some of the above limitations and correlate the drug release kinetics pattern by the spectroscopic tools to their actual action in the cell line in vitro conditions. Fluoroprobe tagged polymer structures were employed as model systems to study the real-time drug release patterns using triggers such as pH and temperature.18,19 Viger et al. demonstrated the application of time-resolved fluorescence technique for probing drug release kinetics of poly(L-glutamic acid) nanoparticles.20 Unfortunately, these studies did not provide information on the cyctotoxicity, cellular uptake mechanism, or cellular uptake capability of their polymer design or the probe altogether. Fluorescence lifetime imaging microscopy (FLIM) is a new technique developed to image the internal organelles of the cellular compartments.21 The fluorophores employed in the FLIM technique exhibit difference in their excited state fluorescence lifetime with respect to their environment such as hydrophilic or hydrophobic which in turn is used for imaging a particular site.22 Organic dyes such as nitrobenzoxadiazole (NBD), laurdan, 8-alkoxy quinoline, and Rhodamine B were employed for the FLIM analysis.23,24 Thus, the real-time drug release kinetics of synthetic polymer vesicles responsive to enzyme and pH are important to be studied for fundamental understanding as well as long-term application based on fluorescence lifetime imaging. Recently, we have reported cell penetrable polysaccharide nanovesicles based on modified dextran derivatives anchored with renewable resource hydrophobic units25−27 (see Figure 1a). These vesicles were found to be excellent candidates for loading water-soluble anticancer drugs such as doxorubicin·HCl 10512

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

The Journal of Physical Chemistry B kinetic analysis by TCSPC. Further, the dextran-PDP (DEXPDP), imine coupled pH and enzyme responsive DEX-IM-5 molecules were also subjected to docking studies with esterase enzyme active site to trace the enzyme-mediated drug release process. The overall investigation revealed that the real-time fluorescence method is a very powerful methodology for understanding the drug release kinetics of cell penetrable polymer vesicles that are potential vectors for cancer treatment.

DLC (%) = {weight of drug encapsulated in vesicles/weight of drug loaded polymer} × 100% = 3.15 wt % for DEX-PDP-5, 2.9 wt % for DEX-IM-5 vesicles. DLE (%) = {Weight of drug encapsulated in vesicles/weight of drug in feed} × 100% = 63% for DEX-PDP-5, 58% for DEXIM-5 vesicles. In Vitro Release Studies. The drug and dye release from the vesicles was performed at pH 7.4 (physiological pH) for DEX-PDP and acidic pH (pH 5.0 and pH6.0) for DEX-IM vesicles using dialysis method as per the procedure reported earlier.25−28 A control experiment was also performed using solution without polymer by just adding DOX·HCl (60 μg) or Rh-B in 3 mL of PBS (pH 7.4). In brief, polysaccharide vesicles encapsulated with Rh-B or DOX·HCl in 3.0 mL of buffer solution (pH 7.4) were placed in a dialysis tube (SPECTRA/ POR, MWCO 8000). Control free drug/dye solution was used for elution experiment without any previous dialysis. The tube carrying drug/dye loaded vesicles was submerged in 100 mL buffer solutions (required pH) taken in a beaker, and the whole solution was incubated at 37 °C. After the appropriate time interval (30 or 60 min), 3.0 mL of dialysate was recovered and restored with an equal volume of fresh buffer. Absorbance of the collected aliquot was measured at 480 nm using UV−vis absorption spectrophotometry, and the amount of the drug/ dye eluted was calculated using Beer’s law, where molar absorption coefficient of Rh-B and DOX was fixed as 115 000 and 11 500, respectively. The effect of esterase enzyme on drug/dye release was studied by mixing 10 U esterase enzyme into the vesicle solution prior to the release studies. Photophysical Studies. The absorption studies were done using PerkinElmer Lambda 45 UV−visible spectrophotometer. Emission spectra were recorded at room temperature using a SPEX Fluorolog HORIBA JOBIN VYON fluorescence spectrophotometer equipped with a photomultiplier tube (PMT, Hamamatsu R928) for detection, a double-grating 0.22 m Spex1680 monochromator, and a 450W Xe lamp as the excitation source. The excitation spectra were collected at 550 nm (for Rh-B experiment) and 480 nm (for DOX experiment), and the emission spectra were recorded by exciting at the excitation maxima. The dye containing samples were purged with N2 gas for at least 15−20 min prior to photophysical experiments. Fluorescence lifetime of both free dye/drug and dye/drug loaded vesicles was measured using a time-correlated single photon counting (TCSPC) system (Horiba) equipped with a NanoLED excitation source of 560 nm for Rh-B and 450 nm for DOX·HCl with 1 MHz impulse repetition rate (Horiba) and a R928P detector (Hamamatsu Photonics, Japan). In Rh-B and DOX experiments, the detector was set to 585 and 560 nm, respectively. The bandpass width was varied between 2 and 14.7 nm to attain suitable α value. A scattering solution of Ludox-40 (Sigma-Aldrich) in water was used as prompt (560 or 450 nm) to obtain the instrument response function (IRF). A total of 4094 channels were used with a time calibration of 0.110 channel/ns. The fluorescence lifetimes were estimated using DAS6 decay analysis software (Horiba), where the goodness of fit was evaluated by fixing χ2 values between 0.95 and 1.20. Drug/dye loaded into vesicles was fitted to a monoexponential as follows:29

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795



EXPERIMENTAL SECTION Materials. Doxorubicin·HCl (DOX), Rhodamine B (Rh-B), and horse liver esterase enzyme were purchased from Aldrich Chemicals. Modified dextran derivatives were synthesized as reported in earlier work and were used for the present studies.25,28 Encapsulation of Rhodamine-B. Water-soluble Rhodamine B was loaded into vesicles inner core by the method described elsewhere.25 Briefly, 20 mg of polysaccharide derivatives (DEX-PDP-5 or DEX-IM-5) and 2 mg of Rhodamine B were dispersed in 3 mL of DMSO solvent. This solution was agitated in a beaker at 25 °C, and 3 mL of phosphate buffered saline (PBS, pH 7.4) was added slowly (dropwise) into it. The resulting dye−polymer mixture was stirred again for another 12 h at 25 °C and then transferred into regenerated cellulose dialysis bag (SPECTRA/POR, MWCO 3,500). The dialysis process was carried on until Rhodamine B leaching out from the dialysis bag was entirely stopped. The vesicles loaded with Rhodamine B dye were recovered from the dialysis bag and diluted to 10 mL with PBS. The Rhodamine B loaded in the vesicles was completely released by dissolving 100 μL of the above stock solution in 2.9 mL of DMSO solvent. The Rhodamine B loading amount was determined by measuring the absorbance of the above diluted solution at 552 nm and substituting it in the Beer’s law equation where the molar absorption coefficient of Rhodamine B was fixed as 111 500. The loading content of Rh-B in vesicles was determined as 2.70 wt %. Encapsulation of Doxorubicin·HCl. Doxorubicin·HCl (DOX·HCl) was encapsulated in vesicles by the method described earlier.25−27 Briefly, 100 mg of polymer (DEX-PDP-5 or DEX-IM-5) and 2 mg of DOX·HCl were codissolved in 3 mL of DMSO and stirred at 25 °C in a light protected condition. Self-assembly of the polymer was induced by slow addition of 3 mL of PBS (pH 7.4) into the above solution. The resulting polymer drug solution was stirred at 25 °C under dark condition for 12 h, transferred to a dialysis bag (SPECTRA/ POR, MWCO 3500), and then extensively dialyzed in a light protected container against PBS (0.01 M, pH 7.4) for 24 h to remove DMSO and unencapsulated DOX molecules. The resulting drug loaded vesicle solution was freeze-dried after passing through 0.45 μM filters. The drug loading content (DLC) in vesicles was estimated using absorption spectroscopy method. Briefly, 3 mg of lyophilized drug loaded polymer derivative was dissolved in DMSO (1 mL) and then 100 μL of this solution was made to 3.0 mL with DMSO. The absorbance of this resulting solution at 480 nm was measured on a UV− visible spectrophotometer, and the amount of drug loaded in the vesicles was determined using Beer’s law where the molar extinction coefficient of DOX was fixed as 11 500. Drug loading content (DLC) and drug loading efficiency (DLE) were estimated using following equations.

A(t ) = α1 exp( −t /τ1) 10513

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

The Journal of Physical Chemistry B

Figure 2. (a) Schematic representation of DOX and Rh-B loading into vesicles. Absorbance spectra of free DOX and DOX loaded in vesicles (b) and free Rh-B and Rh-B loaded vesicles (c). Emission spectra of free DOX and DOX loaded in vesicles (d) and free Rh-B and Rh-B in vesicles (e). Fluorescence lifetime plot of free DOX and DOX loaded in vesicles (f) and free Rh-B and Rh-B in vesicles (g). Insets in (d) and (e) show DOX and Rh-B loaded vesicle solutions in vials under handheld UV light exposure.

Esterase enzyme treated samples were fitted to a sum of three exponentials. Lifetime components with insignificant fractional contributions (fi < 3%) were rejected, and decays were judged to be biexponential, where the sum of the fractional contributions from free Rh-B and Rh-B loaded in vesicles (f1 + f 2) was more than 95%.

enzyme, docking of monomer and trimer units of dextran-PDP and DEX-IM-5 vesicles to human liver carboxyl esterase 1 was performed. The crystal structure of human liver carboxyl esterase 1 (PDB ID: 3K9B)30 was taken as the receptor and catalytic triad composed of serine, histidine, and glutamine in the interface between the subunits of the enzyme was taken as the active site. The quantum chemically optimized geometry of monomer and trimer units of dextran-PDP and DEX-IM vesicles were taken as the ligands. The optimization was done using Hartree−Fock theory with basis set 6-31G* using Gaussian 0331 software. The restrained electrostatic potential (RESP)32 charges on the atoms of the optimized structures were calculated using Antechamber module of AMBERTools.33 For the enzyme, the charges on the atoms were calculated using AMBER99SB force field34 available in AMBER1135 software. During docking, the quantum chemically derived charges on the atoms of ligands and the AMBER99SB charges on the receptor were kept unchanged. The docking was performed using AutoDock software.36,37 The receptor was kept rigid, and the ligands were kept flexible during docking. The grid was generated on active site of human liver carboxyl esterase 1 with grid points 126, 74, and 80 in X, Y, and Z directions,

A(t ) = α1 exp( −t /τ1) + α2 exp(−t /τ2)

The average lifetime (τAVE) was computed as per the equation τAVE = Σf iτi, where f i is the fractional contribution of the individual lifetime of each component (τi).11,29 f1 = [α1τ1/(α1τ1 + α2τ2)] × 100 f2 = [α2τ2/(α1τ1 + α2τ2)] × 100

The experiment of DEX-PDP vesicles encapsulated with Rh-B was performed at pH 7.4 and by adding 10 U esterase enzyme, while that of DEX-IM vesicles loaded with Rh-B were performed under four conditions: (i) pH 6.0, (ii) pH 5.0, (iii) pH 6.0 and 10 U esterase, and (iv) pH 7.4 and 10 U esterase. Docking Protocol. To understand the interaction of dextran-PDP (DEX-PDP) and DEX-IM-5 vesicles to esterase 10514

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

The Journal of Physical Chemistry B

Figure 3. (a) Endocytosis of vesicles and drug release at the intracellular compartments. (b) TCSPC decay profiles of Rh-B in dextran vesicle along with prompt. (c) TCSPC decay profiles of Rh-B in dextran vesicle upon adding enzyme and collecting the data at various time intervals.

respectively, with a grid spacing of 0.192 Å. Genetic algorithm33 was used as the search method for the docking process.

e). This indicated that the steady state absorbance and emission spectroscopic methods were not able to distinguish between the free drug in solution and drug encapsulated in vesicles. Time-resolved fluorescent decay profiles of DOX and Rh-B in their free state and loaded form were recorded by TCSPC techniques. TCSPC decay profiles of the free DOX and DOX in vesicles are given in Figure 2f. The decay profiles of the DOX followed single exponential decay with decay rate constant (τ1) of 1.1 ns in its free form38 and 2.0 ns for loaded in the vesicles. The decay rate constants revealed that the DOX molecules have almost identical lifetime for both free and loaded in the vesicles with a small difference of 0.9 ns. This difference in the DOX lifetime is not very useful for the differentiation of free and polymer bound drug for real-time analysis. Remarkably, significant difference was observed in the TCSPC lifetime of free Rh-B and Rh-B loaded in the vesicles. TCSPC decay profiles of the Rh-B (see Figure 2g) followed single exponential decay with τ1 = 1.74 ns for free39 and 4.90 ns for loaded in the vesicles. Further, the decay profiles also showed substantial difference for free Rh-B and Rh-B in the vesicles. Thus, this large difference in the time-resolved fluorescent decay process could be used as a tool to understand the real-time release kinetics of the drug or dye from polysaccharide vesicles. The quantum yields of the free DOX and free Rh-B were compared in loaded and free form in the vesicular scaffolds. It was found that both the free drug and loaded drug (DOX or Rh-B) displayed identical quantum yield; thus, the larger difference among their τ1 values did not arise due to any variation in their quantum yield. The time-resolved fluorescence offers important information about the Rh-B environment in the aqueous medium. Photoexcited Rh-B molecules return to ground state by transferring the excitation energy into neighboring Rh-B or solvent molecules. When the Rh-B molecules were present in a



RESULTS AND DISCUSSION DOX and Rh-B Loaded DEX-PDP Vesicles. Amphiphilicity was introduced into microorganism derived dextran polysaccahride by coupling with plant derived hydrophobic pentadecyl phenol (PDP) on its backbone via aliphatic ester linkage as described earlier (see Figure 1a).25 The modified dextran derivaties (DEX-PDP) self-assembled in water to produce nanovesicular morphologies. These vesicles were charcatrized by electron microscopy, atomic force microscopy, and dynamic and static light scattering, and the details are published elswehere.25,27 The present investigation is mainly devoted to study the drug releasing capabilities of these vesicles for water-soluble fluorescent anticancer drug doxorubicin·HCl (DOX) and water-soluble fluorophore dye Rhodamine B (RhB). Under identical conditions, DOX and Rh-B were loaded in the dextran vesicles by dissolving the guest molecules and the polymer in dimethyl sulfoxide + water and dialyzed against a large amount of water (6 times water exchange) for 72 h until the dialysate became devoid of released cargoes. At the end of the dialysis, the DOX and Rh-B encapsulated dextran loaded samples were observed as stable red/pink color solution in the dialysis tube. These vesicular structures were characterized by electron microscopy, atomic force microscopy, and dynamic and static light scattering methods, and these details were reported elsewhere.25,27 A schematic representation of DOX and Rh-B loading into vesicles (see Figure 2a) and their photophysical characteristics are shown in Figure 2b−e. DOX and Rh-B loaded vesicles showed identical absorbance and emission maxima for free as well as encapsulated drug in vesicular scaffold (see Figure 2b− 10515

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

The Journal of Physical Chemistry B

Figure 4. (a) Fractional contribution of Rh-B released in vesicles versus time in TCSPC method. (b) In vitro release profile of Rh-B loaded vesicles in dialysis method. (c) Steady state fluorescence spectra of Rh-B released from dextran vesicles. (d) Schematic representation of Rh-B/DOX released from dextran vesicles under normal and enzyme conditions. (e) Dynamic dialysis of Rh-B loaded vesicle solution and schematic representation of dye transport from intravesicular to extravesicular environment and subsequent flow across dialysis membrane.

Enzyme-Responsive Vesicles. The endocytosed vesicles undergo rupturing in the presence of esterase enzyme abundant in lysosomes as shown in Figure 3a. Typically, endocytosis of nanocarrier starts with endosome formation by membrane invagination followed by delivery of cargoes into various vesicular structures and end up in lysosomes.40 Degradative enzymes in lysosomes rupture the carriers and release the loaded cargoes. Detailed cellular uptake studies were done earlier to quantify the fluorescent intensity of the Rh-B loaded vesicles in normal cells (wild type mouse embryonic fibroblasts, WT-MEFs), breast cancer cells (MCF-7), and colon cancer cells (DLD-1).27 The Rh-B vesicular particles were found to be stable and they were not found to affect both normal and cancer cells even though differences exist at the microlevel in

confined space inside the vesicular scaffold, their collision with a large water pool is restricted. This enabled the excited Rh-B molecules to undergo slow decay process which led to the increase in the fluorescence lifetime of the molecule. Unfortunately, only Rh-B fluorophore exhibited sensitivity in TCSPC analysis with respect to scaffold confined and the DOX (anticancer drug-cum-fluorophore) did not show diffrence for free and in the loaded form. Since both DOX and Rh-B showed similar cellular uptake and cell perinuclear accumulation in breast and colon cancer cells; the variation in the TCSPC analysis in Rh-B molecules in the free and loaded form (inside the vesicles) can be employed to study the vesicular membrance in real-time process. 10516

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

The Journal of Physical Chemistry B

normal conditions in PBS (pH 7.4) at 37 °C in the absence of enzyme reached 95%. The fitting of decay profiles aided to estimate the fractional contribution of released free Rh-B and Rh-B still located inside vesicular cavity at time t (see Table ST2). The component having a lifetime close to 4.90 ns was judged to be the Rh-B molecules located inside the vesicular hydrophilic cavity (polymer bound). The lifetime close to 1.70 ns was that of free Rh-B released from vesicles. The fractional contribution of vesicle encapsulated in Rh-B was obtained as 100% at time t = 0 min. Interestingly, the fractional contribution of Rh-B released when vesicles incubated under 10517

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

The Journal of Physical Chemistry B

Figure 5. (a) Chemical structure of DEX-IM-5. TCSPC decay profiles of Rh-B in DEX-IM-5 vesicles at pH = 6.0 (b), at pH = 5.0 (c), at pH 7.4 in the presence of esterase (d), and at pH 6.0 in the presence of esterase enzyme (e).

Figure 5a for chemical structure of DEX-IM-5, where 5 denotes degree of substitution on dextran backbone). Similar to DEXPDP vesicles, the steady state absorbance and emission methods could not identify any distinctness between free RhB and Rh-B loaded in DEX-IM-5 vesicles. Rh-B in DEX-IM-5 vesicles exhibited 3 nm red shift compared to free Rh-B in PBS (see Figure SF-6a). On the other hand, the emission spectra did not display any appreciable difference between Rh-B in loaded and free form (see Figure SF-6b). Analogous to DEX-PDP vesicles, DEX-IM-5 vesicles also showed a significant difference in the TCSPC lifetime of free Rh-B in PBS or water and Rh-B in vesicles. The lifetime of Rh-B in DEX-IM-5 vesicles was 4.82 ns (see Figure SF-6c). Thus, the time-resolved fluorescence decay profile could be used to quantify the release from pH and enzyme responsive DEX-IM-5 vesicles. Drug release studies under acidic pH conditions and in the presence of esterase enzyme were performed by adopting a similar procedure as illustrated in the case of DEX-PDP vesicles. The absorbance of Rh-B loaded DEX-IM-5 vesicular solutions was adjusted to 0.2

not considered in drug release rate determination. A few other groups also reported similar issues with respect to inherent barrier properties of dialysis membrane and drug partitioning between dispersed phases during drug transport processes.16,17 Hence, compared to dynamic dialysis method, the real-time monitoring using TCSPC technique displayed about 20 ± 5% difference under normal conditions (without enzyme) and 20 ± 1% under enzymatic conditions. The actual drug concentration present in the interior hydrophilic cavities of vesicles and the rate of drug release across the vesicular wall are the main factors to be considered. In dialysis method, the properties of dialysis membrane and drug partitioning between vesicle and dialysis membrane may also affect the final release rate. The physical disturbance that occurs due to additional mechanical force exerted during dialysis process hastens the transport of drug/dye across the wall. pH and Enzyme Dual Responsive Vesicles. Rh-B was also loaded into pH and enzyme dual responsive polysaccharide DEX-IM-5 vesicles and the details were given elsewhere.28 (See 10518

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

The Journal of Physical Chemistry B

Figure 6. (a) Cumulative drug release profiles of Rh-B at various pH. (b) Fractional contribution of Rh-B released at various pH in TCSPC method. (c) Cumulative drug release profiles of Rh-B at various pH in the presence of esterase enzyme. (d) Fractional contribution of Rh-B released in at various pH in the presence of esterase enzyme by TCSPC method.

O.D. using phosphate buffer (PB, pH 6.0 or pH 5.0) and used for TCSPC analysis directly or by adding 10 U esterase enzyme as explained in experimental sections. The decay profile collected at pH 6.0 (see Figure 5b), at pH 5.0 (see Figure 5c), at pH 7.4 with esterase (see Figure 5d), and at pH 6.0 with esterase (see Figure 5e) was fitted with DAS6 decay analysis software. As explained earlier, acidic pH can hydrolyze imine linkage connecting PDP and dextran and subsequent leakage of loaded Rh-B content. Therefore, the drug release analysis of Rh-B loaded DEX-IM vesicles were performed using both dynamic dialysis (see Figure 6a) and time-resolved fluorescence method (see Figure 6b) under acidic pH conditions relevant to cancer pH (pH 6.0) and intracellular lysosomal pH (pH 5.0). The Rh-B molecules eluted out from the dialysis bag to reservoir was estimated using absorbance method, and the result showed that about 80% release under pH 6.0 and close to 100% release under lysosomal (pH 5.0) conditions (see Figure 6a). In the TCSPC method, the fractional contribution of Rh-B released ( f1) and polymer bound Rh-B molecule ( f 2) was obtained by fitting method as described for DEX-PDP vesicles (see Figure 6b) (see the Supporting Information for Tables ST3 to ST-6). At pH 6.0, the component having lifetime close to 4.82 ns was decreased with time and reached to 62% (see Figure 6b). It means that the Rh-B molecules located inside the vesicular hydrophilic cavity (polymer bound) eluted out, which declined the fractional contributions of vesicles loaded Rh-B. Remarkably, this value was almost 20% more than that predicted by dynamic dialysis method. The lifetime of Rh-B loaded in the vesicles also reduced to 3.2 ns from 4.82 ns (see Figure SF-7). A similar trend was observed in the case of pH 5.0 (see Figures 6b and SF-8). At low pH, more rapid release of Rh-B was observed from DEX-IM-5 vesicles and the release

reached about 90% in 48 h. The lifetime of Rh-B loaded in the vesicles drastically reduced to 3.0 ns from 4.82 ns (see Figure SF-6). On the other hand, the lifetime of released Rh-B was retained in the range of 1.68−1.78 ns as observed in the case of DEX-PDP vesicles (see Figure SF-8). In Figure 6, the comparison of the dialysis data and TCSPC data revealed that the real-time method provided 20% more accuracy for drug release kinetics at all pH and also in the presence of esterase enzymes. Thus, the time-resolved luminescent decay profiles can be used as tool to make out free Rh-B and Rh-B loaded in vesicles during the release process. This process was verified for two sets of dextran vesicles having esterase and pH plus esterase responsiveness. The plots of fluorescence lifetime values change for Rh-B released and bound to DEX-IM-5 vesicles at pH = 7.4 and pH = 6.0 in the presence of esterase are given in Figures SF-9 and SF-10. As expected, the lifetime of Rh-B eluted into buffer solutions remained in the range of 1.68−1.76 ns (see the Supporting Information for tables). This suggested that the time-resolved luminescent decay profiles can be used as a tool to differentiate free Rh-B and Rh-B loaded in vesicles during the release process. The present studies on the vesicular membrane cleavage by real-time TCSPC analysis provided a new opportunity for studying the release kinetics of a wide range of nanostructures in drug delivery. The vesicular cleavage process was investigated with Rh-B, since it is strong fluorescence dye and preferably suitable for the TCSPC analysis. In general, this approach can be expanded to other polymeric vesicles and also these Rh-B loaded vesicles could be explored as probes. Currently, efforts have been made to check the process for various fluorophore molecules having emission from 400 to 900 nm, and this future analysis may offer more insight into the cellular uptake and real-time TCSPC of polymer vesicles. 10519

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

The Journal of Physical Chemistry B

Figure 7. (a) Ribbon representation of human liver carboxyl esterase 1 (PDB ID: 3K9B. (b) Schematic representation of catalytic triad binding to vesicles and vesicle cleavage. (c) Mechanism of action of Ser-His-Asp catalytic triad on DEX-PDP molecule to get dextran and PDP acid molecules. (d) Optimized structures of monomer of DEX-PDP, trimer of DEX-PDP, monomer of DEX-IM, and trimer of DEX-IM.

Figure 8. Docked configurations of (a) monomer of DEX-PDP, (b) trimer of DEX-PDP, (c) monomer of DEX-IM-5, and (d) trimer of DEX-IM-5.

10520

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

The Journal of Physical Chemistry B Docking Studies of Enzyme and pH Responsive Vesicles. The in vitro drug release studies using dynamic dialysis and TCSPC method showed that the release of Rh-B from DEX-PDP and DEX-IM-5 vesicles was marginally increased in the presence of esterase enzyme. So to understand this rapid dissociation of Rh-B from vesicles in detail, docking of monomer and trimer units DEX-PDP and DEX-IM-5 vesicles to the catalytic site of human liver carboxyl esterase 1 was performed. The studies on esterase enzyme proved that its active site constitutes a catalytic triad composed of three amino acids, serine, histidine, and glutamine, which is responsible for hydrolysis of esters (see Figure 7a and b).42 As a first step, carbonyl carbon of the ester portion of the substrate was attacked by the nucleophilic serine hydroxyl group. This nuleophilic attack breaks the π-bond and forms a tetrahedral intermediate, which is stabilized by the oxanion hole of the enzyme. Later on, cleavage of the intermediate happens to form alcohol and acid which was removed from the active site by diffusion42 (see Figure 7c for mechanism of cleavage illustrated using DEX-PDP as an example). The aim of the docking study was to understand the affinity and preference of monomer and trimer units of DEX-PDP and DEX-IM-5 vesicles toward the amino acids in the catalytic triad of the esterase enzyme. The optimization of monomer and trimer units of DEX-PDP and DEX-IM-5 molecules were performed using Gaussian 03 software (Hartree−Fock theory using 6-31G* basis set),31 and the restrained electrostatic potential (RESP)32 charges on the atoms of the optimized structures were estimated using Antechamber module of AMBERTools33 (see Figure 7d for optimized structures of DEX-PDP and DEX-IM-5). As the initial step in the hydrolysis process was the nucleophilic attack by the hydroxyl group of serine, the extent of cleavage of DEX-PDP vesicles could be correlated to the position and orientation of the carbonyl group of the vesicles with respect to the hydroxyl group of serine (see Figure 7c for mechanism of DEX-PDP ester bond cleavage associated with catalytic triad. A similar mechanism can be predicted for DEX-IM-5). The docked configurations of monomer units of DEX-PDP vesicles showed that they are positioned well inside the catalytic triad (as shown in Figure 8a) with appreciable binding energies (see Table ST-7). The catalytic cleavage of the ester group in monomer unit of DEX-PDP may be faster because oxygen of the carbonyl group faces downward, exposing the carbonyl carbon more toward the serine hydroxyl group (as shown in Figure 8a) for the easy nucleophilic attack. To understand the effect of polymer chain, the docking was extended from monomer to trimer units of DEX-PDP vesicles. The docking results are summarized in Table ST-7, and the corresponding docked configurations are shown in Figure 8b. The interaction between serine hydroxyl group and ligand carbonyl group is shown and marked by the yellow oval shapes. Here also, the binding energy of DEX-PDP trimer was more favorable and the positive binding energy was due to the presence of large number of rotatable bonds in comparison with the corresponding monomers. The analysis of the docked configurations demonstrated that the DEX-PDP trimer was more prone to nucleophilic attack by serine hydroxyl group because of the close vicinity and favorable orientation of the carbonyl group. The docked conformation of DEX-IM monomer and trimer are shown in Figure 8c and d, respectively (see Table ST-7 for binding energy data). In this case, also molecules were positioned inside the catalytic triad of the enzyme. Thus, the

docking studies explained the experimental observation of rapid hydrolysis of DEX-PDP and DEX-IM-5 vesicles by esterase enzyme in terms of favorable binding energy, position, and orientation of carbonyl group with respect to serine hydroxyl group for the initial nucleophilic attack during ester hydrolysis. To summarize, a novel technique based on fluorescent decay profile and lifetime of Rhodamine B fluorescent dye molecule was described to determine the enzymatic release kinetics from polysaccharide vesicles. The limitations of conventional steady state absorbance techniques plus dialysis method in distinguishing species such as Rh-B−PBS (released) and Rh-B−carrier (polymer bound) were effectively overcome using real-time analysis method. The rapid degradation kinetics of vesicles in the presence of enzyme stimuli was precisely calculated explained using TCSPC method and docking studies.



CONCLUSION In conclusion, a new method based on fluorescent lifetime of the Rh-B fluorophore was developed to study the real-time drug release kinetics from enzyme and pH responsive polysaccharide vesicles. Amphiphilic dextran and starch polysaccharide derivatives were encapsulated with Rh-B molecules, and these luminescent vesicles provided a new opportunity to study the real-time drug release profiles in a cell penetrable polymer carrier in cancer therapy. The pitfalls of the widely adopted dynamic dialysis plus absorbance method in drug release kinetics were successfully overcome using timeresolved fluorescent method. The real-time drug release kinetics based on fluorescence lifetime, and their fractional contribution was found to be more accurate by 20 ± 10% compared to the dialysis method in the three vesicles studied. The basis for enhanced hydrolysis rate in the presence of enzyme was investigated using docking studies. The specific positioning of the ester linked amphiphilic unit in the active site of the enzyme and the binding energy data supported the ability of the enzyme to attack on vesicles. The present investigation provided a new insight into the accurate release rate determination from enzyme and pH responsive polysaccharide vesicles that are very good drug delivery vehicles for breast and colon cancers. Though the approach here demonstrated only Rh-B loaded polysaccharide vesicles, in principle, this technique could be adapted for a large number of other polymers. Further, the polysaccharide vesicles loaded with Rh-B are a very good biological probe for fluorescent life imaging microscopes for quantifying or studying the cellular uptake mechanistic pathways and live cell imaging. Currently, research work is focused on these directions to utilize the new polysaccharide probe for studying cancer behaviors for better cancer treatment.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcb.5b05795. Tables containing fluorescent lifetime data, release profiles, and cumulative drug release patterns of vesicles in FBS; absorbance and fluorescence data for ph and enzyme responsive vesicles (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Fax: +91-20-2590 8186. 10521

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

The Journal of Physical Chemistry B Present Address

(16) Moreno-Bautista, G.; Tam, K. C. Evaluation of Dialysis Membrane Process for Quantifying the in Vitro Drug-Release from Colloidal Drug Carriers. Colloids Surf., A 2011, 389, 299−303. (17) Modi, S.; Anderson, B. D. Determination of Drug Release Kinetics from Nanoparticles: Overcoming Pitfalls of the Dynamic Dialysis Method. Mol. Pharmaceutics 2013, 10, 3076−3089. (18) Weinstain, R.; Segal, E.; Satchi-Fainaro, R.; Shabat, D. RealTime Monitoring of Drug Release. Chem. Commun. 2010, 46, 553− 555. (19) Cui, W.; Lu, X.; Cui, K.; Wu, J.; Wei, Y.; Lu, Q. Fluorescent Nanoparticles of Chitosan Complex for Real-Time Monitoring Drug Release. Langmuir 2011, 27, 8384−8390. (20) Viger, M. L.; Sheng, W.; McFearin, C. L.; Berezin, M. Y.; Almutairi, A. Application of Time-Resolved Fluorescence for Direct and Continuous Probing of Release from Polymeric Delivery Vehicles. J. Controlled Release 2013, 171, 308−314. (21) Berezin, M. Y.; Achilefu, S. Fluorescence Lifetime Measurements and Biological Imaging. Chem. Rev. 2010, 110, 2641−2684. (22) Aigner, D.; Dmitriev, R. I.; Borisov, S. M.; Papkovsky, D. B.; Klimant, I. pH-Sensitive Perylenebisimide Probes for Live Cell fluorescence Lifetime Imaging. J. Mater. Chem. B 2014, 2, 6792−6801. (23) Becker, W. Fluorescence Life Time Imaging- Techniques and Applications. J. Microsc. 2012, 247, 119−136. (24) Bunge, A.; Fischlechner, M.; Loew, M.; Arbuzova, A.; Herrmann, A.; Huster, D. Characterization of Lipid Bilayers Adsorbed on Spherical LbL-Support. Soft Matter 2009, 5, 3331−3339. (25) Pramod, P. S.; Takamura, K.; Chaphekar, S.; Balasubramanian, N.; Jayakannan, M. Dextran Vesicular Carriers for Dual Encapsulation of Hydrophilic and Hydrophobic Molecules and Delivery into Cells. Biomacromolecules 2012, 13, 3627−3640. (26) Sridhar, U.; Pramod, P. S.; Jayakannan, M. Creation of Dextrin Vesicles and Their Loading-Delivering Capabilities. RSC Adv. 2013, 3, 21237−21241. (27) Pramod, P. S.; Shah, R.; Chaphekar, S.; Balasubramanian, N.; Jayakannan, M. Polysaccharide Nano-Vesicular Multidrug Carriers for Synergistic Killing of Cancer Cells. Nanoscale 2014, 6, 11841−11855. (28) Pramod, P. S.; Shah, R.; Jayakannan, M. Dual Stimuli Polysaccharide Nanovesicles for Conjugated and Physically Loaded Doxorubicin Delivery in Breast Cancer Cells. Nanoscale 2015, 7, 6636−6652. (29) Lakowicz, J. R. Principles of Fluorescence Spectroscopy; Springer: New York, 2010; Vol. 3, pp 102−106. (30) Hemmert, A. C.; Otto, T. C.; Wierdl, M.; Edwards, C. C.; Fleming, C. D.; MacDonald, M.; Cashman, J. R.; Potter, P. M.; Cerasoli, D. M.; Redinbo, M. R. Human Carboxylesterase 1 Stereoselectively Binds the Nerve Agent Cyclosarin and Spontaneously Hydrolyzes the Nerve Agent Sarin. Mol. Pharmacol. 2010, 77, 508− 516. (31) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03, Revision C.02, Gaussian Inc.: Wallingford, CT, 2003. (32) Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Kollmann, P. A. Application of Resp Charges to Calculate Conformational Energies,

§

(P.S. Pramod) Pharmacy Division, PIPMS, Government Medical College, Thiruvananthapuram, Kerala, India.

Notes

The authors declare no competing financial interest.

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795



ACKNOWLEDGMENTS The authors are thankful for a research grant from Department of Science and Technology (DST), New Delhi, INDIA, under nanomission initiative project SR/NM/NS-42/2009 and SB/ S1/OC-37/2013 under SERB scheme. The authors thank Dr. Nagaraj Balasubramanian, Department of Biology, IISER-Pune for providing cell culture facilities and cellular uptake studies. N.U.D. thanks UGC New Delhi for JRF research fellowship. The authors thank Mrs. D. S. Wilbee, Department of Chemistry, IISER-Pune for docking studies.



REFERENCES

(1) Larson, N.; Ghandehari, H. Polymeric Conjugates for Drug Delivery. Chem. Mater. 2012, 24, 840−853. (2) Nicolas, J.; Mura, S.; Brambilla, D.; Mackiewicz, N.; Couvreur, P. Design, Functionalization Strategies and Biomedical Applications of Targeted Biodegradable/Biocompatible Polymer-Based Nanocarriers for Drug Delivery. Chem. Soc. Rev. 2013, 42, 1147−1235. (3) Ahmad, Z.; Shah, A.; Siddiq, M.; Kraatz, H.-B. Polymeric Micelles as Drug Delivery Vehicles. RSC Adv. 2014, 4, 17028−17038. (4) Brinkhuis, R. P.; Rutjes, F. P. J. T.; van Hest, J. C. M. Polymeric Vesicles in Biomedical Applications. Polym. Chem. 2011, 2, 1449− 1462. (5) Maity, A. R.; Chakraborty, A.; Mondal, A.; Jana, N. R. Carbohydrate Coated, Folate Functionalized Colloidal Graphene as a Nanocarrier for Both Hydrophobic and Hydrophilic Drugs. Nanoscale 2014, 6, 2752−2758. (6) Torchilin, V. P. Multifunctional, Stimuli-Sensitive Nanoparticulate Systems for Drug Delivery. Nat. Rev. Drug Discovery 2014, 13, 813−827. (7) Maeda, H. Tumor-Selective Delivery of Macromolecular Drugs Via the EPR Effect: Background and Future Prospects. Bioconjugate Chem. 2010, 21, 797−802. (8) Fang, J.; Nakamura, H.; Maeda, H. The Epr Effect: Unique Features of Tumor Blood Vessels for Drug Delivery, Factors Involved, and Limitations and Augmentation of the Effect. Adv. Drug Delivery Rev. 2011, 63, 136−151. (9) Du, J.; O'Reilly, R. K. Advances and Challenges in Smart and Functional Polymer Vesicles. Soft Matter 2009, 5, 3544−3561. (10) Meng, F.; Zhong, Z.; Feijen, J. Stimuli-Responsive Polymersomes for Programmed Drug Delivery. Biomacromolecules 2009, 10, 197−209. (11) Saito, G.; Swanson, J. A.; Lee, K.-D. Drug Delivery Strategy Utilizing Conjugation Via Reversible Disulfide Linkages: Role and Site of Cellular Reducing Activities. Adv. Drug Delivery Rev. 2003, 55, 199− 215. (12) Guo, D.-S.; Wang, K.; Wang, Y.-X.; Liu, Y. CholinesteraseResponsive Supramolecular Vesicle. J. Am. Chem. Soc. 2012, 134, 10244−10250. (13) Feng, A.; Yuan, J. Smart Nanocontainers: Progress on Novel Stimuli-Responsive Polymer Vesicles. Macromol. Rapid Commun. 2014, 35, 767−779. (14) Rodriguez, A. R.; Kramer, J. R.; Deming, T. J. Enzyme-Triggered Cargo Release from Methionine Sulfoxide Containing Copolypeptide Vesicles. Biomacromolecules 2013, 14, 3610−3614. (15) Habraken, G. J. M.; Peeters, M.; Thornton, P. D.; Koning, C. E.; Heise, A. Selective Enzymatic Degradation of Self-Assembled Particles from Amphiphilic Block Copolymers Obtained by the Combination of N-Carboxyanhydride and Nitroxide-Mediated Polymerization. Biomacromolecules 2011, 12, 3761−3769. 10522

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523

Article

Downloaded by NANYANG TECHNOLOGICAL UNIV on August 24, 2015 | http://pubs.acs.org Publication Date (Web): August 12, 2015 | doi: 10.1021/acs.jpcb.5b05795

The Journal of Physical Chemistry B Hydrogen Bond Energies, and Free Energies of Solvation. J. Am. Chem. Soc. 1993, 115, 9620−9631. (33) Case, D. A.; Cheatham, T. E., 3rd; Darden, T.; Gohlke, H.; Luo, R.; Merz, K. M., Jr.; Onufriev, A.; Simmerling, C.; Wang, B.; Woods, R. J. The Amber Biomolecular Simulation Programs. J. Comput. Chem. 2005, 26, 1668−88. (34) Hornak, V.; Abel, R.; Okur, A.; Strockbine, B.; Roitberg, A.; Simmerling, C. Comparison of Multiple Amber Force Fields and Development of Improved Protein Backbone Parameters. Proteins: Struct., Funct., Genet. 2006, 65, 712−725. (35) Case, D. A.; Darden, T. A.; T.E. Cheatham, I.; Simmerling, C. L.; Wang, J.; Duke, R. E.; Luo, R.; Walker, R. C.; Zhang, W.; Merz, K. M.; Roberts, B. P.; Wang, B.; Hayik, S.; Roitberg, A.; Seabra, G.; Kolossváry, I.; Wong, K. F.; Paesani, F.; Vanicek, J.; Wu, X.; Brozell, S. R.; Steinbrecher, T.; Gohlke, H.; Cai, Q.; Ye, X.; Wang, J.; Hsieh, M.J.; Cui, G.; Roe, D. R.; Mathews, D. H.; Seetin, M. G.; Sagui, C.; Babin, V.; Luchko, T.; Gusarov, S.; Kovalenko, A.; Kollman, P. A. Amber 11, University of California: San Francisco, 2010. (36) Morris, G. M.; Goodsell, D. S.; Halliday, R. S.; Huey, R.; Hart, W. E.; Belew, R. K.; Olson, A. J. Automated Docking Using a Lamarckian Genetic Algorithm and An Empirical Binding Free Energy Function. J. Comput. Chem. 1998, 19, 1639−1662. (37) Morris, G. M.; Huey, R.; Lindstrom, W.; Sanner, M. F.; Belew, R. K.; Goodsell, D. S.; Olson, A. J. Autodock4 and Autodocktools4: Automated Docking with Selective Receptor Flexibility. J. Comput. Chem. 2009, 30, 2785−2791. (38) Zhang, X.; Shastry, S.; Bradforth, S. E.; Nadeau, J. L. Nuclear Uptake of Ultrasmall Gold-Doxorubicin Conjugates Imaged by Fluorescence Lifetime Imaging Microscopy (Flim) and Electron Microscopy. Nanoscale 2015, 7, 240−251. (39) Sharma, V.; Shivalingaiah, S.; Peng, Y.; Euhus, D.; Gryczynski, Z.; Liu, H. Auto-Fluorescence Lifetime and Light Reflectance Spectroscopy for Breast Cancer Diagnosis: Potential Tools for Intraoperative Margin Detection. Biomed. Opt. Express 2012, 3, 1825−1840. (40) Duncan, R.; Richardson, S. C. W. Endocytosis and Intracellular Trafficking as Gateways for Nanomedicine Delivery: Opportunities and Challenges. Mol. Pharmaceutics 2012, 9, 2380−2402. (41) Xu, G.; Zhang, W.; Ma, M. K. Human Carboxylesterase 2 Is Commonly Expressed in Tumor Tissue and Is Correlated with Activation of Irinotecan. Clin. Cancer Res. 2002, 8, 2605−2611. (42) Richter, F.; Blomberg, R.; Khare, S. D.; Kiss, G.; Kuzin, A. P.; Smith, A. J. T.; Gallaher, J.; Pianowski, Z.; Helgeson, R. C.; Grjasnow, A.; Xiao, R.; Seetharaman, J.; Su, M.; Vorobiev, S.; Lew, S.; Forouhar, F.; Kornhaber, G. J.; Hunt, J. F.; Montelione, G. T.; Tong, L.; Houk, K. N.; Hilvert, D.; Baker, D. Computational Design of Catalytic Dyads and Oxyanion Holes for Ester Hydrolysis. J. Am. Chem. Soc. 2012, 134, 16197−16206.

10523

DOI: 10.1021/acs.jpcb.5b05795 J. Phys. Chem. B 2015, 119, 10511−10523