16
Combining Immune Checkpoint Inhibitors and Kinase-Inhibiting Supramolecular Therapeutics for Enhanced Anticancer Ecacy Ashish Kulkarni, ,,§ Siva Kumar Natarajan, §,Vineethkrishna Chandrasekar, §,Prithvi Raj Pandey, ,and Shiladitya Sengupta* ,,,§,Department of Medicine, Harvard Medical School, Boston, Massachusetts 02115, United States Harvard-MIT Division of Health Sciences and Technology, Cambridge, Massachusetts 02139, United States § Division of Engineering in Medicine, Department of Medicine, Brigham and Womens Hospital, Boston, Massachusetts 02115, United States India Innovation Research Center and Invictus Oncology, New Delhi 110092, India Dana Farber Cancer Institute, Boston, Massachusetts 02115, United States * S Supporting Information ABSTRACT: A major limitation of immune checkpoint inhibitors is that only a small subset of patients achieve durable clinical responses. This necessitates the develop- ment of combinatorial regimens with immunotherapy. However, some combinations, such as MEK- or PI3K- inhibitors with a PD1-PDL1 checkpoint inhibitor, are pharmacologically challenging to implement. We rational- ized that such combinations can be enabled using nanoscale supramolecular targeted therapeutics, which spatially home into tumors and exert temporally sustained inhibition of the target. Here we describe two case studies where nanoscale MEK- and PI3K-targeting supramolecular therapeutics were engineered using a quantum mechanical all-atomistic simulation-based approach. The combinations of nanoscale MEK- and PI3K-targeting supramolecular therapeutics with checkpoint PDL1 and PD1 inhibitors exert enhanced antitumor outcome in melanoma and breast cancers in vivo, respectively. Additionally, the temporal sequence of administration impacts the outcome. The combination of supramolecular therapeutics and immunotherapy could emerge as a paradigm shift in the treatment of cancer. KEYWORDS: cancer, supramolecular therapeutics, immune checkpoint inhibitors, targeted therapy A mong the most promising approaches in the treatment of cancer is the blockade of immune check points. 1,2 Cancer progression occurs as a result of the ability of cancer cells to overcome immunosurveillance. 3,4 Molecules such as the programmed cell death ligand 1 (PDL1), expressed on cancer cells, can ligate with the PD1 receptor that is expressed on cell surface of T cells and cause T cell anergy. 5 A number of antibodies that disrupt the PD1-PDL1 immune checkpoint are currently in the clinics. 6 However, the percentage of patients who show durable response to immunotherapy is small. 7 This stems from the strong immunosuppressive environment within the tumor, and also potential intrinsic, acquired or drug-induced resistance mechanisms that can limit the activity of an immunotherapy agent. 8 As a result, it is increasingly being realized that a combination of an immunotherapy agent with a second therapeutic agent may be the optimal strategy for a maximal antitumor outcome. 911 A large body of data is evolving on the combinations of immune checkpoint inhibitors with chemotherapy, radio- therapy, or targeted therapies. 916 The emerging paradigm is however to move from empirical combinations to one of rational design, based on the compatibility of mechanisms that can act synergistically. 16 For example, there is currently intensive investigation into PD1-PDL1 inhibition in combina- tion with either BRAF inhibitors alone or with BRAF- and MEK-inhibitors. This is based on the observations that vemurafenib, a BRAF inhibitor, increases the responsiveness of T cells against melanoma by enhancing the expression of T Received: March 5, 2016 Accepted: September 22, 2016 Published: September 22, 2016 Article www.acsnano.org © 2016 American Chemical Society 9227 DOI: 10.1021/acsnano.6b01600 ACS Nano 2016, 10, 92279242

Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

Combining Immune Checkpoint Inhibitors andKinase-Inhibiting Supramolecular Therapeuticsfor Enhanced Anticancer EfficacyAshish Kulkarni,†,‡,§ Siva Kumar Natarajan,§,∇ Vineethkrishna Chandrasekar,§,∇ Prithvi Raj Pandey,∥,∇

and Shiladitya Sengupta*,†,‡,§,⊥

†Department of Medicine, Harvard Medical School, Boston, Massachusetts 02115, United States‡Harvard-MIT Division of Health Sciences and Technology, Cambridge, Massachusetts 02139, United States§Division of Engineering in Medicine, Department of Medicine, Brigham and Women’s Hospital, Boston, Massachusetts 02115,United States∥India Innovation Research Center and Invictus Oncology, New Delhi 110092, India⊥Dana Farber Cancer Institute, Boston, Massachusetts 02115, United States

*S Supporting Information

ABSTRACT: A major limitation of immune checkpointinhibitors is that only a small subset of patients achievedurable clinical responses. This necessitates the develop-ment of combinatorial regimens with immunotherapy.However, some combinations, such as MEK- or PI3K-inhibitors with a PD1-PDL1 checkpoint inhibitor, arepharmacologically challenging to implement. We rational-ized that such combinations can be enabled using nanoscalesupramolecular targeted therapeutics, which spatially homeinto tumors and exert temporally sustained inhibition of thetarget. Here we describe two case studies where nanoscaleMEK- and PI3K-targeting supramolecular therapeutics were engineered using a quantum mechanical all-atomisticsimulation-based approach. The combinations of nanoscale MEK- and PI3K-targeting supramolecular therapeutics withcheckpoint PDL1 and PD1 inhibitors exert enhanced antitumor outcome in melanoma and breast cancers in vivo,respectively. Additionally, the temporal sequence of administration impacts the outcome. The combination ofsupramolecular therapeutics and immunotherapy could emerge as a paradigm shift in the treatment of cancer.

KEYWORDS: cancer, supramolecular therapeutics, immune checkpoint inhibitors, targeted therapy

Among the most promising approaches in the treatmentof cancer is the blockade of immune check points.1,2

Cancer progression occurs as a result of the ability ofcancer cells to overcome immunosurveillance.3,4 Moleculessuch as the programmed cell death ligand 1 (PDL1), expressedon cancer cells, can ligate with the PD1 receptor that isexpressed on cell surface of T cells and cause T cell anergy.5 Anumber of antibodies that disrupt the PD1-PDL1 immunecheckpoint are currently in the clinics.6 However, thepercentage of patients who show durable response toimmunotherapy is small.7 This stems from the strongimmunosuppressive environment within the tumor, and alsopotential intrinsic, acquired or drug-induced resistancemechanisms that can limit the activity of an immunotherapyagent.8 As a result, it is increasingly being realized that acombination of an immunotherapy agent with a second

therapeutic agent may be the optimal strategy for a maximalantitumor outcome.9−11

A large body of data is evolving on the combinations ofimmune checkpoint inhibitors with chemotherapy, radio-therapy, or targeted therapies.9−16 The emerging paradigm ishowever to move from empirical combinations to one ofrational design, based on the compatibility of mechanisms thatcan act synergistically.16 For example, there is currentlyintensive investigation into PD1-PDL1 inhibition in combina-tion with either BRAF inhibitors alone or with BRAF- andMEK-inhibitors. This is based on the observations thatvemurafenib, a BRAF inhibitor, increases the responsivenessof T cells against melanoma by enhancing the expression of T

Received: March 5, 2016Accepted: September 22, 2016Published: September 22, 2016

Artic

lewww.acsnano.org

© 2016 American Chemical Society 9227 DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

Page 2: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

cell antigens.16,17 Similarly, the PI3K-AKT pathway, which ishighly dysregulated in breast cancer, has also been implicated inconferring resistance to T cell-mediated killing throughincreased expression of anti-apoptotic proteins.18,19 BothPI3K and MAPK pathways have been reported to regulatethe expression of PDL1, although the effect is variable acrossdifferent cancers.20,21 Together, these observations indicate thattargeted therapeutics that inhibit PI3K or MEK could berationally combined with immunotherapies that modulate Tcell functions. However, concerns exist about the potentialcumulative toxicities with these combinations.11,16,22 Further-more, both PI3K and MAPK pathways are implicated in theactivation of T cells. PDL1-PD1 ligation downregulates thesepathways in the T cells, leading to T cell exhaustion.23,24 Thus,a systemic shutdown of these pathways could potentially limitthe effect of a PD1-PDL1 inhibition. We rationalized that thepreferential targeting of the inhibitors of these aberrantpathways to the cancer cells could emerge as an alternativeapproach to enable a combination regimen with T cellactivating treatments.Nanoscale technologies allow the preferential delivery of

therapeutic payloads to tumors by evading the systemic

immune system.25,26 While nanoparticles have recently beenused as vaccines or as mimics of professional antigen-presentingcells to induce an immune reaction against the cancer cells,27

combination regimens of immunotherapy with nanoparticlesthat target specific signaling pathways in cancer are yet to bestudied. We have previously demonstrated that a polymericnanoparticle that targeted MAPK could preferentially accumu-late in the tumor and exert an enhanced antitumor effect.28 Inanother study, we demonstrated that nanoparticles can exert asustained inhibition of PI3K signaling in the tumor in vivo.29,30

Taken together, these pharmacodynamics properties and thepreferential biodistribution of MEK- and PI3K-targetingnanoparticles to the tumors could potentially be harnessedfor combining with PD1-PDL1 checkpoint inhibitors.As described by Jean-Marie Lehn, supramolecular assembly

defines the assembly of complex structures from molecularsubunit via weak interactions, such as coordination bonds, vander Waal’s forces, hydrophobic−hydrophilic bonds, etc.31 In arecent study, we have demonstrated that a quantum mechanical(QM) all-atomistic simulation-based algorithm could be usedto engineer molecular building blocks that are pharmacologi-cally active and can self-assemble into nanoscale structures via

Figure 1. Computer-aided design of a MEK-inhibiting supramolecular therapeutic. (a) Selumetinib forms highly unstable lipid nanoparticlesand precipitates out within 4 h of synthesis. (b) MD-simulation optimized structure of selumetinib. (c) Snapshot of all-atomistic simulation ofselumetinib interacting with SOPC lipid bilayer at 500 ns shows aggregates of selumetinib at the interface of hydrophilic lipid head and waterinterface. Lipid hydrophilic heads are shown in orange and blue spheres, and lipid tails are shown in gray color. Selumetinib is shown in redcolor. (d) MD simulation optimized structure of supratinib (molecular subunit of supramolecular nanostructure STB). (e) Snapshot of all-atomistic simulation of supratinib interacting with SOPC lipid bilayer at 500 ns shows stable supramolecular structure. Lipid hydrophilicheads are shown in orange and blue spheres, and lipid tails are shown in gray color. Supratinib is shown in red color. (f) Schematicrepresentation for tilt angle calculation. Tilt angle is the angle between vector joining center of mass of phospholipid tails and z-axis (axisperpendicular to bilayer plane). Value of tilt angle is positive or negative depending on direction of the ripple; its value is close to 0° when noripples form. (g) Distribution of tilt angle averaged over last 5 ns of MD trajectory. Broader the distribution larger is the tilt angle, hence,higher is the extent of bilayer instability. (h) Schematic representation of angle (θ) between vector defined on C−C bond on SOPC tail withZ-axis (axis perpendicular to bilayer plane). Deuterium order parameter (SCD) is calculated using θ. SCD is calculated on each carbon atom ofphospholipid tail. The higher the SCD, the higher is the lipid tail ordering. (i) Deuterium order parameter on each methylene group onsaturated tail of SOPC is depicted. Lipid tail ordering for selumetinib-containing SOPC bilayer is the least. Lipid tail ordering of supratinib-containing SOPC bilayer is similar to pure SOPC bilayer. (j) Probability of finding water molecules per unit volume was calculatednormalizing the count per unit volume with number of frames in last 5 ns of MD trajectory. (k) High-resolution cryo-TEM image of STBsupramolecular therapeutic showing size of ∼100 nm and spherical morphology (scale bar =100 nm).

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9228

Page 3: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

supramolecular assembly. These nanoscale supramolecularstructures overcome the challenges of stability and drug loadingassociated with traditional lipid-based nanoparticles.32 Here wedescribe the use of this computational algorithm to designinhibitors of MEK and PI3K, which act as molecular subunits inthe assembly of supramolecular nanostructures. We demon-strate that the nansocale supramolecular therapeutics preferen-tially accumulate in the tumor, and the combination of theMEK-inhibiting supramolecular nanoparticle with a PDL1inhibitor and the combination of the PI3K-inhibiting supra-molecule with a PD1 blockade can result in an enhancedantitumor outcome in in vivo models of melanoma and breastcancer, respectively. The combination of nanoscale supra-molecular targeted therapeutics and immunotherapy canemerge as a paradigm shift in the treatment of cancer.

RESULTS AND DISCUSSION

Engineering a MEK-Targeting Nanoscale Supramo-lecular Therapeutic. Classical approaches toward engineering

a nanoparticle, such as a liposome, involves the entrapment ofthe drug in lipid or polymeric matrices. However, attempts toentrap selumetinib, a clinical stage MEK inhibitor, in aliposome (at 20 mol % of drug) failed with the drugprecipitating out within 4 h (Figure 1a). Indeed, the stabilityof nanoparticles at drug concentrations required for clinicalefficacy has been a major challenge in the clinical translation ofnanomedicines. We rationalized that this hurdle could beovercome by designing a molecular subunit that facilitates theassembly into a nanoscale structure through supramolecularinteractions while retaining its biological MEK-inhibitingactivity.As the first step, we used a QM all-atomistic simulation-based

platform technology32 to understand the mechanism under-lying the failure of selumetinib to stay in the lipid-basednanoparticle. As shown in Figure 1b, we first obtained energy-minimized structures of selumetinib using QM optimizationand a molecular dynamics (MD) simulation in vacuum. Inparallel, we designed an analogue, which we term supratinib, by

Figure 2. Characterization of MEK-inhibiting supramolecular therapeutics. (a) Graph shows distribution of hydrodynamic diameter of MEK-inhibiting supramolecular therapeutics (STB) as measured using DLS. (b) Graph shows the changes in size and ζ potential of STB over timeas a measure of physical stability of STB during storage condition at 4 °C. Absence of change reflects structural integrity. (c) Graph showsrelease kinetics profiles of the drug from STBs in PBS (pH 7.4) or cell lysate. Data shown are mean ± SEM (at least triplicates at each timepoint). The drug is released in a sustained manner in PBS, whereas a rapid release is observed in cell lysate, consistent with the fasterdegradation of supramolecular theraeutics under enzymatic conditions. (d) Graphs show effect of increasing concentrations of STB orselumetinib on viability of A549 lung cancer cells and B16-F10 melanoma cells. (e) Table shows the IC50 values of selumetinib and STBs inA549 and B16-F10 cells. (f) Western blot shows expression of phospho and total ERK 1/2 in B16-F10 cells at different time points post-treatment with either 5 μM of selumetinib or STB. Graph shows internalization of selumetinib and STBs at 3 and 18 h in (g) A549 lung cancercells and (h) B16-F10 melanoma cells. Statistical analysis was performed with student t test. Data show mean ± SEM; n.s., not significant; **p< 0.01; ***p < 0.001.

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9229

Page 4: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

Figure 3. Combination of MEK-inhibiting supramolecular therapeutic and PDL1-immune checkpoint inhibitor increases antitumor efficacy ina B16-F10 melanoma model. (a) Graph shows the effect of MEK-inhibiting supramolecular therapeutic (STB) treatments on tumor volume inB16-F10 tumor bearing mice as compared to selumetinib- and vehicle-treated groups (n = 4). Each animal was injected intravenously withfour doses of vehicle (as control group), 5 mg/kg of selumetinib, or STB (equivalent to selumetinib dose). (b) Western blot showingexpressions of phospho and total ERK 1/2, and β-actin for normalization, in tumors after different treatments. Graph shows quantification ofexpressions of pERK/Total ERK in different treatment groups after normalization with β-actin levels. Statistical analysis was performed withone-way ANOVA test. Data shows mean ± SEM; n.s., not significant; ***p < 0.001. (c) Representative images show biodistribution of NIR-dye labeled STBs in tumor-bearing mice at different time points post-i.v. administration. The tumor position is shown dotted circle. (d) Graphshows quantification of NIR-dye labeled STBs in different organs. Data represents mean ± SEM (n = 3−4); **p < 0.01 vs other organs. (e)Graph shows changes in tumor volume after four doses of selumetinib (5 mg/kg), 2.5 mg/kg αPD-L1, selumetinib (5 mg/kg) + αPD-L1 (2.5mg/kg), or STB (5 mg/kg) + αPD-L1 (2.5 mg/kg) on each alternate days. First day of treatment was considered as day 0. Tumor volumeswere measured on every alternate day for 10 days. Statistical analysis was performed with one way ANOVA test. Data shown are mean ± SEM

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9230

Page 5: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

derivatizing selumetinib with a cholesterol-based tether throughan optimal linker. This was based on previous findings thatcholesterol can affect a bilayer by decreasing its free volume,increasing lipid tail order and packing, and reducing the lipidmobility.33 Indeed, in a previous study with taxanes, we haveobserved that such molecular subunits can facilitate theassembly into a stable supramolecular nanostructure. Theenergy-minimized structure of supratinib using QM optimiza-tion and a short MD simulation in vacuum is shown in Figure1c. We next simulated the interaction of a high (20 mol %)concentrations of selumetinib or supratinib with a modelcolipid, 1-octadecanoyl-2-(9Z-octadecenoyl)-sn-glycero-3-phos-phocholine (SOPC). The force field parameters for SOPC wereadapted from an atomistic CHARMM-based force fieldproposed for phosphatidylcholine lipids34,35 and using aTIP3P water model.36,37 Models were constructed such thatthe system contained 256 molecules of SOPC and 64 moleculesof selumetinib or supratinib, i.e., to mimic 20 mol % of the drugin the lipid bilayer. We have previously observed that threedistinct parameters: (1) ripple formation in the bilayer, (2)decreased lipid tail ordering, and (3) the probability of watermolecules entering the bilayer are good predictors ofinstability.32 The unbiased atomistic simulations did not revealany significant ripple formations for either selumetinib- orsupratinib-containing bilayers (Figure 1c, and e, SupplementaryMovies 1 and 2), quantified as the angle between vector joiningthe center of mass of lipid tails and the phosphorus (P) atomon the headgroup of SOPC molecule with z-axis (bilayernormal), where the greater the angle, the higher is the degree ofinstability (Figure 1f and g). However, selumetinib did decreasethe ordering of the aliphatic tail groups that was quantified bycalculating the deuterium order parameter (SCD) for the SOPClipid saturated tail from last 5 ns of the trajectory (Figure 1hand i), which could explain the instability seen in experimentalobservations. In contrast, simulating the interaction of a high(20 mol %) concentrations of supratinib with SOPC revealed astable supramolecular assembly, characterized by absence ofripple formation (Figure 1e−g), a lipid tail ordering similar tothe SOPC-only bilayer (Figure 1h and i), and no permeation ofwater into the supramolecular structures (Figure 1j).Given that supratinib-based supramolecular nanostructure

(STB) satisfied all three parameters of stability in silico, we nextsynthesized (Supplementary Figure 1) and tested its ability toform a stable supramolecular nanostructure with colipidexcipients using a thin-film hydration technique. Indeed, asshown in Figure 1k, cryo-transmission electron microscopy(cryo-TEM) revealed that supratinib was able to form stable

supramolecular nanostructures (STB) with an average meandiameter of 108.6 ± 27 nm and loading efficiency of 87 ± 5%.

Characterization of the MEK-Inhibiting Supramolecu-lar Nanotherapeutic (STB). We next physicochemicallycharacterized the MEK-inhibiting supramolecular therapeutic(STB). Dynamic laser light scattering revealed that thehydrodynamic diameter of STB was 191 ± 38 nm. Thedifference in dimension compared with the findings fromelectron microscopy is consistent with PEGylation of thestructures (Figure 2a). The ζ potential was measured to be 18.4± 6.48 mV. Both the radius and the ζ potential remained withina constant range over a sustained period of time at 4 °C orwhen incubated with serum, suggesting that the structures werestable and that the surface PEGylation prevents theopsonization by serum proteins (Figure 2b, SupplementaryFigure 2a and b). Incubating STB with cancer cell lysate, tomimic the tumor environment, resulted in a sustained andcomplete release within a 7 day period (Figure 2c). Similarly, afaster release was observed when incubated at an acidic pH(Supplementary Figure 2c). In contrast, limited release of thedrug was observed at physiological pH 7.4 (Figure 2c) or inlysate from a normal splenocyte (Supplementary Figure 2d).We next tested the efficacy of STB in melanoma B16/F10

and lung carcinoma A549 cell lines in vitro. As shown in Figure2d and e, STB was more potent than selumetinib in reducingcell viability of both cell lines, as evident from the lower IC50values. At the same concentration range, neither selumetinibnor STB had any significant inhibitory effect on the viability ofT cells (Supplementary Figure 3a). Furthermore, as shown inSupplementary Figure 3b, treatment with STB resulted in anincrease in PDL1 expression of B16/F10 melanoma cells. Wedid not observe a similar increase in PD1 expression on T cellsin response to STB (Supplementary Figure 3c). Additionally,Western blotting revealed that treatment of melanoma cellswith selumetinib resulted in an early inhibition of ERKphosphorylation. However, this inhibition was not long lasting.In contrast, we observed a sustained inhibition of ERKphosphorylation following STB treatment (Figure 2f). A similarsustained inhibition of phosphorylation of ERK was alsoobserved when A549 cells were incubated with STB(Supplementary Figure 3b). Furthermore, the quantificationof intracellular drug concentration revealed that while at anearly time point (3 h post-treatment) both selumetinib andSTB attained similar levels, the latter resulted in a greater drugbuild up at a later time point (Figure 2g and h), which couldexplain the sustained inhibition of ERK phosphorylation andthe increased efficacy.

Figure 3. continued

(n = 3−4); n.s., not significant; *p < 0.05. STB in combination with immune checkpoint inhibitor antibody PDL1 showed the most tumorgrowth inhibition as compared to other treatment groups. (f) Graph shows overall body weight of the animals as a measure of systemic drugtoxicity. (g) Western blot showing differences in expression levels of phospho and total ERK 1/2 in tumors after treatments with PD-L1antibody and STB + PD-L1 antibody combination. β-actin was used for normalization. (h) IHC of tumor tissues shows p-ERK levels indifferent treatment groups; antileukocyte common antigen (CD45) antibody staining (red) of tumor sections in different treatment groupsshows increase in TIL infiltration in STB + PD-L1 group compared to other treatments. Nuclei were stained with DAPI. Representativeimages of tumor sections from animals treated as above were labeled for apoptosis using TUNEL (red) and counterstained with DAPI (blue).(i) Graph shows the quantification of relative TIL infiltration in DAPI-labeled tumor sections as a percentage of CD45+ve cells to total nucleistaining. (j) Graph shows the quantification of apoptosis from the labeled tumor sections as a percentage of TUNEL+ve cells as a function oftotal nuclei. Data shown are mean ± SEM (n = 3−4); ANOVA followed by Newman Keuls post-hoc test (*p < 0.05, **p < 0.01). (k)Representative FACS data from tumors of B16-F10 tumor bearing mice after different treatments. The left panel shows percentage of CD4+ vsCD8+ cells in the isolated lymphocyte population from tumors. The right panel shows the percentage of activated CD8+ cells(CD44HiCD62LLo) in the tumor. (l) Graph shows quantification of number of activated CD8+ T cells (CD44HiCD62LLo) per gram of tissuefrom cell sorted FACS plots of tumors from the above study. (m) Graph shows percentage of Foxp3+ regulatory T cells from CD4+ cells.

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9231

Page 6: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

Combination with a PDL1 Antibody in Vivo. Based onthe in vitro observations, we next tested the efficacy of STB in asyngeneic murine melanoma model. A large subset ofmelanoma presents with constitutive activation of the MAPKpathway, underpinning the use of MEK inhibitors in thetreatment of melanoma.38 Consistent with our in vitro findings,we observed an increase in the antitumor efficacy with STBcompared with selumetinib (Figure 3a). It should be noted thathere we used submaximal doses so that we could dissectwhether the combination chemotherapy results in enhancedefficacy. Western blotting of in vivo tumor analytes revealed asignificantly greater inhibition of MAPK signaling (i.e., reducedphosphorylated ERK levels) with STB treatment as comparedwith vehicle- or selumetinib-treated control groups, consistentwith the observed efficacy (Figure 3b).

Tumors that overexpress PDL1 are reported to respond wellto PDL1 immune checkpoint inhibitors.39 Based on our priorobservation that STB induces PDL1 expression in melanoma,we rationalized that a combination of STB and a PDL1-immune checkpoint inhibitor could result in an increasedantitumor outcome. However, previous studies have reportedthat T cell activation is maintained through ERK signaling,40

and MEK inhibitors can adversely affect T cell proliferation,cytokine production, and antigen-specific expansion,24 suggest-ing that for a MEK inhibitor to be rationally combined with animmune check point inhibitor, it should ideally be preferentiallydelivered to the tumor. Although our in vitro studies showedthat STB has no effect on T cell proliferation at therapeuticconcentrations, we first tested if the supramolecular nanostruc-tures can preferentially target the tumor. Tumor-bearing mice

Figure 4. Computer simulations predicts stable PI3K-inhibiting supramolecular nanostructures. (a) MD simulation optimized structure ofPI103. (b) Snapshot of all-atomistic simulation of PI103 interacting with SOPC lipid bilayer at 500 ns shows aromatic stacking-mediatedaggregates of PI103 coming out of the bilayer. Lipid hydrophilic heads are shown in orange and blue spheres, and lipid tails are shown in graycolor. PI103 is shown in red color. (c) MD simulation optimized structure of SPI molecular subunit. Bottom panel shows snapshot of all-atomistic simulation of SPI molecular subunit interacting with SOPC lipid bilayer at 500 ns, which shows a stable supramolecular structure.(d) Graph depicts the cluster size distribution of PI103 and SPI molecular subunit. PI103 shows higher occurrence of larger cluster sizes ascompared to SPI molecular subunit. (e) Graph shows the estimation of number of pairs of PI103 and SPI molecular subunits stacking in lipidbilayer. (f) Deuterium order parameter on each methylene group on saturated tail of SOPC is depicted. Lipid tail ordering for PI103-containing SOPC bilayer is least. Lipid tail ordering of SPI-containing SOPC bilayer is higher than pure SOPC bilayer, which shows that SPImolecular subunit stabilizes the supramolecular structure. (g) Probability of finding water molecules per unit volume was calculatednormalizing the count per unit volume with number of frames in last 5 ns of MD trajectory. (h) Distribution of tilt angle averaged over last 5ns of MD trajectory. (i) High-resolution cryo-TEM of supramolecular structure (scale bar = 100 nm).

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9232

Page 7: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

were injected with a NIR dye-tagged supramolecularnanostructure and imaged at different time points. At 48 hpost-administration, the animals were sacrificed, and thebiodistribution of the supramolecular nanostructure wasnoted. As shown in Figure 3c, the supramolecular nanostruc-tures were found to progressively accumulate in the tumor,reaching saturation by 24 h. Additionally, a negligible quantityof the nanoparticles distributed to normal tissues like lungs,heart and kidneys, and the amount/g of tissue was still lower inthe major reticuloendothelial system organs such as liver andspleen as compared with the levels in the tumor (Figure 3d),consistent with previous findings.29 Furthermore, an analysis ofthe liver and kidney of animals treated with STB or selumetinibexhibited similar levels of cleaved caspase 3, a marker ofapoptosis, compared with vehicle-treated control animals(Supplementary Figure 4a). Taken together, these resultsindicated that STB could emerge as an attractive approach topreferentially inhibit MEK in the tumor.We next studied whether the combination of a PDL1

inhibitor and the STB exerted a greater antitumor efficacycompared with the immune checkpoint inhibition alone.Melanoma-bearing animals were injected with three cycles ofPDL1 antibodies (2.5 mg/kg) alone or in combination withMEK-inhibiting supramolecular nanoparticles (dose equivalentto 5 mg/kg of selumetinib). As a control, we also included agroup where the tumor-bearing animals were treated with acombination of selumetinib and PDL1 antibodies. As shown inFigure 3e, we observed that while PDL1 inhibition did inhibittumor growth, a significantly enhanced antitumor effect wasevident when it was combined with STB. In contrast, we didnot see a similar increase in efficacy when the PDL1 inhibitorwas combined with selumetinib. We did not observe anychange in body weights following the treatments (Figure 3f).To mechanistically dissect the enhanced efficacy seen with

the combination, we performed Western blotting of tumoranalytes, which revealed that the treatment with STB resultedin a decrease in the phosphorylated ERK levels as comparedwith PDL1 inhibition alone (Figure 3g). Immunolabeling thetumor sections for phospho-ERK confirmed that thephosphorylated ERK levels were significantly lower in thetumors treated with the combination of STB and PDL1inhibitor as opposed to the latter alone (Figure 3h).Additionally, labeling the tumor sections for CD45, a markerfor lymphocytes, or using H&E staining (Supplementary Figure4b) revealed that the checkpoint inhibition increased thenumber of tumor infiltrating lymphocytes (TILs). The levelwas significantly enhanced when the immune checkpointinhibitor was combined with STB as compared to thecombination of PDL1 antibody and selumetinib (Figure 3hand i). Furthermore, an analysis of the tumor sections forapoptosis using TUNEL revealed that the treatment with STBinduced significantly greater cell death as compared withselumetinib (Figure 3h and j). The combination of STB and theimmune checkpoint inhibition resulted in greater tumorapoptosis as compared with PDL1 inhibition alone or incombination with selumetinib. Taken together, the mechanisticobservations were consistent with the tumor growth inhibition.We next validated these results by analyzing the tumor

immune cells using fluorescence activated cell sorting (FACS).As shown in Figure 3k and l, the treatment with STB as well asthe PDL1 inhibitor increased the intratumoral ratio of CD8cells to CD4 cells as compared with selumetinib or vehicletreatments. The combination of STB with the PDL1 inhibitor

further increased CD8+ T cells in the tumor compared with thecombination of the immune checkpoint inhibitor withselumetinib. Additionally, quantifying CD44HiCD62Lo (acti-vated) fraction of the TILs revealed that treatment with thecombination of the PDL1 inhibitor and STB resulted in agreater percentage of activated TILs as compared with thecombination of PDL1 inhibitor and selumetinib (Figure 3l).None of the treatments altered the levels of TRegs in the tumor(Figure 3m and Supplementary Figure 4c). Taken together,these findings indicate that a MEK-inhibiting supramoleculartherapeutic is not associated with T cell inhibitory effects andcan be rationally combined with immune check point inhibitorsfor enhanced antitumor efficacy. Additionally, the enhancedefficacy relies on the enhanced recruitment of activated TILs.

Engineering a PI3K-Targeting Supramolecular Ther-apeutic. While the above results suggested that a nanoscalesupramolecular therapeutic could preferentially accumulate intumors and enable a combinatorial approach with immunecheckpoint inhibitors, we rationalized that the findings neededto be validated in a different setting using a distinct target andtumor model. The PTEN/PI3K pathway is a major oncogenicdriver implicated in 30−40% of primary breast tumors.41

However, PI3K is also the dominant downstream signalingpathway that is needed for the activation and proliferation of Tcells.42 The ligation of PD1 by PDL1 leads to inhibition ofPI3K signaling, resulting in T cell exhaustion,43 indicating that aPI3K inhibitor should ideally be preferentially targeted to thetumor to be combined rationally with an immune checkpointinhibitor, as systemically they can be counteractive.PI103 (Figure 4a) is a well-established inhibitor of class IA

PI3Ks as well as mTOR.44 However, like selumetinib, itprecipitates out of a lipid-based nanoparticle at concentrationsrequired for clinical translation.29 To understand theincompatibility of PI103 with formulation in a classicalliposome, we ran an all-atomistic simulation of PI103interacting with colipids in a bilayer, which revealed that thePI103 molecules precipitate out into the water layer afterstacking with each other to form clusters (Figure 4b andSupplementary Movie 3). We rationalized that the approach weused for selumetinib, i.e., bioconjugation with cholesterol toform a molecular subunit, could be applied to PI103 to preventstacking and facilitate a stable supramolecular nanoscaleassembly. The molecular subunit, indeed formed a stablesupramolecular structure, which we termed as SPI (Figure 4cand Supplementary Movie 4). To compute stacking, a vectorwas defined among the atoms shown on PI103 molecule. Theangle between this vector on different molecules was calculatedwithin a distance of 0.5 nm. Stacking was considered if theangle was >125° or <75°. The count of such stacked pairs wasplotted for each frame in the last 20 ns of the trajectory. Thenumber of molecules stacked was found to be higher for PI103as compared to the SPI system (Figure 4d and e). Additionally,we observed that 20 mol % PI103 system reduced the lipid tailordering, while it was significantly elevated in the 20 mol % SPIsystem compared with a pure SOPC (Figure 4f). This wasagain reflected in the calculation of number of water moleculesper unit volume in the lipid bilayer, i.e., SPI showed the leastperforation among all the systems (Figure 4g). It should benoted that neither system exhibited any ripple formation(Figure 4h). Consistent with the theoretical model, SPIresulted in the formation of extremely stable nanoparticles of∼150 nm diameter (Figure 4i) with a loading efficiency of 65 ±10%, which were stable even in the presence of serum proteins

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9233

Page 8: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

Figure 5. Combination of PI3K-inhibiting supramolecule with PD1 antibody enhances antitumor efficacy in 4T1 breast cancer model. (a)Growth curves show effect of PI3K-inhibiting supramolecule (SPI) treatment on tumor volume in 4T1 tumor bearing mice as compared toPI103- and vehicle-treated groups. Each animals were injected intravenously with three doses of either vehicle (for control group), 2.5 mg/kgof PI103, or SPI (∼2.5 mg/kg of PI103). SPI shows statistically significant tumor growth inhibition as compared to PI103. Data shown aremean ± SEM (n = 5); ***p < 0.001. (b) Western blot showing expressions of phospho and total Akt and mTOR in tumors after differenttreatments. Graph shows quantification of expressions of pAkt/Total Akt in different treatment groups after normalization with β-actin levels.Statistical analysis was performed with one-way ANOVA test. Data show mean ± SEM; *p < 0.05; **p < 0.01. (c) Graph shows changes intumor volume after three doses of αPD-1 (5 mg/kg), PI103 (2.5 mg/kg) + αPD-1 (5 mg/kg), or SPI (2.5 mg/kg) + αPD-1 (5 mg/kg) onalternate days. First day of treatment was considered as day 0. Tumor volumes were measured on every alternate day for 17 days. End pointfor each animal was tumor size >4000 mm3 or tumor ulceration or necrosis or animal death. Statistical analysis was performed with one way

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9234

Page 9: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

Figure 5. continued

ANOVA test. Data show mean ± SEM; **p < 0.01; ***p < 0.001. SPI in combination with immune checkpoint inhibitor antibody PD1 showsthe greatest tumor growth inhibition as compared to other treatment groups. (d) Graph shows effect of treatments on overall body weight as ameasure of any systemic toxicity. (e) Representative FACS data from 4T1 tumors after different treatments (day 17). The left panel showspercentage of CD4+ vs CD8+ cells in the isolated lymphocyte population. The right panel shows the percentage of activated CD8+ cells(CD44HiCD62LLo). (f) Representative FACS data from spleen of 4T1 tumor bearing mice after different treatments (day 17). The left panelshows percentage of CD4+ vs CD8+ cells in the isolated lymphocyte population from spleen. The right panel shows the percentage ofactivated CD8+ cells (CD44HiCD62LLo) in the spleen. (g) Graph shows the number of CD8+ cells in 4T1 tumors (normalized to tumor mass)after different treatments as quantified using FACS. (h) Graph shows the number of activated CD8+ cells in 4T1 tumors (normalized to tumormass) after different treatments as quantified using FACS. Data represents mean ± SEM (n = 3); *p < 0.05 (ANOVA with Newman Keulspost-hoc test). (i) Graph shows the ratio of activated CD8+ T cells to naiv̈e T cells in spleen in 4T1-tumor bearing mice after differenttreatments. Data represent mean ± SEM (n = 3); *p < 0.05 (ANOVA with Newman Keuls post-hoc test).

Figure 6. Efficacy of PI3K-inhibiting supramolecule and PD1 antibody combination is enhanced by temporally sequencing the administration.(a) Schematic shows the sequence of dosing schedules of PI3K-inhibiting supramolecule (SPI) and PD1 antibody combination. SPI wasadministered on days 0, 2, and 4 for both the schedules. PD-1 (immediate) represents the dosing schedule 1 where the PD-1 antibody wasinjected on every alternate day from day 1, and PD-1 (after) represents the dosing schedule 2 where PD-1 antibody is injected on days 6, 8,and 10, i.e. after the SPI-treatment cycle. (b) Graph shows tumor volumes of 4T1 tumor bearing mice after treatments with different doingschedules (day 16). The efficacy of SPI with PD-1 antibody combination is improved with dosing schedule 2 where PD-1 antibody isadministered after SPI treatment as compared to schedule 1. (c) Representative epifluorescent images of tumor sections from animals afterdifferent treatments (day 16), labeled for apoptosis using TUNEL (red) and counterstained with DAPI (blue). (d) Graph shows thequantification of apoptosis from the labeled tumor sections as a percentage of TUNEL+ve cells as a function of total nuclei. Statistical analysiswas performed with ANOVA with Newman Keuls post-hoc test. Data shown are mean ± SEM (n = 3−5); *p < 0.05. (e) Kaplan Meir curvesshowing effect of treatments on survival (n = 5). (f) Representative FACS data from 4T1 tumors after treatments with different dosingschedules (day 16). The FACS data show the percentage of activated CD8+ cells (CD44HiCD62LLo). (g) Graph shows the number of activatedCD8+ cells in 4T1 tumors (normalized to tumor mass) after different treatments (n = 6). Statistical analysis was performed with student t test.Data shown are mean ± SEM; *p < 0.05.

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9235

Page 10: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

(Supplementary Figure 5a and b) and exhibited limited drugrelease in lysate from normal splenocytes (SupplementaryFigure 5c).Combination of PI3K-Inhibiting Supramolecule and

PD1 Inhibitor in Vivo. We first tested the effect of PI103 andSPI on T cell viability in vitro. As shown in SupplementaryFigure 5d, PI103 was more toxic to T cells than SPI. In aseparate study, we treated naiv̈e T cells with PI103 or SPI (atconcentration equivalent 500 nM of PI103) for 36 h and thenactivated the cells using CD3/CD28 Dynabeads. Interestingly,both PI103 and SPI did affect the activation of T cells in thisstatic system, which mimics a systemic exposure (Supple-mentary Figure 5e and f). Interestingly, neither selumetinib norSTB exhibited a similar effect (Supplementary Figure 3a),which highlights the differences in the role of the PI3K andMAPK signaling in T cell activation. This finding furthervalidated that it may be necessary to preferentially target thePI3K inhibitor to the tumor for combining it with an immunecheckpoint inhibitor. Indeed, in a separate study, wedemonstrated that the PI3K-inhibiting supramolecular ther-apeutic (SPI) can preferentially accumulate in the tumor andenables a sustained inhibition of PI3K and downstreamtargets.29 Consistent with these findings, we observed thatSPI exerted a greater antitumor effect than PI103 (Figure 5a).Furthermore, a Western blot analysis of the tumors revealedthat the treatment with SPI results in a significantly greaterinhibition of the PI3K signaling pathway, i.e., reducedphosphorylation of Akt and downstream mTOR, as comparedwith PI103- or vehicle-treated tumors (Figure 5b).PI3K inhibitors have been reported to decrease the

expression of PDL1 in breast cancer,45 and hence a PDL1inhibitor is unlikely to be effective in such a setting. For thisreason, we used a PD1-blocking antibody as the immunecheckpoint inhibitor in this study for combining with the PI3Kinhibitors. In our previous study, we have shown that a higherconcentration of SPI can ablate the tumor.29 As a result, for thecurrent study, we used a submaximal dose of the SPI to be ableto dissect the effect of combining it with the PD1 inhibitor.Treatment with the anti-PD1 antibody inhibited breast tumorgrowth as compared with the vehicle-treated group. Interest-ingly, the combination of PI103 with PD1 inhibitor decreasedthe antitumor efficacy of the latter. In contrast, animals treatedwith the combination of SPI and PD1 inhibitor exhibitedgreater antitumor outcome as opposed to the PD1 inhibitor-treated animals (Figure 5c). We did not observe any changes inthe body weight as a result of drug treatment in any of thegroups (Figure 5d).We next studied the effect of different treatments on the

intratumoral recruitment and activation of T cells. As shown inFigure 5e and g, the different treatments did not significantlyaffect the total number of CD8+ T cells recruited to the tumor.However, an analysis of the activation status of the intratumoralCD8+ T cells by monitoring for CD44HiCD62LLo expressionrevealed that the groups treated with the PD1 inhibitor as wellas SPI exhibited a larger fraction of activated T cells comparedwith vehicle-treated controls (Figure 5e and h). Thecombination of the PD1 inhibitor with the PI3K-inhibitingsupramolecular therapeutic did not further increase the fractionof activated CD44HiCD62LLo CD8+ T cells. The fraction ofactivated T cells in the vehicle- and the PI3K inhibitor (PI103)-treated groups were found to be similar. We additionallymonitored the splenocytes for CD8 positive T cells counts aswell as the effect of treatment on the activation of CD8+ T

cells. As shown in Figure 5f and i, increased ratio of activatedCD8+ T cells to naiv̈e cells was observed following treatmentwith PD1 inhibitor and the PI3K-inhibiting supramoleculartherapeutic as compared with PI103 treatment.

Sequencing Drug Administration Affects Outcome.Recent reports have highlighted the importance of sequencingthe administration of cancer therapies on outcomes.30,46,47 Wetherefore used two distinct schedules to treat tumor-bearinganimals with a combination of PD1 antibody and SPI. In oneschedule, animals were injected with the immune checkpointinhibitor on alternate days of administration of SPI(“immediate” schedule). In the second schedule, the PD1inhibitor was administered after the cycle of SPI treatment wascomplete (“after” schedule) (Figure 6a). Interestingly, as shownin Figure 6b, the antitumor efficacy was significantly greaterwith the “after” schedule. These findings were validated byquantifying the degree of cell death within the tumors. Asshown in Figure 6c and d, the “after” schedule resulted inmaximal levels of intratumoral apoptosis as compared to the“immediate” schedule. Excitingly, the “after” schedule resultedin a significant increase in life span as compared to vehicle orthe regimen where both drugs were administered simulta-neously. It should be noted that a combination of PI103 andPD1 did not increase life span as compared with vehicle-treatedanimals (Figure 6e). Interestingly, we observed an increase inthe mean level of activated intratumoral CD8+ T cells with theafter schedule as compared with the immediate schedule(Figure 6f and g).

CONCLUSIONSCombinations of immunotherapy with targeted therapeutics,such as MEK or PI3K inhibitors, are currently the holy grails inthe management of cancer. Here we demonstrate that theability of nanoscale supramolecular therapeutics to preferen-tially home into tumors can be leveraged to rationally combinetargeted therapeutics with immunotherapy. We show that thecombinations of a nanoscale MEK-targeting supramolecularnanotherapeutics with a PDL1 antibody, and a PI3K-inhibitingsupramolecular nanotherapeutic with a PD1 inhibitor, canresult in increased antitumor efficacy in immunocompetentmelanoma and breast carcinoma models, respectively. Suchcombinations can shift the paradigm in the treatment of cancer.While nanoengineering can enhance the efficacy of

immunostimulatory molecules by altering their colocalization,biodistribution, and release kinetics,27 a paradigm shift isneeded in how nanomedicines are designed. Classicalapproaches that rely on entrapping a drug in the polymericnanoparticles or liposomes have had limited success withclinical translation due to the instability of the formednanostructures. Here we used a distinct approach, where theactive agents were first redesigned based on a QM all-atomisticsimulation to facilitate self-assembly into stable nanoscalesupramolecular structures. Computational simulations, such asthose based on coarse grain models, have recently been used tounderstand the interactions between the drug and excipientsthat can assemble into a nanostructure.48 However, all-atomisticsimulations can enable resolutions at the nanoscale. Indeed, ourresults revealed significant insights, for example, bothselumetinib and PI103 induced instability in the bilayer butthrough distinct mechanisms. It should be noted that suchatomistic resolutions can be challenging to obtain fromexperimental studies due to thermal fluctuations of soft matter.The correlation between the simulation and the experimental

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9236

Page 11: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

observations suggests that computer-aided design can indeedemerge as a powerful tool for the fabrication of stablenanostructures.Sterically stabilized nanoparticles have been reported to

preferentially home into tumors by leveraging the leakypathophysiology of the tumor blood vessels and result ingreater intratumoral concentrations of the active drug ascompared to when the drug is administered in free form.26,27

Our observation from the biodistribution experiment in thecurrent study, where the supramolecular nanotherapeuticspreferentially accumulated in the tumors, is consistent withthese previous reports.49 Furthermore, we observed that unlikepacking the drugs in a liposome, which result in instability, theredesign of the molecular subunit to facilitate supramolecularassembly increases the stability of the nanostructure at high mol%. In a recent study, we have demonstrated that this stability isnot a simple function of conjugating a drug to cholesterol andthat the linker can also influence the supramolecularassembly.32 This ability of the subunit to facilitate thesupramolecular assembly is a key distinction from fabricatinga classical drug-loaded liposome. Interestingly, the cholesterolderivatization does not mean that the drug partitions into thelipid raft-rich cell membrane. Confocal imaging of fluorophore-tagged cholesterol-based supramolecular nanostructures re-vealed that the structures are internalized into the cytoplasm(Supplementary Figure 6). Furthermore, we observed asustained release of the active agent in the tumor lysate andan accelerated release at an acidic pH. Interestingly, the releaserate was slower in physiological pH and lysates of normal cells.Indeed, previous studies have reported differences in esterasesand pH between normal and cancerous environments thatcould contribute to this difference.50 Taken together, theseattributes can lead to a higher and sustained concentration ofthe active drug in the cancer cells as compared to free drug.Indeed, this is validated by pharmacodynamics observations,where the nanoscale supramolecular therapeutics exerted aprolonged and greater inhibition of the target pathway as wellas greater antitumor efficacy compared to selumetinib or PI103treatment. Additionally, in a separate study, we have shown thatthe nanoscale supramolecular therapeutics are less toxicsystemically.29 In the current study, we did observe that thePI103 was more toxic against splenocytes than SPI. Takentogether, these observations support the hypothesis thatnanoscale supramolecular therapeutics could emerge as a betterchoice than current inhibitors of PI3K or MAPK for combiningwith immune checkpoint inhibitors.Interestingly, while both PI3K- and MEK-inhibiting supra-

molecular therapeutics resulted in increased antitumor efficacywhen combined with the immune checkpoint inhibitors, ourresults indicate that the underlying mechanisms driving such anoutcome are distinct. The combination of PDL1 inhibition withMEK-inhibiting supramolecular therapeutic resulted in anincrease in the total activated TILs and increased intratumoralcytotoxicity as compared with the animals treated with eitherthe PDL1 antibody or the MEK-inhibiting supramoleculartherapeutic alone. In a recent study, the triple therapycombination of trametinib, a MEK inhibitor, with a BRAFinhibitor and adoptive cell transfer was indeed shown toincrease T cell infiltration and improved in vivo efficacy.51 Therationale for combining a MEK inhibitor was to reduce thetoxicity associated with BRAF inhibitors attributed to theparadoxical activation of MAPK by the BRAF inhibitors inBRAF wild-type cells.52 The current findings suggest that the

ability of the supramolecular nanotherapeutic to preferentiallyhome into the tumor can offer an alternative and simplerapproach toward combination immunotherapy. Additionally,MEK inhibitors have also been reported to upregulatemelanocyte differentiation antigens and MHC class Iexpression, which could underlie the increased TILs seenwith STB treatment.22 In contrast, treatment with the PI3K-inhibiting supramolecular nanotherapeutic and the PD1inhibitor or the combination did not increase total CD8+ Tcell infiltration, but both significantly increased the fraction ofactivated CD44HiCD62LLo CD8+ TILs. We did not observeany further increase in TILs when the PD1 inhibitor wascombined with the PI3K-inhibiting supramolecular therapeutic,indicating that we had reached a maximal ceiling in the level ofactivated T cells. This suggests that the increased efficacyobserved in the case of simultaneous combination treatmentwith the PD1 inhibitor and the PI3K-inhibiting supramoleculartherapeutic is likely due to a direct anticancer effect as a resultof PI3K-inhibition, and not via the augmentation of the activityof the immune checkpoint inhibitor. However, we did observean increase in the efficacy as well as mean activated TILs whenthe sequence of administration of the combination wastemporally staggered. Indeed, previous reports have suggestedthat PI3K-AKT inhibitors can diminish the expression ofantiapoptotic proteins such as cIAP1 and 2, BCL-2, and BCL-XL, which can then sensitize the tumor cells to immune-mediated destruction.16 Indeed, SPI induced greater apoptosiscompared with the parent drug. Taken together with theobservation that PI3K-inhibiting supramolecule enables drugrelease in a sustained manner, it is logical that sufficient drugbuildup induces maximal apoptosis of cancer cells, which canthen amplify the outcome to immunotherapies. Thus,temporality or the sequence of administration is likely toemerge as a key parameter in the design of combinationregimens with immunotherapies.16 Interestingly, PI103 resultedin a decrease in the efficacy of the PD1-inhibitor, consistentwith the dominant role of PI3K signaling in the activation of Tcells.Several aspects of this study can facilitate future therapy in

humans. First, the ability of the nanoscale supramoleculartherapeutics to preferentially home into tumors indicates thatmany of the adverse effects that have emerged in recent clinicaltrials with combination of targeted therapy with immunother-apy could potentially be minimized. Second, the increasedintratumoral buildup of the active agent can result in asustained pharmacodynamic effect, which can increase theantitumor efficacy. Third, the versatility of the computer-aideddesign of the supramolecular nanotherapeutics means that arepertoire of such agents can be designed for additionalimmunomodulatory targets. Finally, the nanoscale supra-molecular therapeutics can indeed emerge as a natural choicefor enabling combination therapy with immune checkpointinhibitors that are pharmacologically incompatible, thusenabling a paradigm shift toward a comprehensive “integrative”approach in the treatment of cancer.

METHODSMaterials. All chemical reagents were of analytical grade and used

as supplied without further purification unless indicated. All reactionswere performed under inert conditions unless otherwise indicated.Anhydrous dichloromethane (DCM), methanol, cholesteryl hemi-succinate, dimethylamino pyridine (DMAP), succinic anhydride, 1-ethyl-3-(3-(dimethylamino)propyl) carbodiimide (EDC), L-α-phos-

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9237

Page 12: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

phatidylcholine, and sephadex G-25 were purchased from Sigma-Aldrich. PI103 and selumetinib were purchased from LC Laboratories.PD-1 and PD-L1 antibodies were purchased from BioXCell. 1,2-Distearoyl-sn-glycero-3-phosphoethanolamine-N-[amino(polythyleneglycol)2000] and the mini hand-held extruder kit (including 0.2 and0.4 μm polycarbonate membrane, Whatman filter supports, and 1.0mL Hamiltonian syringes) were bought from Avanti Polar Lipids Inc.1H NMR spectra were recorded on Bruker DPX 400 MHzspectrometer. Electrospray ionization mass spectra were recorded ona Micromass Q Tof 2 (Waters), and data were analyzed withMassLynx 4.0 (Waters) software. A549, B16-F10, and 4T1 cell lineswere obtained from ATCC and used within 6 months of resuscitationof frozen stock. MTS reagent was supplied by Promega. EasySepMouse CD90.2 positive selection kit was purchased form StemcellTechnologies. FACS antibodies for CD4, CD8, CD44, and CD62Land isotypes were purchased from Biolegend, and intracellular Foxp3staining kit was purchased from eBioscience. Pierce BCA protein assaykit and Vybrant Lipid Raft labeling kit and protease and phosphataseinhibitors were purchased from Thermo Fisher Scientific.All-Atomistic Computational Simulations. As the first step, we

performed geometry optimization of the molecular subunits using QMmethods to obtain the lowest-energy conformation of the molecularsubunit. Following the QM optimization step, partial charges on eachatom of the drug molecules were computed using Gaussian 09 withB3LYP exchange−correlation functional and 6311G basis set, andelectrostatic potentials were fitted with CHELPG scheme. Next wedeveloped the force field parameters of the molecular subunit. Bondand angle potential parameters were taken from CHARMM force field.Most of the dihedral potentials were taken from CHARMM force field,and those that were not present in the CHARMM force field werecalculated and parametrized. In terms of nonbonded potentials,Lennard-Jones (LJ) parameters were adapted from CHARMM forcefield, while for coulomb potential, partial charges on each atomobtained from QM calculations were used. Following the force fielddevelopment step, the subunit was energy optimized using thedeveloped force field using steepest descent algorithm. We nextperformed a short MD simulation of the molecular subunit in vacuumusing the developed force field, and the output structure obtained wasmatched with the structure obtained after QM optimization of thedrug molecule. If the structures obtained did not match, we loopedback to the force field development circle, and different parameterswere tweaked. This iteration continued until the structures obtainedafter QM optimization and MD in vacuum matched. All the MDsimulations were performed using GROMACS-4.6.1 package. QMgeometry-optimized structures of selumetinib, STB, PI103, and SPIwere considered as the starting structure for MD simulations. All theother bonded and nonbonded parameters were taken from CHARMMforce field. All covalent bonds in SOPC and the drugs wereconstrained with LINCS algorithm. SETTLE method was used toconstrain the covalent bonds in water. Time step of 2 fs was used forall the simulations. Isothermal and isobaric ensemble (NPT) andperiodic boundary condition (PBC) were used. Temperature was keptconstant at 300 K for all the systems using Nose−Hoover thermostatwith a coupling constant of 0.5 ps. Pressure was kept constant at 1 atmusing Parrinello−Rahman barostat with coupling constant of 10 ps.Semi-isotropic pressure coupling was used. Neighbor searching wasdone with Verlet algorithm as implemented in Gromacs 4.6.1. Forlong-range electrostatic interactions, we used a particle mesh Ewald(PME) scheme with real space cut off of 1.4 nm. For van der Waalsinteractions, 1.5 nm cut off was applied with a switching function at 1.4nm. Selu, PI103 or STB, SPI molecules were inserted inside the lipidbilayer structure and equilibrated until 10 ns for all the systems.Models were constructed such that at the high drug concentration(20%), the system contained 256 molecules of SOPC and 64molecules of geometry-optimized structure of selumetinib or PI103.The trajectories were saved for further analysis.Synthesis of Active Molecular Subunits STB and SPI.

Selumetinib (32 mg, 0.072 mmol) was dissolved in 3 mL anhydrousDCM followed by addition of cholesterol-succinic acid (0.216 mmol,105 mg), EDC (0.216 mmol, 41.4 mg), and DMAP (0.216, 26 mg).

The reaction mixture was stirred at rt for 12 h under argon. Uponcompletion of reaction as monitored by TLC, the reaction mixture wasdiluted with 10 mL DCM and washed with dilute HCl and water. Theorganic layers were separated, combined, and dried over anhydroussodium sulfate. The solvent was evaporated under vacuum, and thecrude product was purified by using column chromatography, elutingwith methanol:methylene chloride gradient, to give the supratinib as alight-yellow solid (54 mg, 82%). 1H NMR (CDCl3, 400 MHz): δ 9.75(s, 1H), 8.03 (m, 1H), 7.9 (s, 1H), 7.54 (d, 1H), 7.1 (m, 4H), 5.36 (s,1H), 4.67 (m, 1H), 4.3 (m, 1H), 4.12 (m, 1H), 3.95 (m, 2H), 3.85 (s,1H), 2.65 (m, 3H), 2.31 (m, 2H), 2.06 (m, 2H), 1.86 (m, 3H), 1.72 (s,3H), 1.29 (m, 33H), 0.74 (m, 3H). HRMS calculated for[C48H63BrClFN4O6 + H]+: 925.37, found: 925.39. Cholesterylhemisuccinate was similarly conjugated to PI103 using EDC andDMAP coupling reaction to form molecular subunit of SPI.29

Synthesis and Characterization of Nanoscale Supramolec-ular Therapeutics (STB and SPI). For supramolecule synthesis, 6.5mg (50 mol %) of L-α-phosphatidylcholine, 3.1 mg supratinib, or 2.7mg molecular subunit of SPI (20 mol %) and 14.1 mg (30 mol %) of1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[amino-(polythylene glycol)2000] (DSPE-PEG) were dissolved in 1.0 mLDCM. Solvent was evaporated into a thin and uniform lipid-drug filmusing a rotary evaporator. The lipid-drug film was then hydrated with1.0 mL H2O for 2 h at 70 °C. It was extruded at 70 °C using 400 nmand then 200 nm membrane with 100 μL sample volume to obtainsub-200 nm particles. The nanoscale supramolecular therapeutics werefurther passed to Sephadex G-25 column to remove free molecularsubunits. Drugs incorporated in the STB or SPI were determined usingUV−vis spectroscopy. The incorporation efficiency was determined asthe percentage of drug recovered from the nanoscale supramoleculartherapeutic fractions compared to the initial loading amount. Themean particle size of the nanoscale supramolecular therapeutics wasmeasured by dynamic light scattering (DLS) method using ZetasizerNano ZS90 (Malvern, U.K.). The ζ potential was measured using aZetasizer ZS90 with the nanoscale supramolecular therapeutics dilutedin water for measurement according to the manufacturer’s manual.The physical stability of nanoscale supramolecular therapeutics wasevaluated by measuring changes in mean particle size and ζ potentialduring storage condition at 4 °C. For serum stability study, STB andSPI (100 μL) were incubated with 100 μL of human serum plasma.The particle size and ζ potential were measured as a function of timeusing Malvern Zetasizer.

Cryo-Transmission Electron Microscopy. The sample waspreserved in vitrified ice supported by holey carbon films on 400mesh copper grids. The sample was prepared by applying 3 μL ofsample suspension to a cleaned grid, blotting away with filter paperand immediately proceeding with vitrification in liquid ethane. Gridswere stored under liquid nitrogen until transferred to the electronmicroscope for imaging. Electron microscopy was performed using anFEI Tecnai Cryo-Bio 200 kV FEG TEM, operating at 120 keVequipped with 2 Gatan Sirius CCD cameras one 2K × 2K and one 4K× 4K pixel. Vitreous ice grids were transferred into the electronmicroscope using a cryostage that maintains the grids at a temperaturebelow −170 °C. Images of the grid were acquired at multiple scales toassess the overall distribution of the specimen. After identifyingpotentially suitable target areas for imaging at lower magnification,high-magnification images were acquired at nominal magnification of52,000× (0.21 nm/pixel) and 21,000× (0.50 nm/pixel). Images wereacquired at a nominal underfocus of −5 μm (21,000×) and −4 μm(52,000×) at electron doses of ∼10−15 e/A°2.

Release Kinetics Studies. Nanoscale supramolecular therapeutics(1 mg drug/mL, 5 mL) were suspended in PBS buffer (pH 7.4) oracidic conditions (pH 5.5) or A549 cell lysate and sealed in a dialysistube (MWCO = 3500 Da, Spectrum Lab). The dialysis tube wassuspended in 1L PBS pH 7.4 pH with gentle stirring to simulate theinfinite sink tank condition. A 100 μL portion of the aliquot wascollected from the incubation medium at predetermined time intervalsand replaced by equal volume of PBS buffer, and the released drug wasquantified by UV−vis spectrophotometer and plotted as cumulativedrug release.

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9238

Page 13: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

Release Kinetics from Splenocytes Lysate. Spleen was isolatedfrom Balb/C (female, 4−6 weeks old) mice, and lysate was preparedusing NP40 lysis buffer in the presence of Halt protease andphosphatase inhibitor. STB and SPI (equivalent to 3 mg/mL of drug, 5mL) were suspended in splenocytes lysate and sealed in a dialysis tube(MWCO = 3500 Da, Spectrum Lab). The dialysis tube was suspendedin 1 L PBS 7.4 pH with gentle stirring to simulate the infinite sink tankcondition. A 100 μL portion of the aliquot was collected from theincubation medium at predetermined time intervals and replaced byequal volume of PBS buffer, and the released drug was quantified byUV−vis spectrophotometer and plotted as cumulative drug release.In Vitro Uptake of FITC-Cholesterol Supramolecular Nano-

particles. B16-F10 melanoma cells were labeled and cross-linkedusing Vybrant 555 lipid raft labeling kit according to manufacturer’sprotocols. The labeled cells were then seeded on 8 chamber tissueculture-treated slides at 50,000 cells/well density and were allowed toattain 70% subconfluency. The cells were treated with FITC-cholesterol tagged supramolecular nanoparticles for 30 min, 6 h, and12 h. After treatment, the cells were washed with ice-cold PBS toremove supramolecular nanoparticles that have not been internalized.The cells were later fixed using 4% paraformaldehyde and weremounted using ProLong Gold Antifade reagent and were imaged usingOlympus confocal microscope.In Vitro Assays. A549 human lung cancer cells and B16-F10

mouse melanoma cells were cultured in DMEM, supplemented with10% FBS and 1% of antibiotic-antimycotic 100× solution (Invitrogen,15240-062). For cell viability studies, cells (4 × 103) were seeded into96-well flat-bottomed plates and incubated with free drug or nanoscalesupramolecular therapeutics (normalized to equivalent amounts of freedrug) for desired time periods. Cell viability was quantified using theCellTiter 96 Aqueous One Solution assay (Promega).For Western blot, 2 × 105 cells were seeded in each well of a 6-well

plate and incubated with free drug or SNPs (normalized to equivalentamounts of free drug) for desired time periods. After the desired timepoints, the cells were lysed, and protein lysates were electrophoresedand transferred to membranes, which were incubated with antibodiesagainst phosphorylated forms of proteins (Phospho-p44/42 MAPK(Erk1/2) (1:1000 dilution), p44/42 MAPK (Erk1/2) (1:1000dilution), and β-actin (1:1000 dilution) antibodies) and probed withhorseradish peroxidase-conjugated secondary antibody. Expression wasnormalized to total expression of the specific protein or β-actin. Tostudy drug internalization, 1 × 106 cancer cells were incubated withselumetinib or an equivalent amount of nanoscale supramoleculartherapeutics for 3 and 18 h. The cells were then lysed, and drugconcentrations of the drug were quantified using spectroscopy.In Vitro Cell Viability Assay of T Cells. T cells were isolated from

the spleen of Balb/C mice (female, 4−6 weeks old) using mouseCD90.2 T cell isolation kit, according to manufacturer’s protocol, wereseeded on a 96-well plate at the density of 10,000 cells/well, weretreated with different concentrations of PI103, SPI, selumetinib, andSTB, and were allowed to incubate for 48 h. The cells were thencentrifuged at 2000 rpm and washed with PBS. The cells were exposedto CellTiter Aqueous One solution and were incubated at 37 °C. Afterdesired incubation in a 5% CO2 atmosphere at 37 °C, the absorbancein each well was recorded at 490 nm using an Epoch plate reader(Biotek instruments, VT). The absorbance reflects the number ofsurviving cells. Blanks were subtracted from all data, and resultsanalyzed using GraphPad Prism software (GraphPad, San Diego, CA).Each experiment was independently repeated thrice, and data shownare mean ± SEM of n = 3.In Vitro FACS To Determine PDL-1 Expression on B16-F10

Melanoma Cells. Five ×105 B16-F10 melanoma cells were seeded ineach well of a 6-well plate. Then, selumetinib or STB (2.5 μM) wasadded in DMEM complete media (10% FBS). After 48 h of incubationin 5% CO2 atmosphere at 37 °C, cells were washed once with ice-coldPBS and collected for flow cytometric analysis. The cell surfaceexpression of PDL-1 was evaluated by surface staining using themanufacturer’s protocol (PE-Anti-PDL-1, Biolegend). The PDL-1positive cells were quantified using BD Accuri C6 cytometer, and datawere analyzed using FlowJo.

In Vitro Flow Cytometric Surface Marker Analysis of SpleenDerived T Cells. Spleens were harvested from C57BL/6 mice (male,4−6 weeks old), and a single cell suspension was prepared from thespleen samples. The single cell suspension was passed through 40 μmfilter to remove debris and fat. Washing was carried out using RPMI-1640 media by centrifuging at 2000 rpm for 5 min (2 times). Thepellet was resuspended in 1 mL of recommended media for CD90.2+T cell isolation. The isolation was carried out using Easy Sep MouseCD90.2 positive selection kit II according to the manufacturer’sprotocol (Stem Cell Technologies). The isolated CD90.2 positive Tcells were cultured in Immunocult ex vivo expansion medium andtreated with free drug or equivalent amount of supramoleculartherapeutics (0.5 μm). After 36 h of incubation, the cells wereactivated for 12 h using CD3/CD28 Dynabeads according to themanufacturer’s protocol. The cells were collected from both sets ofexperiments and quantified using a BD J LSR II without removingDynabeads. The cell surface expressions of CD4 vs CD8 and CD44 vsCD62L were evaluated by surface staining in order to identify thepercentage of immune effector cells (CD4+, CD8+ T cells) andactivated T cells (CD44HI62LLo).

In Vivo Biodistribution Study. For biodistribution studies, DiRdye-labeled STB-nanoscale supramolecular therapeutics were synthe-sized by lipid-film hydration method. Briefly, 4.3 mg (50 mol %) of L-α-phosphatidylcholine, 1.9 mg (19 mol %) of STB, 0.12 mg (1 mol %)of DiR dye, and 9.3 mg (30 mol %) of 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[amino(polythylene glycol)2000] (DSPE-PEG) were dissolved in 1.0 mL DCM. Solvent was evaporated intoa thin and uniform lipid-drug film using a rotary evaporator. The lipid-drug film was then hydrated with 1.0 mL H2O for 2 h at 70 °C. B16-F10 melanoma cells (5 × 105) were inoculated subcutaneously into theflank of each male C57B/L6 mice (4−6 weeks old, weight ∼20 g).After the tumor reached ∼500 mm3 volume, dye-labeled nanoparticleswere injected through the tail vein of the tumor bearing mice. Theimaging was performed at 1, 3, 6, 24, and 48 h post-injection usingIVIS filter set (excitation 710 nm and emission 760 nm). After theimaging, mice were sacrificed, and major organs were collected andimaged. All the images were captured using Maestro (CRI) smallanimal in vivo fluorescence imaging system. For the purposes of thisstudy, the exposure times for acquiring the data were kept the same. Inaddition, the stage settings (which influences the distance from thecamera) and the fiber optic adjustable illuminator arms were also keptat the same settings for the duration of the study. After the studyperiod, the tumors and other organs were isolated for ex-vivofluorescence imaging. Fluorescence signals were normalized andquantified using Meastro Software. The Maestro Software was usedto conduct spectral unmixing, i.e., separate individual fluorophoresfrom each other and background autofluorescence. This feature wasused to produce the spectrally unmixed images. Using a controlmouse, we assigned a spectrum for autofluorescence/background. Aphantom containing the dye was used to assign a spectrum for signal.Using the Maestro Software, we established a “spectral library” whichindicates a spectrum for in vivo signal from the dye, minusautofluorescence. This spectral library, once established, was appliedto all raw data to produce the “spectrally unmixed images”.Quantification of fluorescence intensity in the organs was performedby calculating the ratio of relative fluorescent intensity in each organwith the sum of the total intensity accumulated for all organs.

Western Blotting to Determine Liver and Kidney Toxicity.Liver and kidney from all treatment groups stored in −80 °C werehomogenized to form a single cell suspension. Protein was extractedusing NP40 lysis buffer in the presence of Halt protease andphosphatase inhibitor. Protein quantification was performed using theBCA assay, and equal amounts of protein lysates were electrophoresedon a polyacrylamide gel, then transferred to polyvinylidene difluoridemembrane (Bio-Rad), and blocked in 5% milk solution. Themembranes were incubated with cleaved-caspase-3, caspase-3, and β-actin primary antibodies (1:1000 dilution) (all antibodies werepurchased from Cell Signaling Technology) overnight at 4 °C,followed by horseradish peroxidase-conjugated secondary antibody

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9239

Page 14: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

incubation for 1 h. Detection was carried out using G-box fromSyngene.Combination Efficacy Studies. B16-F10 melanoma cells (5 ×

105) were implanted subcutaneously in the right flank of 4-week-oldC57BL/6 mice (male). When the tumor volume reached ∼50 mm3,animals were randomized into different treatment groups. The animalswere administered via tail vain with following treatments: (1) vehicle,(2) selumetinib (5 mg/kg), and (3) STB (at dose equivalent to 5 mg/kg of the selumetinib); (4) PDL-1 antibody (2.5 mg/kg), (5)selumetinib (5 mg/kg) + PDL-1 antibody (2.5 mg/kg), and (6) STB(at dose equivalent to 5 mg/kg of the selumetinib) + PDL-1 antibody(2.5 mg/kg). Separately, 4T1 breast cancer cells (1 × 106) wereimplanted subcutaneously in the flanks of 4-week-old female BALB/cmice. The drug therapy was started on day 10. Each animal wasinjected intravenously with three doses of either vehicle (for controlgroup), 2.5 mg/kg of free PI103, 2.5 mg/kg of SPI, 5 mg/kg αPD-1,PI103 + αPD-1, or SPI + αPD-1 on each alternate days. Foroptimizing the dosing schedules, 4T1 tumor bearing female BALB/cmice were injected with either 5 mg/kg αPD-1 on alternate days ofadministration of SPI (2.5 mg/kg) or 5 mg/kg αPD-1 andadministered after the cycle of SPI treatment (2.5 mg/kg) wascomplete. The tumors were measured every alternate days using aVernier caliper, and tumor volume (Vt) was calculated using theformula, L × B2/2, where L is the longest, and W is the shortestdimensions. Tissues were harvested for FACS analysis and furtherstudies. All animal procedures were approved by the HarvardInstitutional Use and Care of Animals Committee.Tumor Histocytochemistry. Immunohistochemical analysis was

performed on formalin-fixed and paraffin-embedded 5 μm tumorsections. Briefly, the tissue sections were heated to 60 °C andrehydrated in xylene and graded alcohols. Antigen retrieval was carriedout in 0.1 M citrate buffer (pH 6.0) for 10 min in a pressure cookerand allowed to cool to room temperature followed by rinsing in PBSand PBST (0.1% Tween). The sections were then incubated in 3%hydrogen peroxide for 5 min to ablate endogenous tissue peroxidaseactivity. Each incubation step was carried out at room temperature andwas followed by three 5 min washes in PBST. The tissue sections werethen incubated with primary antibodies pERK (Cell Signaling) inantibody dilution buffer (Envision) for 45 min at room temperaturefollowed by incubations with biotinylated secondary antibody for 30min. The slides were then stained using DAKO EnVision+ Systemhorseradish peroxidase detection kit and counterstained withhematoxylin. Images were obtained using Nikon 90i microscope.Tumors were cryosectioned, tissues were stained with CD45 antibody(for tumor infiltrating leukocytes analysis), and nuclei were labeledwith DAPI. For studying apoptosis, formalin-fixed tumor sections werestained with standard TMR red fluorescent terminal deoxynucleotidyltranferase-mediated dUTP nick end labeling (TUNEL) kit followingthe manufacturer’s protocol (In Situ Cell Death Detection Kit, TMR-Red, Roche). Images were captured using Nikon C1 confocalmicroscope.FACS Analysis. The harvested tumor tissues were thoroughly

minced and suspended in separate 4 mL type-I collagenase solution(175 U/mg). A single cell suspension was prepared for each tumorsample and incubated for 1 h at 37 °C and 5% CO2 and finally passedthrough 40 μm filter. Washing was carried out using RPMI-1640media by centrifuging at 2000 rpm for 5 min (2 times). The pellet wasresuspended in 1 mL of recommended media for CD90.2+ T cellisolation. The isolation was carried out using EasySep Mouse CD90.2positive selection kit II according to the manufacturer’s protocol (StemCell Technologies). The phenotype of the isolated CD90.2+ T cellsfrom the tumor single cell suspension was quantified using a BD J LSRII cytometer. The cell surface expression of CD4, CD8, CD44,CD62L, and intracellular Foxp3 was evaluated by respective surfaceand intracellular staining in order to identify the percentage of immuneeffector cells (CD4+, CD8+ T cells and TRegs) and activated T cells(CD44Hi62LLo) in the tumor bearing mice.Western Blot Analysis of in Vivo Tumor Samples. Tumor

samples were homogenized into single cell suspensions in serum-freeRPMI-1640 media containing 175 U/mL of collagenase type I. The

cell suspensions were then lysed in NP40 cell lysis buffer containingprotease and phosphatase inhibitors. The cellular debris was removedby centrifugation at 15000 rpm, and the supernatant was taken for theexperiment. Protein estimation was done by BCA protein assay kit. 20μg protein was loaded in each well and probed for phospho-ERK1/2,total ERK 1/2, and β-actin (for STB animal study) and phospho Akt,total Akt, phospho mTOR, total mTOR, and β-actin (for SPI animalstudy). Detection was done using G-box from Syngene, anddensitometric quantification was done by image J software.

Statistics. The statistical analysis was done using two-tailedStudent’s t test or one-way ANOVA followed by Newman Keulspost-hoc test, with P < 0.05 as the threshold for significance.

ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acsnano.6b01600.

Supporting results and figures (PDF)Movie 1: All-atomistic simulation of Selumetinibinteracting with colipids in a bilayer (AVI)Movie 2: All-atomistic simulation of Supratinib interact-ing with colipids in a bilayer (AVI)Movie 3: All-atomistic simulation of PI103 interactingwith colipids in a bilayer (AVI)Movie 4: All-atomistic simulation of SPI molecularsubunit interacting with colipids in a bilayer (AVI)

AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected].

Author Contributions∇These authors contributed equally.

NotesThe authors declare the following competing financialinterest(s): S.S. is a cofounder, owns equity, and is a memberof the board at Invictus Oncology.

ACKNOWLEDGMENTSThe authors would like to thank the Dana Farber CancerInstitute Flow Cytometry Core Facility for their expertise,consulting, and assistance with flow cytometry experiments.The cryo-TEM analysis was performed in part at the Center forNanoscale Systems (CNS), a member of the NationalNanotechnology Coordinated Infrastructure Network(NNCI), which is supported by the National ScienceFoundation under NSF award no. 1541959. CNS is part ofHarvard University. The authors would like to thank Dr. SudipRoy (National Chemical Laboratory, Pune, India) for providingthe support in computational analysis. This work was supportedby a DoD BCRP Collaborative Innovator Grant (W81XWH-09-1-0700), a NIH RO1 (1R01CA135242-01A2), a DoDBreakthrough Award (BC132168), an American Lung Associ-ation Innovation Award (LCD-259932-N), and an Indo-USJoint Center Grant from IUSSTF to S.S.

REFERENCES(1) Sharma, P.; Allison, J. P. The Future of Immune CheckpointTherapy. Science 2015, 348, 56−61.(2) Topalian, S. L.; Drake, C. G.; Pardoll, D. M. Immune CheckpointBlockade: A Common Denominator Approach to Cancer Therapy.Cancer Cell 2015, 27, 450−461.

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9240

Page 15: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

(3) Schreiber, R. D.; Old, L. J.; Smyth, M. J. Cancer Immunoediting:Integrating Immunity’s Roles in Cancer Suppression and Promotion.Science 2011, 331, 1565−1570.(4) Kroemer, G.; Senovilla, L.; Galluzzi, L.; Andre, F.; Zitvogel, L.Natural and Therapy-Induced Immunosurveillance in Breast Cancer.Nat. Med. 2015, 21, 1128−1138.(5) Wherry, E. J.; Kurachi, M. Molecular and Cellular Insights into TCell Exhaustion. Nat. Rev. Immunol. 2015, 15, 486−499.(6) Postow, M. A.; Callahan, M. K.; Wolchok, J. D. ImmuneCheckpoint Blockade in Cancer Therapy. J. Clin. Oncol. 2015, 33,1974−1982.(7) Kyi, C.; Postow, M. A. Checkpoint Blocking Antibodies inCancer Immunotherapy. FEBS Lett. 2014, 588, 368−376.(8) Kelderman, S.; Schumacher, T. N.; Haanen, J. B. Acquired andIntrinsic Resistance in Cancer Immunotherapy. Mol. Oncol. 2014, 8,1132−1139.(9) Mellman, I.; Coukos, G.; Dranoff, G. Cancer ImmunotherapyComes of Age. Nature 2011, 480, 480−489.(10) Mahoney, K. M.; Rennert, P. D.; Freeman, G. J. CombinationCancer Immunotherapy and New Immunomodulatory Targets. Nat.Rev. Drug Discovery 2015, 14, 561−584.(11) Melero, I.; Berman, D. M.; Aznar, M. A.; Korman, A. J.; PerezGracia, J. L.; Haanen, J. Evolving Synergistic Combinations ofTargeted Immunotherapies to Combat Cancer. Nat. Rev. Cancer2015, 15, 457−472.(12) Aris, M.; Barrio, M. M. Combining Immunotherapy withOncogene-Targeted Therapy: A New Road for Melanoma Treatment.Front. Immunol. 2015, 6, 46.(13) Higuchi, T.; Flies, D. B.; Marjon, N. A.; Mantia-Smaldone, G.;Ronner, L.; Gimotty, P. A.; Adams, S. F. CTLA-4 Blockade SynergizesTherapeutically with PARP Inhibition in BRCA1-Deficient OvarianCancer. Cancer Immunol. Res. 2015, 3, 1257−1268.(14) Sharabi, A. B.; Lim, M.; DeWeese, T. L.; Drake, C. G. Radiationand Checkpoint Blockade Immunotherapy: Radiosensitisation andPotential Mechanisms of Synergy. Lancet Oncol. 2015, 16, e498−509.(15) Sharma, P.; Allison, J. P. Immune Checkpoint Targeting inCancer Therapy: Toward Combination Strategies with CurativePotential. Cell 2015, 161, 205−214.(16) Vanneman, M.; Dranoff, G. Combining Immunotherapy andTargeted Therapies in Cancer Treatment. Nat. Rev. Cancer 2012, 12,237−251.(17) Boni, A.; Cogdill, A. P.; Dang, P.; Udayakumar, D.; Njauw, C.N.; Sloss, C. M.; Ferrone, C. R.; Flaherty, K. T.; Lawrence, D. P.;Fisher, D. E.; Tsao, H.; Wargo, J. A. Selective BRAFV600E InhibitionEnhances T-Cell Recognition of Melanoma without AffectingLymphocyte Function. Cancer Res. 2010, 70, 5213−5219.(18) Hahnel, P. S.; Thaler, S.; Antunes, E.; Huber, C.; Theobald, M.;Schuler, M. Targeting AKT Signaling Sensitizes Cancer to CellularImmunotherapy. Cancer Res. 2008, 68, 3899−3906.(19) Noh, K. H.; Kang, T. H.; Kim, J. H.; Pai, S. I.; Lin, K. Y.; Hung,C. F.; Wu, T. C.; Kim, T. W. Activation of AKT as a Mechanism forTumor Immune Evasion. Mol. Ther. 2009, 17, 439−447.(20) Jiang, X.; Zhou, J.; Giobbie-Hurder, A.; Wargo, J.; Hodi, F. S.The Activation of MAPK in Melanoma Cells Resistant to BRAFInhibition Promotes PD-L1 Expression That Is Reversible by MEKand PI3K Inhibition. Clin. Cancer Res. 2013, 19, 598−609.(21) Atefi, M.; Avramis, E.; Lassen, A.; Wong, D. J.; Robert, L.;Foulad, D.; Cerniglia, M.; Titz, B.; Chodon, T.; Graeber, T. G.;Comin-Anduix, B.; Ribas, A. Effects of MAPK and PI3K Pathways onPD-L1 Expression in Melanoma. Clin. Cancer Res. 2014, 20, 3446−3457.(22) Kim, T.; Amaria, R. N.; Spencer, C.; Reuben, A.; Cooper, Z. A.;Wargo, J. A. Combining Targeted Therapy and Immune CheckpointInhibitors in the Treatment of Metastatic Melanoma. Cancer Biol. Med.2014, 11, 237−246.(23) Francisco, L. M.; Sage, P. T.; Sharpe, A. H. The PD-1 Pathwayin Tolerance and Autoimmunity. Immunol. Rev. 2010, 236, 219−242.(24) Vella, L. J.; Pasam, A.; Dimopoulos, N.; Andrews, M.; Knights,A.; Puaux, A. L.; Louahed, J.; Chen, W.; Woods, K.; Cebon, J. S. MEk

Inhibition, Alone or in Combination with BRAF Inhibition, AffectsMultiple Functions of Isolated Normal Human Lymphocytes andDendritic Cells. Cancer Immunol. Res. 2014, 2, 351−360.(25) Sengupta, S.; Kulkarni, A. Design Principles for Clinical Efficacyof Cancer Nanomedicine: A Look into the Basics. ACS Nano 2013, 7,2878−2882.(26) Maeda, H.; Wu, J.; Sawa, T.; Matsumura, Y.; Hori, K. TumorVascular Permeability and the EPR Effect in MacromolecularTherapeutics: A Review. J. Controlled Release 2000, 65, 271−284.(27) Goldberg, M. S. Immunoengineering: How NanotechnologyCan Enhance Cancer Immunotherapy. Cell 2015, 161, 201−204.(28) Basu, S.; Harfouche, R.; Soni, S.; Chimote, G.; Mashelkar, R. A.;Sengupta, S. Nanoparticle-Mediated Targeting of MAPK SignalingPredisposes Tumor to Chemotherapy. Proc. Natl. Acad. Sci. U. S. A.2009, 106, 7957−7961.(29) Kulkarni, A. A.; Roy, B.; Rao, P. S.; Wyant, G. A.; Mahmoud, A.;Ramachandran, M.; Sengupta, P.; Goldman, A.; Kotamraju, V. R.;Basu, S.; Mashelkar, R. A.; Ruoslahti, E.; Dinulescu, D. M.; Sengupta,S. Supramolecular Nanoparticles That Target Phosphoinositide-3-Kinase Overcome Insulin Resistance and Exert Pronounced AntitumorEfficacy. Cancer Res. 2013, 73, 6987−6997.(30) Pandey, A.; Kulkarni, A.; Roy, B.; Goldman, A.; Sarangi, S.;Sengupta, P.; Phipps, C.; Kopparam, J.; Oh, M.; Basu, S.; Kohandel,M.; Sengupta, S. Sequential Application of a Cytotoxic Nanoparticleand a PI3K Inhibitor Enhances Antitumor Efficacy. Cancer Res. 2014,74, 675−685.(31) Lehn, J.-M. Supramolecular ChemistryScope and PerspectivesMolecules, Supermolecules, and Molecular Devices (Nobel Lecture).Angew. Chem., Int. Ed. Engl. 1988, 27, 89−112.(32) Kulkarni, A.; Pandey, P.; Rao, P.; Mahmoud, A.; Goldman, A.;Sabbisetti, V.; Parcha, S.; Natarajan, S. K.; Chandrasekar, V.;Dinulescu, D.; Roy, S.; Sengupta, S. Algorithm for DesigningNanoscale Supramolecular Therapeutics with Increased AnticancerEfficacy. ACS Nano 2016, 10, 8154−8168.(33) Falck, E.; Patra, M.; Karttunen, M.; Hyvonen, M. T.;Vattulainen, I. Lessons of Slicing Membranes: Interplay of Packing,Free Area, and Lateral Diffusion in Phospholipid/Cholesterol Bilayers.Biophys. J. 2004, 87, 1076−1091.(34) Klauda, J. B.; Venable, R. M.; Freites, J. A.; O’Connor, J. W.;Tobias, D. J.; Mondragon-Ramirez, C.; Vorobyov, I.; MacKerell, A. D.,Jr.; Pastor, R. W. Update of the Charmm All-Atom Additive ForceField for Lipids: Validation on Six Lipid Types. J. Phys. Chem. B 2010,114, 7830−7843.(35) Jambeck, J. P.; Lyubartsev, A. P. Derivation and SystematicValidation of a Refined All-Atom Force Field for PhosphatidylcholineLipids. J. Phys. Chem. B 2012, 116, 3164−3179.(36) Jambeck, J. P.; Lyubartsev, A. P. An Extension and FurtherValidation of an All-Atomistic Force Field for Biological Membranes. J.Chem. Theory Comput. 2012, 8, 2938−2948.(37) William, L.; Jorgensen, J. C.; Jeffry, D.; Madura; Roger, W.;Impey; Michael, L.; Klein. Comparison of Simple Potential Functionsfor Simulating Liquid Water. J. Chem. Phys. 1983, 79, 926.(38) Flaherty, K. T.; Robert, C.; Hersey, P.; Nathan, P.; Garbe, C.;Milhem, M.; Demidov, L. V.; Hassel, J. C.; Rutkowski, P.; Mohr, P.;Dummer, R.; Trefzer, U.; Larkin, J. M.; Utikal, J.; Dreno, B.; Nyakas,M.; Middleton, M. R.; Becker, J. C.; Casey, M.; Sherman, L. J.; et al.Group, M. S.: Improved Survival with MEK Inhibition in BRAF-Mutated Melanoma. N. Engl. J. Med. 2012, 367, 107−114.(39) Chen, L.; Han, X. Anti-PD-1/PD-L1 Therapy of HumanCancer: Past, Present, and Future. J. Clin. Invest. 2015, 125, 3384−3391.(40) Callahan, M. K.; Masters, G.; Pratilas, C. A.; Ariyan, C.; Katz, J.;Kitano, S.; Russell, V.; Gordon, R. A.; Vyas, S.; Yuan, J.; Gupta, A.;Wigginton, J. M.; Rosen, N.; Merghoub, T.; Jure-Kunkel, M.;Wolchok, J. D. Paradoxical Activation of T Cells Via AugmentedERK Signaling Mediated by a RAF Inhibitor. Cancer Immunol. Res.2014, 2, 70−79.(41) Gonzalez-Angulo, A. M.; Ferrer-Lozano, J.; Stemke-Hale, K.;Sahin, A.; Liu, S.; Barrera, J. A.; Burgues, O.; Lluch, A. M.; Chen, H.;

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9241

Page 16: Combining Immune Checkpoint Inhibitors and Article Kinase ...kulkarnigroup.com/wp-content/uploads/2017/07/Publication-15.pdf · The combination of supramolecular therapeutics and

Hortobagyi, G. N.; Mills, G. B.; Meric-Bernstam, F. PI3K PathwayMutations and PTEN Levels in Primary and Metastatic Breast Cancer.Mol. Cancer Ther. 2011, 10, 1093−1101.(42) Pages̀ F, R. M.; Rottapel, R.; Truneh, A.; Nunes, J.; Imbert, J.;Olive, D. Binding of Phosphatidylinositol-3-OH Kinase to CD28 IsRequired for T-Cell Signalling. Nature 1994, 370, 327−329.(43) Parry, R. V.; Chemnitz, J. M.; Frauwirth, K. A.; Lanfranco, A. R.;Braunstein, I.; Kobayashi, S. V.; Linsley, P. S.; Thompson, C. B.; Riley,J. L. CTLA-4 and PD-1 Receptors Inhibit T-Cell Activation by DistinctMechanisms. Mol. Cell. Biol. 2005, 25, 9543−9553.(44) Raynaud, F. I.; Eccles, S.; Clarke, P. A.; Hayes, A.; Nutley, B.;Alix, S.; Henley, A.; Di-Stefano, F.; Ahmad, Z.; Guillard, S.; Bjerke, L.M.; Kelland, L.; Valenti, M.; Patterson, L.; Gowan, S.; de HavenBrandon, A.; Hayakawa, M.; Kaizawa, H.; Koizumi, T.; Ohishi, T.;et al. Pharmacologic Characterization of a Potent Inhibitor of Class IPhosphatidylinositide 3-Kinases. Cancer Res. 2007, 67, 5840−5850.(45) Mittendorf, E. A.; Philips, A. V.; Meric-Bernstam, F.; Qiao, N.;Wu, Y.; Harrington, S.; Su, X.; Wang, Y.; Gonzalez-Angulo, A. M.;Akcakanat, A.; Chawla, A.; Curran, M.; Hwu, P.; Sharma, P.; Litton, J.K.; Molldrem, J. J.; Alatrash, G. PD-L1 Expression in Triple-NegativeBreast Cancer. Cancer Immunol. Res. 2014, 2, 361−370.(46) Goldman, A.; Majumder, B.; Dhawan, A.; Ravi, S.; Goldman, D.;Kohandel, M.; Majumder, P. K.; Sengupta, S. Temporally SequencedAnticancer Drugs Overcome Adaptive Resistance by Targeting aVulnerable Chemotherapy-Induced Phenotypic Transition. Nat.Commun. 2015, 6, 6139.(47) Lee, M. J.; Ye, A. S.; Gardino, A. K.; Heijink, A. M.; Sorger, P.K.; MacBeath, G.; Yaffe, M. B. Sequential Application of AnticancerDrugs Enhances Cell Death by Rewiring Apoptotic SignalingNetworks. Cell 2012, 149, 780−794.(48) Jambeck, J. P.; Eriksson, E. S.; Laaksonen, A.; Lyubartsev, A. P.;Eriksson, L. A. Molecular Dynamics Studies of Liposomes as Carriersfor Photosensitizing Drugs: Development, Validation, and Simulationswith a Coarse-Grained Model. J. Chem. Theory Comput. 2014, 10, 5−13.(49) Sengupta, P.; Basu, S.; Soni, S.; Pandey, A.; Roy, B.; Oh, M. S.;Chin, K. T.; Paraskar, A. S.; Sarangi, S.; Connor, Y.; Sabbisetti, V. S.;Kopparam, J.; Kulkarni, A.; Muto, K.; Amarasiriwardena, C.;Jayawardene, I.; Lupoli, N.; Dinulescu, D. M.; Bonventre, J. V.;Mashelkar, R. A.; et al. Cholesterol-Tethered Platinum II-BasedSupramolecular Nanoparticle Increases Antitumor Efficacy andReduces Nephrotoxicity. Proc. Natl. Acad. Sci. U. S. A. 2012, 109,11294−11299.(50) Yamazaki, Y.; Ogawa, Y.; Afify, A. S.; Kageyama, Y.; Okada, T.;Okuno, H.; Yoshii, Y.; Nose, T. Difference between Cancer Cells andthe Corresponding Normal Tissue in View of StereoselectiveHydrolysis of Synthetic Esters. Biochim. Biophys. Acta, Gen. Subj.1995, 1243, 300−308.(51) Hu-Lieskovan, S.; Mok, S.; Homet Moreno, B.; Tsoi, J.; Robert,L.; Goedert, L.; Pinheiro, E. M.; Koya, R. C.; Graeber, T. G.; Comin-Anduix, B.; Ribas, A. Improved Antitumor Activity of Immunotherapywith BRAF and MEK Inhibitors in BRAF(V600E) Melanoma. Sci.Transl. Med. 2015, 7, 279ra41.(52) Su, F.; Viros, A.; Milagre, C.; Trunzer, K.; Bollag, G.; Spleiss, O.;Reis-Filho, J. S.; Kong, X.; Koya, R. C.; Flaherty, K. T.; Chapman, P.B.; Kim, M. J.; Hayward, R.; Martin, M.; Yang, H.; Wang, Q.; Hilton,H.; Hang, J. S.; Noe, J.; Lambros, M.; Geyer, F.; et al. RAS Mutationsin Cutaneous Squamous-Cell Carcinomas in Patients Treated withBRAF Inhibitors. N. Engl. J. Med. 2012, 366, 207−215.

ACS Nano Article

DOI: 10.1021/acsnano.6b01600ACS Nano 2016, 10, 9227−9242

9242