Abstract
Although photodynamic immunotherapy represents a promising therapeutic approach against malignant tumors, its efficacy is often hampered by the hypoxia and immunosuppressive conditions within the tumor microenvironment (TME) following photodynamic therapy (PDT). In this study, we report the design guidelines towards efficient Type-I semiconducting polymer photosensitizer and modify the best-performing polymer into a hypoxia-tolerant polymeric photosensitizer prodrug (HTPSNiclo) for cancer photo-immunotherapy. HTPSNiclo not only performs Type-I PDT process to partially overcome the limitation of hypoxic tumors in PDT by recycling oxygen but also specifically releases a Signal Transducer and Activator of Transcription-3 (STAT3) inhibitor (Niclosamide) in response to a cancer biomarker in the TME. Consequently, HTPSNiclo inhibits the phosphorylation of STAT3, and suppresses the expression of hypoxia-inducible factor-1α. The synergistic effect results in the enhanced activation of immune cells (including mature dendritic cells, cytotoxic T cells) and production of immunostimulatory cytokines compared to Type-I PDT alone. Thus, HTPSNiclo-mediated photodynamic immunotherapy enhances tumor inhibition rate from 75.53% to 91.23%, prolongs the 100% survival from 39 days to 60 days as compared to Type-I PDT alone. This study not only provides the generic approach towards design of polymer-based Type-I photosensitizers but also uncovers effective strategies to counteract the immunosuppressive TME for enhanced photo-immunotherapy in 4T1 tumor bearing female BALB/c mice.
Similar content being viewed by others
Introduction
Combination cancer immunotherapy has emerged as a prevalent strategy for cancer treatment, primarily because single treatment method is usually insufficient to address the challenges posed by tumor heterogeneity and the complex immunosuppressive tumor microenvironment (TME)1,2. In comparison to surgery, phototherapy3 in particularly, photodynamic therapy (PDT)4,5 offers a minimal invasive way to ablate tumor6 and triggers immunogenic cell death (ICD)7 to initiate tumor-specific immune response8,9,10. Thus, it has evolved as a prominent clinical therapeutic approach in multiple tumor types11. To amplify the antitumor response, photodynamic-immunotherapeutic nanomedicines12 composed of photosensitizers and immune checkpoint blockade (ICB) molecules13, tumor vaccine14, or cytokines15 have been developed. These nanomedicines can achieve controllable release of immunotherapeutic agents in the TME16, offering precise immune modulation17 with the potential to minimize immune-related adverse events (irAEs)18.
Currently, photodynamic-immunotherapeutic nano-agents19 face two main challenges20,21: tumor hypoxia and the immunosuppressive TME following phototherapy22,23. Because most photosensitizers are involved in Type-II photodynamic process, which rely on the molecular oxygen (O2) to generate singlet oxygen (1O2)24; their antitumor efficacy is hindered by hypoxic TME25. Furthermore, O2 consumption in Type-II PDT often worsens tumor hypoxia and upregulates the hypoxia-inducible factor-1α (HIF-1α), leading to the accumulation of immunosuppressive factors26. For example, elevated regulatory T cells (Treg) activity in immunosuppressive TME hinders immune function and promotes immune tolerance post-PDT27. In contrast, Type-I photosensitizers28 exhibit diminished oxygen-dependent generation of reactive oxygen species, mainly, superoxide anion (O2•–) and hydroxyl radical (·OH) via transfer electrons to O229. In the process, O2 could be recycled after O2•– transfers electrons to other molecules such as H2O, lipid, DNA, and protein30. As such, Type-I photo-immunotherapeutic nanoagents hold the potential to mitigate these two shortcomings, which unfortunately most photosensitizers used in this modality are inorganic materials31, and such nanoagents need to be better paired to achieve a synergetic effect.
Semiconducting polymer nanoparticles (SPNs)32,33,34 have been widely employed as efficient photosensitizers and drug carriers35,36 for cancer photo-immunotherapy due to their structural versatility and controllable photophysical properties37. To reverse the immunosuppressive TME, SPN-based photo-immunotherapeutic nano-agents have been developed by incorporating SPNs with ICB (PD-1/PD-L1, CTLA-4 checkpoint pathway) therapy38,39, immune-metabolic agents (adenosine deaminase, ADA)40 or immune activation agents (Toll-like receptor and STING agonist)41. Furthermore, to minimize irAEs induced by off-target immunotherapy, their immunomodulation components are often devised to be activated by either external (light or ultrasound)42,43 or internal stimuli (cancer-overexpressed enzymes)44. These activatable SPNs-based photo-immunotherapeutic nano-agents enhance tumor-specific immune responses45, leading to precision therapy with reduced off-target effects40. However, there are limited SPNs that can exert Type-I PDT process46,47, and they have been rarely exploited for Type-I photo-immunotherapy.
Herein, we report the development of a hypoxia-tolerant semiconducting polymeric photosensitizer Niclosamide prodrug (HTPSNiclo) for cancer photo-immunotherapy (Fig. 1). Through screening a library of semiconducting polymers (SPs), we identify an optimal Type-I SPs with high efficiency in generating O2•– and •OH (higher cytotoxicity compared with 1O2) as the Type-I PDT component in HTPSNiclo. Furthermore, to alleviate the immunosuppressive TME following PDT, a Signal Transducer and Activator of Transcription 3 (STAT3) inhibitor (Niclosamide) is conjugated to HTPS via Cathepsin B (CatB)-cleavable linker to afford HTPSNiclo. HTPSNiclo can efficiently produce O2•– in an oxygen-recyclable manner with reduced O2 consumption, which is beneficial for relieving hypoxic TME. Moreover, HTPSNiclo can specifically release Niclosamide in response to the overexpressed CatB in the TME to inhibit the phosphorylation of STAT3, which not only increase the secretion of immunostimulatory cytokines such as IL-12 and IFN-γ but also decrease the expression of HIF-1α48. Due to the synergistic effect of Type-I PDT and immune modulation, HTPSNiclo transforms the immunosuppressive TME into an active phenotype, resulting in tumor suppression, lung metastasis prevention and prolonged survival rate in 4T1 tumor-bearing mice.
a Chemical structure, self-assembly and mechanism of CatB-specific activation of HTPSNiclo. b Schematic illustration of single-mode Type-II PDT and HTPSNiclo-mediated photo-immunotherapy. (i) In single-mode Type-II PDT, NIR irradiation converts O2 into non-recyclable 1O2, inducing low cytotoxicity and a hypoxic TME that promotes immune tolerance and tumor metastasis. (ii) In Type-I photo-immunotherapy, HTPSNiclo generates highly cytotoxic O2•- and •OH in a O2 recyclable manner to reduce O2 consumption and HIF-1α expression and releases Niclosamide in situ to further downregulate the transcription of HIF-1α. The synergistic effect of Type-I PDT mediated enhanced ICD and immune modulation achieve robust immune activation, tumor suppression and metastasis prevention.
Results
Screening polymeric photosensitizers for Type-I PDT
A library of potential Type-I polymeric photosensitizers was constructed by copolymerization of a non-fullerene oligomer acceptor 2,2′-((2Z,2′Z)-((12,13-bis(2-octyldodecyl)-3,9-diundecyl-12,13-dihydro-[1,2,5]thiadiazolo[3,4-e]thieno[2″,3″:4’,5’]thieno[2′,3′:4,5]pyrrolo-[3,2-g]thieno[2′,3′:4,5]thieno[3,2-b]indole-2,10-diyl)bis(methanylylidene))bis(6-bromo-3-oxo-2,3-dihydro-1H-indene-2,1-diylidene))dimalononitrile (BTP-2Br-OD) with four types of donors including 4,7-Bis(5-(trimethylstannyl)thiophen-2-yl)benzo[c][1,2,5]thiadiazole (BDT), 2,5‐ bis(triMethylstannyl)thieno[3,2‐b]thiophene (TT), 4,8-bis((2-ethylhexyl)oxy)benzo[1,2b:4,5-b′]dithiophene-2,6-diyl)bis(trimethylstannane) (OT), 2,5-Bis(2-ethylhexyl)-3,6-bis(5-(trimethylstannyl)thiophen-2-yl)pyrrolo[3,4-c]pyrrole-1,4(2H,5H)-dione (DPP) via Stille coupling, denoted as P1-P4 (Fig. 2a). To screen the most efficient photosensitizer for generating Type-I PDT ROS, P1-P4 were assembled with the assistance of amphiphilic polymer poly(ethylene oxide)–poly(propylene oxide)–poly(ethylene oxide) (PEO–PPO–PEO, F127)49 into aqueous solution to form nanoparticles (denoted as PN1-PN4), showing absorption maxima at 766, 775, 770, and 775 nm, respectively (Fig. 2b). Under same concentration, PN3 exhibited the highest fluorescence emission at 819 nm (Fig. 2c) followed by PN2 and PN1; while PN4 had negligible fluorescence peak above 800 nm. The photodynamic properties of PN1-PN4 were studied. The generation of 1O2, O2•- and •OH for PN1-PN4 under NIR irradiation were detected by singlet oxygen sensor green (SOSG), dihydrorhodamine 123 (DHR123) and hydroxyphenyl fluorescein (HPF), respectively. The fluorescence intensity of SOSG gradually increased for PN1-PN4 under continuous NIR irradiation, proving the generation of 1O2. Among them, PN2 produced the highest amount of 1O2, increasing the fluorescence signal of SOSG by 21.41-fold, followed by PN1 with a 12.46-fold increase (Fig. 2d). Different from the trend in the Type-II PDT ROS generation, PN1 exhibited the highest Type-I PDT ROS generation capability among all nanoparticles (Fig. 2e, f), showing 22.83-fold and 12.98-fold fluorescence increases for DHR123 (for O2•- determination) and HPF (for •OH determination), respectively. This is significantly higher than PN2-PN3 with the DHR123 fluorescence enhancements ranging from 8.13 to 15.90-fold and the HPF enhancements ranging from 4.71 to 12.34-fold.
a Chemical structure of P1-P4. b UV-Vis absorption and c Fluorescence spectra of F127 coated P1-P4 (PN1-PN4) in H2O, arb. units, arbitrary units. Generation of d 1O2, e O2•- and f •OH determined by fold changes of fluorescence intensity of SOSG, DHR123 and HPF in NP1-NP4 ([BTP-OD] = 5 μg/mL) solution after NIR irradiation (808 nm, 0.2 W/cm2) for different time (n = 3 independent experiments). g Molecular energy level of P1-P4 using cyclic voltammetry (0.1 mol L−1 Bu4NPF6 dichlorobenzene/ac etonitrile solutions at a scan rate of 50 mVs−1). h Mechanism of Type-I PDT. The mean values and SD are presented. Source data are provided as a Source Data file.
The mechanism of Type-I ROS generation was investigated by determining the HOMO and LUMO energy levels of P1-P4 using cyclic voltammetry (CV) (Fig. 2g, Supplementary Fig. 1). The LUMO energy levels of P1, P2, P3, and P4 were −4.29 eV, −4.30 eV, −4.23 eV and −4.21 eV, respectively. Given that the LUMO levels of P1 and P2 are closer to the redox potentials of O2/O2•− (−4.30 eV)50 compared to those of P3 and P4, electrons can more easily transfer from excited PN1 and PN2 to O2, resulting in more efficient generation of O2•−. However, the HOMO energy levels of P1-P4 (−5.03 eV, −5.02 eV, −5.37 eV, and −5.25 eV) were higher than EH of H2O/•OH (−6.70 eV), which hindered •OH generation by direct oxidation of H2O. Thus, •OH generated from PN1 was likely derived from the transfer of electrons in O2•– to H2O31. The mechanism for the efficient Type-I ROS generation by PN was proposed in Fig. 2h. Under NIR irradiation, the electrons of PN are excited from ground states (S0) to higher energy excited states (Sn) to obtain singlet excitons, which can be efficiently populated from the triplet states (Tn) to the triplet excitons at the lowest level (T1) state via intersystem crossing. Subsequently, electrons can be transferred to O2 to generate O2•-, and then further transferred to H2O to produce •OH. Due to PN1 had the highest O2•- and •OH generation ability among all nanoparticles, P1 was selected as the optimal Type-I photosensitizer backbone for construction of hypoxia-tolerant polymeric photosensitizer (HTPS) in the following studies.
Synthesis and characterization of polymeric photosensitizer prodrug
Based on the best-performing polymer P1, the BDT-BTP-BDT-DPP-Br with Type-I PDT ability was first synthesized via Stille polycondensation of 3 monomers, including BDT, BTP-2Br-OD, and DPP (Fig. 3a). BDT-BTP-BDT-DPP-Br was then reacted with sodium azide to afford BDT-BTP-BDT-DPP-N3 for post-functionalization. To obtain CatB-responsive Niclosamide prodrug, the alkynyl modified CatB cleavable peptide (Alk-Val-Cit) was synthesized and conjugated with Niclosamide through a self-immolative linker to get Alk-Val-Cit-PAB-Niclo. Then, Alk-Val-Cit-PAB-Niclo was further modified into NH2-Val-Cit-PAB-Niclo and reacted with the carboxyl group on polyethylene glycol (PEG, Mw=2000), affording Niclosamide prodrug-modified alkyne-PEG (Alk-PEG2k-Niclo), with 85.7% grafting efficiency base on NMR analysis (Supplementary Figs. 2–8). Thereafter, BTP-BDT-DPP-N3 was reacted with Alk-PEG2k-Niclo through click reaction and then self-assembled into HTPSNiclo in aqueous solution. The control nanoparticle HTPS without Niclosamide prodrug was synthesized through the click reaction between BTP-BDT-DPP-N3 and Alk-m PEG2k (Supplementary Fig. 9).
a Synthetic route of HTPS and HTPSNiclo. (i) Tris(dibenzylideneacetone)dipalladium(0) Pd2(dba)3, Tri(o-tolyl)phosphine, chlorobenzene, 120 °C, 2 h; (ii) Sodium azide, tetrahydrofuran (THF)/N,N-dimethylformamide (DMF), r.t., 12 h; (iii) Alk-PEG2k, CuBr, N,N,N’,N”,N”’ pentamethyldiethylenetriamine (PMDETA), THF, r.t., 24 h; (iv) Alk-PEG2k -Niclo, CuBr, N,N,N’,N”,N”’ pentamethyldiethylenetriamine (PMDETA), THF, r.t., 24 h; (v) Aqueous self-assembly of BDT-BTP-BDT-DPP-PEG; (vi) Aqueous self-assembly of BDT-BTP-BDT-DPP-PEG-Niclo. b DLS profile and representative TEM images of HTPS and HTPSNiclo. c Fluorescence spectra of HTPS and HTPSNiclo with an excitation wavelength at 760 nm, arb. units, arbitrary units. d Representative confocal fluorescence images of 4T1 cancer cells after incubation with HTPS and HTPSNiclo ([BTP-OD] = 1 μg/mL) for 24 h, followed by staining with lysosome tracker (Green DND-26) and cell nuclei dye (Hoechst 33342) (Scale bar: 10 μm). Generation of e 1O2, f O2•- and g •OH determined by fold changes of fluorescence intensity of SOSG, DHR123, and HPF in HTPS and HTPSNiclo ([BTP-OD] = 2.5 μg/mL) solution after NIR irradiation (808 nm, 0.2 W/cm2) for different time (n = 3 independent experiments). Relative cell viabilities of 4T1 cells after incubation with HTPS and HTPSNiclo for 12 h at different SPN concentrations with or without NIR irradiation (0.2 W/cm2 at 808 nm) for 5 min (n = 3 independent experiments) in (h) hypoxia or i normal O2 conditions. *p = 0.04 HTPSNiclo versus HTPS, ****p = 2.7 × 10−5 HTPS + NIR versus HTPS, ****p = 9.2 × 10−5 HTPSNiclo + NIR versus HTPS in (h). ***p = 5.8 × 10−4 HTPS + NIR versus HTPS, ***p = 5.9 × 10−4 HTPSNiclo + NIR versus HTPS in (i). Confocal fluorescence images (j, l) and MFI (k, m) showing intracelluar generation of ROS and O2•- in 4T1 cells after incubation with HTPSNiclo ([BTP-OD] = 5 μg/mL) for 12 h and NIR irradiation (808 nm, 0.2 W/cm2) for 5 min under normoxia and hypoxia conditions (1: PBS; 2: HTPS; 3: HTPSNiclo) (Scale bar: 50 μm), arb. units, arbitrary units. ***p = 2.7 × 10-4 HTPS + NIR + O2 versus PBS + NIR + O2, ***p = 2.6 × 10−4 HTPSNiclo + NIR + O2 versus PBS + NIR + O2. ****p = 4.6 × 10−5 HTPS + NIR - O2 versus PBS + NIR - O2, ****p = 6.2 × 10−5 HTPSNiclo + NIR - O2 versus PBS + NIR - O2 in k. ***p = 8.3 × 10−4 HTPS + NIR + O2 versus PBS + NIR + O2, ***p = 1.0 × 10−4 HTPSNiclo + NIR + O2 versus PBS + NIR + O2. ****p = 2.1 × 10−5 HTPS + NIR - O2 versus PBS + NIR - O2, ****p = 5.9 × 10−5 HTPSNiclo + NIR - O2 versus PBS + NIR - O2 in m. The experiments in (b, d) were repeated independently three times with similar results. Statistical significance was calculated via one-way ANOVA with a Tukey posthoc test. The mean values and SD are presented. Source data are provided as a Source Data file.
As shown in Transmission Electron Microscopy (TEM) images, both HTPS and HTPSNiclo exhibited spherical morphology with uniform size distribution. Dynamic Light Scattering measurements revealed that hydrodynamic diameters of HTPS and HTPSNiclo ranged from 19 nm to 26 nm with good dispersity in PBS (Fig. 3b). The stabilities of HTPS and HTPSNiclo were evaluated by measuring their hydrodynamic size in presence of PBS and FBS. Both of them exhibited minimal hydrodynamic size changes up to 72 h, indicating their good biostability (Supplementary Fig. 10). A slightly red-shifted fluorescence emission for HTPSNiclo (838 nm) was observed compared to HTPS (833 nm) (Fig. 3c). In addition, hematological analysis was performed by incubation both HTPS and HTPSNiclo with red blood cells (RBCs) for 24 h, and no hemolysis were detected, indicating their good biosafety. (Supplementary Fig. 11). The cellular uptake of HTPS and HTPSNiclo was evaluated in 4T1 cells using confocal laser scanning microscopy. The obvious fluorescent signals of HTPS and HTPSNiclo were detected, which were similarly 11-fold higher than PBS group. Moreover, the signals of nanoparticles colocalized well with lysosomal tracker, suggesting that both HTPS and HTPSNiclo underwent endocytosis through the lysosomal pathway (Fig. 3d and Supplementary Fig. 12). Then, photodynamic ROS generation properties of HTPS and HTPSNiclo were evaluated. After NIR irradiation (808 nm, 0.2 W/cm2) for 3 min, the fluorescence intensity of DHR123 in HTPS and HTPSNiclo solutions increased by ~5.65-fold and 5.05-fold, respectively. Similarly, the HPF fluorescence intensity in HTPS and HTPSNiclo solutions increased by about 6.24-fold and 5.98-fold, respectively. In contrast, the SOSG fluorescence intensity increased slightly, which was 1.75-fold and 1.65-fold for HTPS and HTPSNiclo, respectively (Fig. 3e–g). These data demonstrated that HTPS and HTPSNiclo exhibited effective Type-I rather than Type-II photodynamic properties.
Afterward, HTPS and HTPSNiclo photodynamic cytotoxicity was evaluated by 3-(4,5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium (MTS) assay under normoxic and hypoxic conditions. Both nanoparticles only exhibited strong cytotoxicities after NIR irradiation (808 nm, 0.2 W/cm2) for 5 min. Under normoxic conditions, HTPS and HTPSNiclo showed significant photodynamic cytotoxicity to 4T1 cells in a concentration-dependent manner, with cell viabilities of 24.78% and 19.20%, respectively, at a concentration of 20 μg/mL. More importantly, both HTPS and HTPSNiclo induced obvious photodynamic cytotoxicity under hypoxic conditions, demonstrating their excellent hypoxia tolerance in PDT (Fig. 3h, i). The cytotoxic effect of Niclosamide on 4T1 cells was evaluated. Niclosamide exhibited slight cytotoxicity towards 4T1 cells at the concentrations up to 2 μM (Supplementary Fig. 14).
To investigate the mechanism for the cellular cytotoxicity, the intracellular photodynamic ROS and O2•- generation were evaluated by turn on probes 2′,7′-dichlorodihydrofluorescein (H2DCFDA) and DHR123 under normoxic and hypoxic conditions (Fig. 3j–m). Under normoxic conditions, the mean fluorescence intensities (MFI) of DCF in HTPS and HTPSNiclo + NIR group were 4.07-fold and 3.80-fold higher than the PBS group, respectively. In addition, MFI of DHR123 in HTPS and HTPSNiclo + NIR group were 6.58-fold and 6.29-fold stronger than the PBS group, respectively. Under hypoxic conditions, the photodynamic total ROS and O2•- generation abilities of HTPS and HTPSNiclo were comparable to their performance under normoxic conditions. Similar results could be got from the flow cytometry analysis (Supplementary Fig. 13), where significant DCF and DHR123 emission was detected, indicating that both HTPS and HTPSNiclo-treated 4T1 cells could generate photodynamic total ROS and O2•- after NIR irradiation (808 nm, 0.2 W/cm2) for 5 min. The hypoxia tolerance of HTPS and HTPSNiclo mediated Type-I PDT can be explained by the recycling of O2, leading to less O2 consumption than Type-II PDT.
In vitro photo-immunotherapy
To confirm the CatB-specific drug release behavior of HTPSNiclo, high performance liquid chromatography (HPLC) analysis was conducted to monitor the Niclosamide release after incubating HTPSNiclo with CatB. A new elution peak at 11.91 min, attributed to free Niclosamide, gradually became higher with increased incubation time (Fig. 4a). It was calculated that 96.47% Niclosamide was released from HTPSNiclo after incubation with CatB for 5 h (Fig. 4b). In contrast, no Niclosamide peak was detected in the absence of CatB (Supplementary Fig. 15). To determine whether CatB was overexpressed in the 4T1 cell line, immunofluorescence (IF) staining and western immunoblot analysis of expression levels of CatB was performed. The results clearly showed that the expression of CatB in 4T1 cancer cells was significantly higher compared to normal 3T3 cells (Fig. 4c and Supplementary Fig. 16). Then, the cellular STAT3 inhibition ability of HTPSNiclo was studied. After incubating HTPSNiclo with 4T1 cells for 24 h, the phosphorylation of STAT3 (p-STAT3) was 11.67-fold lower compared to PBS group, as shown in Fig. 4d. This can be attributed to the successful release of free Niclosamide in CatB-overexpressed 4T1 cells. To further confirm this, CA-074, a CatB inhibitor, was used to inhibit its cleavage activity51. As quantified in the western blotting data, the p-STAT3 expression in HTPSNiclo + CA-074 group exhibited 76.0-fold higher compared to the group treated with HTPSNiclo alone (Fig. 4d and Supplementary Fig. 17).
a HPLC profile for the release of Niclosamide from PEG2k in the absence or presence of CatB (0.2 U/mL). b Released profiles of Niclosamide from HTPSNiclo after incubation with CatB for different times (n = 3 independent experiments). c Representative confocal fluorescence images of CatB expression in 4T1 and 3T3 cells (Scale bar: 50 μm). d Representative western immunoblot analysis of expression levels of p-STAT3 in 4T1 cells with or without CA-074, the CatB inhibitor. Integrated density was measured by Image J software and labeled on the image (normalized to actin protein). MFI of e CRT expression in cell surface and f HMGB1 in the nuclei of 4T1 cells after different treatments. 1: PBS treatment, 2: HTPS treatment, 3: HTPSNiclo treatment. NIR irradiation: 0.2 W/cm2 at 808 nm for 5 min (n = 3 independent experiments), arb. units, arbitrary units. ***p = 2.7 × 10−4 and 2.8 × 10-4 group 2, 3 (+NIR) versus group Ctr in (e). *p = 1.1 × 10−2 and 1.5 × 10−2 group 2, 3 (+NIR) versus group Ctr in (f). g Relative levels of TNF-α and IL-12 secreted by BMDCs (n = 3 independent experiments). **p = 1.3 × 10−2 group 3 (- NIR) versus group Ctr (- NIR), ****p = 2.7 × 10−5 group 2 (+NIR) versus group Ctr (- NIR), ****p = 1.0 × 10−6 group 3 (+NIR) versus group Ctr (- NIR) for TNF-α. **p = 2.7 × 10−3 group 3 (- NIR) versus group Ctr (- NIR), ***p = 1.5 × 10−4 group 2 (+NIR) versus group Ctr (- NIR), ****p = 2.0 × 10−5 group 3 (+NIR) versus group Ctr (- NIR) for IL-12. The proportion of h matured DCs (CD80+CD86+) and i M2 macrophages (F4/80+CD11b+CD206+) after incubation with 4T1 cell supernatants after different treatments (n = 3 independent experiments) ***p = 2.9 × 10−4 group 3 (- NIR) versus group Ctr (- NIR), ****p = 7.6 × 10−5 group 2 (+NIR) versus group Ctr (- NIR), ****p = 8.8 × 10−5 group 3 (+NIR) versus group Ctr (- NIR) for DCs. ***p = 1.1 × 10−4 group 3 (- NIR) versus group Ctr (- NIR), **p = 4.0 × 10−3 group 2 (+NIR) versus group Ctr (- NIR), ****p = 1.9 × 10−5 group 3 (+NIR) versus group Ctr (- NIR) for macrophages. j Mechanism of ICD and p-STAT3 inhibition in inducing DCs maturation and downregulating M2 macrophage percentage. Statistical significance was calculated via one-way ANOVA with a Tukey posthoc test. The experiments in c and d were repeated independently three times with similar results. Group arrangement: Ctr: Untreated BMDCs, 1: PBS treated 4T1 cells, 2: HTPS treated 4T1 cells, 3: HTPSNiclo treated 4T1 cells; *p < 0.05, **p < 0.01, ***p < 0.001, and ****p < 0.0001. The mean values and SD are presented.
To study the ICD of 4T1 cells after HTPSNiclo-mediated Type-I PDT, the calreticulin (CRT) and high-mobility group box 1 (HMGB1) were then evaluated (Fig. 4e, f and Supplementary Figs. 18, 19). In contrast to the PBS group, cells treated by HTPS and HTPSNiclo-mediated PDT showed a 6.12-fold and 6.52-fold higher MFI of CRT signals on the plasma membrane, respectively. A more quantitative evaluation of CRT exposure by flow cytometry confirmed a 4.92-fold and 4.32-fold higher proportion of 4T1 cells undergoing ICD after HTPS and HTPSNiclo-mediated PDT treatment, respectively, compared to PBS treatment. In addition, the nuclear HMGB-1 signals after HTPS and HTPSNiclo-mediated PDT were 7.65-fold and 5.16-fold lower compared to PBS group, respectively. These results indicated that potent ICD was induced by HTPSNiclo-mediated Type-I PDT. To study the immune activation effect of ICD and immune modulation ability of HTPSNiclo, bone marrow derived dendritic cells (BMDCs) and macrophages were incubated with 4T1 cells’ supernatants that underwent different treatments. As shown in Fig. 4g, TNF-α and IL-12 secretion in BMDCs treated with 4T1 tumor supernatants underwent HTPSNiclo-mediated PDT (HTPSNiclo + NIR) was 1.36-fold and 1.69-fold higher than that in HTPS + NIR group, respectively. This could be attributed to p-STAT3 inhibition by the released Niclosamide from HTPSNiclo. The flow cytometry results showed that the matured DCs in HTPS + NIR group and HTPSNiclo + NIR group was 3.39-fold and 2.61-fold higher than the untreated group, respectively (Fig. 4h and Supplementary Fig. 20a, b). Moreover, HTPSNiclo + NIR treatment group induced the lowest percentage of M2 macrophages, which was 8.22-fold lower than HTPS + NIR group (Fig. 4i and Supplementary Fig. 20c). The immune activation of Niclosamide towards DC maturation and downregulation of M2 macrophage was also investigated as shown in Supplementary Fig. 21. After 24 h of Niclosamide treatment, the proportion of mature DCs increased, while M2 macrophages decreased in a concentration-dependent manner. In summary, the synergistic effects of ICD and p-STAT3 inhibition effectively induced DC maturation and downregulated M2 macrophage percentage, with the potential to reverse the immunosuppressive TME and enhance the tumor suppression.
In vivo photodynamic immunotherapy
The pharmacokinetics of HTPS and HTPSNiclo were investigated in Supplementary Fig. 22, indicating that HTPS could be metabolized normally in vivo, demonstrating their good biosafety. The potential off-target effects and best timepoint for in vivo photo-irradiation study of HTPS and HTPSNiclo were first evaluated by evaluating the accumulation of nanoparticles in tumor tissues using NIR fluorescence (NIRF) imaging in 4T1 tumor-bearing mice. The NIRF intensity in tumors of HTPSNiclo- and HTPS-injected mice gradually increased, reaching similar maximum levels at 12 h post-injection, showing 15.33-fold and 15.42-fold increase compared to the background, respectively (Fig. 5a, b). Ex vivo fluorescence imaging at 48 h post-injection demonstrated that the highest nanoparticle accumulation was in tumors, followed by spleen and liver (Supplementary Fig. 23). To verify the photodynamic activity of HTPSNiclo and HTPS in vivo, DHR123 was used to measure the generation of O2•- in tumors. The fluorescence signals of DHR123 in tumors treated with HTPSNiclo + NIR and HTPS + NIR group were 5.03-fold and 5.05-fold higher compared to saline group, respectively (Fig. 5c and Supplementary Fig. 24).
NIR fluorescence imaging (a) and MFI b of 4T1 tumor-bearing BALB/c mice after intravenous injection of HTPS and HTPSNiclo at different time points (200 µL, [BTP-OD] = 200 μg/mL) (n = 3 independent experiments). c Quantitative analysis of DHR123 MFIs in primary tumor tissues from mice in different group. NIR irradiation: 0.2 W/cm2 at 808 nm for 5 min (n = 3 independent experiments). ***p = 6.5 × 10-4 HTPS versus saline, ***p = 2.8 × 10-4 HTPSNiclo versus saline. d Schedule for remote tumor model implantation and HTPSNiclo-mediated photo-immunotherapy. Growth curves of e primary tumors and f distant tumors on 4T1 remote tumor model (n = 5 independent mice). **p = 1.1 × 10−3 HTPSNiclo + NIR versus HTPS + NIR in (e), ****p = 4.3 × 10−5 HTPSNiclo + NIR versus HTPS + NIR in (f). g Histological H&E staining of lung in 4T1 tumor-bearing mice after different treatment. Images are representative of three biologically independent mice (Scale bar: 500 μm). h Tumor metastasis area in lungs of mice in 4T1 tumor-bearing mice after different treatment. **p = 2.9 × 10−3 HTPSNiclo versus saline, *p = 1.2 × 10−2 HTPS + NIR versus saline, **p = 1.1 × 10−3 HTPSNiclo + NIR versus saline. i Survival curves for the mice after different treatments (n = 5 independent mice). j Schedule for rechallenged tumor model establishment and HTPSNiclo-mediated photo-immunotherapy. k Growth curves of rechallenged tumors (n = 5 independent mice). ****p = 1.4 × 10−5 HTPSNiclo + NIR versus saline. l Number of lung metastasis nodules after different treatments on rechallenged tumor model (n = 3 independent experiments). ****p = 4.1 × 10−5 HTPSNiclo + NIR versus saline. m Images of lung metastasis after staining with India ink. Statistical significance was calculated via one-way ANOVA with a Tukey posthoc test. *p < 0.05, **p < 0.01, ***p < 0.001, and ****p < 0.0001. The mean values and SD are presented. Source data are provided as a Source Data file.
The therapeutic efficacy of HTPSNiclo-mediated Type-I photo-immunotherapy was then evaluated using a bilateral animal model (Fig. 5d). NIR irradiation (808 nm, 0.2 W/cm2, 5 min) was locally conducted on primary tumor 12 h post HTPSNiclo injection. The growth of both primary and distant tumors in different treatment groups, as well as the survival of mice, were monitored. In the HTPS + NIR group, the growth of primary tumors was inhibited until Day 10, but the tumor growth accelerated afterwards. In contrast, the tumors in the HTPSNiclo + NIR group remained well-inhibited throughout the experiment. The inhibition rate of primary tumor growth in the HTPSNiclo + NIR and HTPS + NIR group was 91.23% and 75.53%, respectively (Fig. 5e). A more pronounced difference was observed in the distant tumor growth curve. The HTPSNiclo + NIR group shown an inhibition rate of 88.22%, much higher than the 37.99% observed in the HTPS + NIR group (Fig. 5f). To further confirm the inhibition of tumor metastasis by HTPSNiclo-mediated photo-immunotherapy, the hematoxylin and eosin (H&E) staining of lung tissues was conducted. No lung metastasis lesions were observed in lung tissue from the mice in HTPSNiclo + NIR group. In contrast, the lung metastasis area was 64.52%, 13.41% and 14.17% in the saline group, HTPS Niclo group, and HTPS + NIR group, respectively (Fig. 5g, h). In the HTPS + NIR group, 100% survival was observed up to Day 39. By Day 60, only 60% of the mice remained alive (Fig. 5i). In contrast, all mice in the HTPSNiclo + NIR group survived through Day 60. The data demonstrated that HTPSNiclo-based photo-immunotherapy not only more effectively inhibited the growth of both primary and distant tumors but also significantly extended mice survival, compared to HTPS-based single PDT. The therapeutic effects were further confirmed by IF staining of caspase-3 and H&E staining analysis. Significant tumor cell apoptosis was observed in both primary and distant tumors in the HTPSNiclo + NIR group. In contrast, no apoptosis was observed in the distant tumors in the HTPS + NIR group (Supplementary Fig. 25).
To further investigate the long-term anti-tumor recurrence and anti-metastasis ability of HTPSNiclo-mediated photo-immunotherapy, a 4T1 tumor rechallenge model was established (Fig. 5j). The growth of rechallenged tumors in HTPSNiclo + NIR group was greatly suppressed compared to that of the saline group (Fig. 5k and Supplementary Fig. 26). At the end of rechallenged tumor monitoring, the lungs of mice were stained with India ink and the number of lung metastasis nodules was counted. Negligible tumor metastasis was found in the lungs of mice treated with HTPSNiclo + NIR on Day 57, while more than 40 white nodules were found in saline group (Fig. 5l, m). These results further demonstrated that HTPSNiclo-mediated photo-immunotherapy can effectively prevent tumor recurrence and lung metastasis. During the treatment process, the body weight of HTPSNiclo- and HTPS-injected mice did not change significantly compared to saline group. The immune response over time after HTPSNiclo injection was evaluated by the detection of IL-12, TNF-α, IL-10, IFN-γ in the serum. These cytokines were upregulated (IL-12, TNF-α, IFN-γ) or downregulated (IL-10) on Day 3 compared to Day 0 and returned to normal levels by Day 7. In addition, H&E staining of the major organs (heart, liver, spleen, kidney) in HTPSNiclo- and HTPS-injected mice showed similar physiological morphologies compared to those in saline group, indicating the good biocompatibilities of HTPSNiclo and HTPS. Pharmacokinetics, biodistribution, H&E staining of the major organs and immune response over time have proven the long-term Safety of HTPSNiclo (Supplementary Figs. 26–29).
Mechanistic study of photodynamic immunotherapy
To gain insight into the mechanism of HTPSNiclo-mediated photo-immunotherapy, immunological characterizations including tumor ICD, immune cell populations, immunostimulatory cytokines levels, angiogenesis level and p-STAT3 inhibition in nanoparticle-injected tumor-bearing mice were analyzed. First, IF staining of HMGB1 and CRT was used to visualized ICD in primary tumors. The MFI of HMGB1 in the tumor tissues of HTPS + NIR and HTPSNiclo + NIR group was 4.90- and 5.19-fold higher compared to the saline group, respectively (Fig. 6a and Supplementary Fig. 30). Similarly, MFI of CRT in HTPS + NIR and HTPSNiclo + NIR group was 3.07- and 3.11-fold higher than that in saline group, respectively (Fig. 6b and Supplementary Fig. 31). Then, flow cytometry was used to analyze changes in immune cell populations in tumor-draining lymph nodes (TDLNs) and tumors after HTPSNiclo-mediated photo-immunotherapy (Supplementary Fig. 32). The mature DC ratio in TDLNs of HTPSNiclo + NIR group was 4.29-fold higher than that in the saline group and 1.51-fold higher than that in the HTPS + NIR group, respectively. (Fig. 6c and Supplementary Fig. 33). Moreover, the immunosuppressive M2 macrophages in tumors of HTPSNiclo + NIR group was 14.4-fold lower compared to saline-treated mice and 14.1-fold lower compared to HTPS + NIR group, respectively (Fig. 6d and Supplementary Fig. 33). Considering the essential role of tumor-infiltrating T lymphocytes in combating tumors, the CD8+ T cells, CD4+ T cells and regulatory T cells in primary and distant tumors were further evaluated. The HTPSNiclo-mediated photo-immunotherapy induced the highest CD8+/Treg and CD4+/Treg ratio in both primary and distant tumors (Fig. 6e, f and Supplementary Figs. 34, 35). In detail, the CD8+/Treg ratio in the primary and distant tumor after HTPS + NIR treatment was 2.32-fold and 2.17-fold higher relative to that in saline group, respectively. In contrast, HTPS-mediated PDT could not upregulate the ratio of CD8+/Treg and CD4+/Treg in distant tumor. These results demonstrated that HTPSNiclo-mediated photo-immunotherapy promoted DC maturation, downregulated M2 macrophage percentage, and induced a robust T cell response in both primary and distant tumors.
Quantification of a HMGB1 and b CRT expression in the tumor after different treatments (n = 3 independent experiments), arb. units, arbitrary units. ***p = 1.0 × 10−3 HTPS + NIR versus saline, ***p = 4.4 × 10−4 HTPSNiclo + NIR versus saline for HMGB1 expression. ****p = 1.0 × 10−3 HTPS + NIR versus saline, ***p = 1.3 × 10−4 HTPSNiclo + NIR versus saline for CRT expression. Proportion of c matured DCs from the TDLNs and d M2 macrophages in the tumor of 4T1 tumor-bearing mice after different treatments (n = 3 independent experiments). **p = 4.8 × 10−3 HTPSNiclo versus saline, **p = 1.4 × 10−3 HTPS + NIR versus saline, ****p = 4.1 × 10−5 HTPSNiclo + NIR versus saline for the ratio of mature DCs. **p = 1.0 × 10−2 HTPSNiclo versus saline, **p = 1.0 × 10−2 HTPS + NIR versus saline, ***p = 1.4 × 10-4 HTPSNiclo + NIR versus saline for the ratio of macrophages. e Ratio of CD8+ T cell to Treg and f CD4+ T cell to Treg in the primary and distant tumor after different treatments (n = 3 independent experiments). **p = 9.0 × 10-3 HTPS + NIR versus saline, **p = 1.1 × 10−3 HTPSNiclo + NIR versus saline for the ratio of CD8+ T cell to Treg in primary tumor. ***p = 8.7 × 10−4 HTPSNiclo + NIR versus saline for the ratio of CD8+ T cell to Treg in distant tumor. **p = 4.8 × 10−3 HTPSNiclo versus saline, **p = 9.8 × 10−3 HTPS + NIR versus saline, ***p = 2.6 × 10−4 HTPSNiclo + NIR versus saline for the ratio of CD4+ T cell to Treg in primary tumor. **p = 5.8 × 10−3 HTPSNiclo versus saline, ***p = 1.0 × 10−4 HTPSNiclo + NIR versus saline for the ratio of CD4+ T cell to Treg in distant tumor. g Cytokine detection in the serum from 4T1 tumor-bearing mice after different treatments (n = 3 independent experiments). ***p = 6.6 × 10−4 HTPSNiclo versus saline, *p = 5.0 × 10−2 HTPS + NIR versus saline, ***p = 2.4 × 10−4 HTPSNiclo + NIR versus saline for IL-12 secretion. **p = 4.6 × 10-3 HTPS + NIR versus saline, ****p = 9.4 × 10−5 HTPSNiclo + NIR versus saline for TNF-α secretion. ***p = 1.9 × 10−4 HTPSNiclo versus saline, ***p = 3.6 × 10−4 HTPS + NIR versus saline, ***p = 2.2 × 10−4 HTPSNiclo + NIR versus saline for IL-10 secretion. **p = 9.8 × 10−3 HTPSNiclo versus saline, ***p = 5.9 × 10−4 HTPS + NIR versus saline, ***p = 1.8 × 10−4 HTPSNiclo + NIR versus saline for IFN-γ secretion. h HIF-1α and CD31 expression in tumors after different treatments (Scale bar: 100 μm). i Western blotting analysis of expression levels of p-STAT3 in tumors after different treatments. Integrated density was measured by Image J software and labeled on the image (normalized to actin protein). j, k Quantification of effector memory T cell in HTPSNiclo + NIR group and saline group (n = 3 independent experiments). ***p = 1.4 × 10−4 HTPSNiclo + NIR versus saline. l Mechanism of immune activation after HTPSNiclo-mediated photo-immunotherapy. The experiments in (h, i) were repeated independently three times with similar results. Statistical significance was calculated via one-way ANOVA with a Tukey posthoc test. *p < 0.05, **p < 0.01, ***p < 0.001, and ****p < 0.0001. The mean values and SD are presented. Source data are provided as a Source Data file.
Next, immunostimulatory cytokines including IL-12, IFN-γ, TNF-α and immunosuppressive cytokine IL-10 in the serum of mice were analyzed by enzyme-linked immunosorbent assay (ELISA) after various treatments. The serum concentration of TNF-α, IL-12, and IFN-γ in HTPSNiclo + NIR treated mice were 1.40-, 3.42- and 1.40-fold higher than that in saline treated mice, respectively. In addition, the IL-10 level in HTPSNiclo + NIR group was 3.45-fold lower compared with saline treated group (Fig. 6g). In general, the oxygen consumption and damaged vessels after PDT are two main factors for hypoxia condition, leading to the expression of HIF-1α and further angiogenesis in TME. Therefore, the HIF-1α and angiogenesis level in immunosuppressive TME after HTPSNiclo-mediated photo-immunotherapy were further evaluated. The expression of HIF-1α in tumor tissues after HTPS + NIR treatment was 1.26-fold higher compared to the saline group, indicating that single-mode HTPS-mediated PDT could not completely overcome the hypoxic TME (Fig. 6h and Supplementary Fig. 36). In contrast, HTPSNiclo-mediated photo-immunotherapy resulted in 3.28-fold lower HIF-1α expression compared to the saline group. Considering that HIF-1α is involved in tumor angiogenesis, CD31, the indicator of newly generated vessels, was stained in the tumors after different treatment. As expected, the new vascular area in HTPS + NIR and HTPSNiclo + NIR treated mice were 1.84-fold larger and 4.14-fold smaller compared to saline group, respectively (Fig. 6h and Supplementary Fig. 37), indicating the higher angiogenesis level in single-mode HTPS-mediated PDT treated mice compared to that of HTPSNiclo-mediated photo-immunotherapy. To further evaluate HTPSNiclo could inhibit HIF-1α and angiogenesis level through STAT3 pathway, western blotting analysis on the p-STAT3 in tumor lysis was conducted. In both the NIR irradiation and non-irradiation group, HTPSNiclo treatment significantly downregulated the p-STAT3 level, which can be attributed to the release of STAT3 inhibitor Niclosamide (Fig. 6i and Supplementary Fig. 38). The immune memory effect induced by HTPSNiclo-mediated photo-immunotherapy was further evaluated. As shown in Fig. 6j, k, the proportion of effector memory T cells (CD8+CD44+CD62L-) in HTPSNiclo + NIR group was 1.49-fold higher than that in the saline group.
Collectively, the synergistic effects of Type-I PDT and p-STAT3 inhibition in HTPSNiclo-mediated photo-immunotherapy was summarized in Fig. 6l. In a traditional Type-II PDT, O2 was largely consumed to generate 1O2. This consumption leads to hypoxia environment and HIF-1α upregulation, resulting in tumor angiogenesis. Although ICD of tumor cells induces a tumor-specific immune response, the immune tolerance after PDT leads to an immunosuppressive TME. This dilemma was effectively solved by HTPSNiclo-mediated photo-immunotherapy. This approach first generates Type-I ROS in an O2 recyclable manner, enhancing hypoxia-tolerance. Additionally, it releases Niclosamide through CatB cleavage, which inhibits pSTAT-3, leading to a reduction in HIF-1α, immunosuppressive cells and cytokines. Simultaneously, it promotes the secretion of immunostimulatory cytokines, enhances DCs maturation and T cell activation. These effects collectively reverse the immunosuppressive TME and improve tumor treatment efficacy.
Discussion
By tuning the structure units, we synthesized a series of photodynamic semiconducting polymers (SPs) and evaluated their efficiency in exerting the Type-I PDT process to generate O2•– and •OH, which were more toxic than 1O2. The mechanism for the difference in generating O2•– and •OH by SPs can be explained by how their HOMO/LUMO energy levels align with EH of O2/ O2•– (−4.3 eV) and EH of H2O/•OH (−6.7 eV). Although the LUMO energy levels of the four SPs all exceeded −4.3 eV, the LUMO of P1 was the closest to −4.3 eV among all SPs (Fig. 2g). It is thus the easiest for electrons to be transferred from excited P1 to O2, resulting in the most efficient production of O2•– from PN1. Since the HOMO energy levels of SPs were higher than EH of H2O/•OH (−6.7 eV), it was theoretically impossible for them to directly oxidize H2O into •OH. Thus, •OH was most likely to be derived from O2•– through its electron transfer to H2O. Given its highest efficiency in generating O2•– and •OH, P1 was further constructed into the hypoxia-tolerant polymeric photosensitizer (HTPS) for tumor-specific cancer photo-immunotherapy.
By virtue of the advantages of hypoxia tolerant capability of Type-I PDT, HTPS-based single therapy had already enhanced tumor inhibition efficiency (from 50% to 75.53%) and extended the 100% survival rate (from 25 days to 39 days) in 4T1 tumor-bearing mice compared to reported Type-II PDT52. Yet, it failed to effectively inhibit the growth of distant tumors (Fig. 5f). Furthermore, lung metastasis was observed due to the HIF-1α upregulation after HTPS-based single-mode PDT (Fig. 5g). In contrast, HTPSNiclo-mediated photo-immunotherapy released Niclosamide in the TME, inhibiting the phosphorylation of STAT3 and downregulating HIF-α expression by 7.63-fold compared to HTPS-based PDT. In addition, HTPSNiclo-based therapy could polarize the immunosuppressive TME into an active phenotype, resulting in higher DC maturation, CD8+/Treg and CD4+/Treg ratios in both primary and distant tumor compared to HTPS-based therapy (Fig. 6c–f). Furthermore, the levels of immunostimulatory cytokines TNF-α, IL-12, and IFN-γ after HTPSNiclo-mediated photo-immunotherapy were 1.40-, 3.42- and 1.40-fold higher compared to HTPS-based PDT, respectively (Fig. 6g). As a result of synergetic effect, HTPSNiclo-mediated photo-immunotherapy significantly improved tumor inhibition rate (91.23%) and reduced lung metastasis area (1.56%) compared to HTPS-based PDT, which showed 75.53% tumor inhibition rate and 14.17% lung metastasis area. Finally, HTPSNiclo-mediated photo-immunotherapy significantly prolonged the 100% survival rate into 60 days, compared to 39 days in HTPS-based single-mode PDT.
In summary, we developed a TME-activatable hypoxia-tolerant polymeric photosensitizer prodrug HTPSNiclo that overcame the limitations of PDT. We revealed the importance of aligning HOMO/LUMO energy levels of polymeric photosensitizer with the redox potentials of (O2/ O2•–) to optimize the generation of Type-I photodynamic ROS generation and identified various Type-I polymeric photosensitizers. Coupled with the immune modulator Niclosamide to inhibit the phosphoralation of STAT3, HTPSNiclo-based photo-immunotherapy upregulated the level of immunostimulatory cytokines and downregulated HIF-1α, further mitigating the immunosuppressive TME following PDT. This study not only develops a set of polymer-based Type-I photosensitizers but also uncovers strategies to counteract the immunosuppressive TME for enhanced immunotherapeutic efficacy. Despite its effectiveness, PDT faces limitations such as its dependency on oxygen and challenges posed by the immunosuppressive TME, which can reduce therapeutic efficacy. However, advancements in the development of Type-I photosensitizers to address oxygen dependency, along with the integration of PDT with other treatments, particularly immunotherapies, present promising avenues for overcoming these challenges and improving cancer treatment outcomes.
Methods
Materials
All chemicals and reagents were purchased from Sigma-Aldrich unless otherwise stated. Y6-OD was purchased from derthon optoelectronics materials science technology co ltd. 4,7-bis(5-trimethylstannyl-2-thienyl)-2,1,3-benzothiadiazole, 2,5-bis(triMethylstannyl)th ieno[3,2‐b]thiophene, 1,1′-[4,8-Bis[(2-ethylhexyl)oxy]benzo[1,2-b:4,5-b′]dithiophene-2,6-diyl]bis[1,1,1-trimethylstannane] and 2,5-Bis(2-ethylhexyl)-3,6-bis(5-(trimethylstannyl)thiophen-2-yl)pyrrolo[3,4-c]pyrrole-1,4(2H,5H)-dione, 2,5-Bis(6-bromohexyl)-3,6-bis(5-bromothiophen-2-yl)pyrrolo[3,4-c]pyrrole-1,4(2H, 5H)-dione were purchased from BLD Pharmatech Ltd. PEG (Mw=2000) was purchased from Biopharma PEG Scientific Inc. Agarose powder (Certified Molecular Biology Agarose) was purchased from Bio-Rad (Hercules, CA, USA). SOSG was purchased from Molecular Probes Inc. (Carlsbad, CA, USA). 3-(4,5-Dimethylthiazol-2-yl)-5-(3carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium, inner salt (MTS) solution was purchased from Promega Corp. (Madison, WI, USA). Dulbecco’s modified eagle medium (DMEM), Trypsin–EDTA (0.05%), penicillin–streptomycin (10,000 U/mL), fetal bovine serum (FBS), ACK lysis, type I collagenase, and type IV collagenase were purchased from Gibco. Anti-HMGB1 and anti-caspase 3 were purchased from Abcam Inc. (Cambridge, CA, USA). ELISA kits for TNF-α, IFN-γ, IL-10, and IL-12 detection, anti-STAT3, anti-p-STAT3, anti-CD16/32, anti-CD8a, anti-CD3, anti-CD4, anti-CD80, anti-CD86, and anti-CD11c antibodies were purchased from BioLegend (USA). BCA protein assay kit was purchased from Thermo Fisher Scientific. All antibodies were purchased from Biolegend (USA) unless otherwise indicated. Antibody against CD16/32 (93, 1:50) was used to block the non-specific binding. Antibodies against CD11b (M1/70, 1:200), F4/80 (BM8, 1:20) and CD206 (C068C2, 1:100) were used for M2 macrophage staining. Antibodies against CD3 (17A2, 1:100) and CD8 (53-6.7, 1:100) were used for cytotoxic T cell staining. Antibodies against CD4 (RM4-5, 1:100) and FOXP3 (MF-14, 1:200) were used for Treg cell staining. Antibodies against CD62L (MEL-14, 1:100) and CD44 (IM7, 1:20) were used for memory T cell staining. Antibodies against CD11c (N418, 1:100), CD80 (16-10A1, 1:50), and CD86 (A17199A, 1:50) were used for mature dendritic cell staining. Tumour slices were stained with antibodies against high mobility group box 1 (HMGB-1) (3E3, Alexa Fluor 488-labeled, 1:200, Biolegend), Caspase-3 (#9664, rabbit anti-mouse, 1:400, Cell Signalling Technology), CD31 (MEC13.3, rat anti-mouse, 1:200, Biolegend), HIF-1A (mgc3, 1:200, Thermo Fisher Scientific) and rabbit anti-mouse calreticulin (PA3-900, 1:200, Thermo Fisher Scientific). Alexa Fluor 488-conjugated secondary antibody (goat anti-rabbit, 1:500, Thermo Fisher Scientific) was used in caspase-3 and calreticulin staining.
Instrumentation
Proton nuclear magnetic resonance (1H NMR) spectra were recorded on a Bruker AVIII 400 MHz NMR system (Bruker Physik AG, Germany). Electrospray ionization-mass spectrometry (ESI-MS) spectra were recorded by Thermo Finnigan Polaris Q quadrupole ion trap mass spectrometer. UV-Vis spectra were recorded on a Shimadzu UV-2450 spectrophotometer. Fluorescence measurements were performed on a Fluorolog 3TCSPC spectrofluorometer (Horiba Jobin Yvon). Diameter size and zeta potential were measured by a Malvern Nano-ZS Particle Sizer. The absorbance in 96-well plates was measured using a SpectraMax M5 microplate reader (Molecular Devices). TEM images were captured from a JEM 1400 transmission microscope with accelerating voltage ranging from 40 to 120 kV. Fluorescence imaging was performed on an IVIS imaging system (IVIS-CT machine, PerkinElmer). Fluorescence images of cells were acquired using a Laser Scanning Microscope LSM800 (Zeiss). Tissue sectioning was performed using a crystat (Leica). Western blotting images were obtained from iBright CL750 Imaging system.
Synthesis of P1-P4
For P1, Y6-OD-2Br (50.00 mg, 0.027 mmol), 4,7-bis(5-trimethylstannyl-2-thienyl)-2,1,3-benzothiadiazole (16.70 mg, 0.027 mmol), Tris(dibenzylideneacetone)dipalladium(0) (5 mg, 0.005 mmol) and Tri(o-tolyl)phosphine (10 mg, 0.0329 mmol) were dissolved in anhydrous chlorobenzene. The reaction mixture was stirred at 120 °C after being degassed by freeze-thaw cycles three times. The reaction was monitored by UV-Vis absorption. The product was precipitated in cold methanol, collected through centrifugation at 1000 × g for 10 min, washed thrice with methanol and dried under vacuum to obtain P1. P2-P4 was synthesized using 2,5-bis(triMethylstannyl)thieno[3,2‐b]thiophene (P2), 1,1′-[4,8-Bis[(2-ethylhexyl)oxy]benzo[1,2-b:4,5-b′]dithiophene-2,6-diyl]bis[1,1,1-trimethylstannane] (P3), and 2,5-Bis(2-ethylhexyl)-3,6-bis(5-(trimethylstannyl)thiophen-2-yl)pyrrolo[3,4-c]pyrrole-1,4(2H,5H)-dione (P4) under similar conditions.
In vitro generation of 1O2, O2 •–and •OH
The generation of 1O2 by P1-P4 was measured by adding 1 µL SOSG probe (500 µM) to 1 mL nanoparticle solutions ([BTP-OD] = 5 μg/mL), followed by NIR irradiation (808 nm, 0.2 W/cm2) for 5 min. The fluorescence intensity at 520 nm was recorded every 30 s. The generation of 1O2 was assessed as F/F0, where F0 represents the fluorescence intensity of each sample before NIR irradiation. DHR123 and HPF were used to measure the generation of O2•–and •OH respectively, with a method like that of the SOSG probe.
Electrochemical cyclic voltammetry (CV) tests and energy of conduction band and valence band calculation
The cyclic CV tests were conducted according to previously reported literature53. 0.1 mol/L tetrabutylammonium hexafluorophosphate (Bu4NPF6) dichlorobenzene/acetonitrile (5:1, v/v) solution as the supporting electrolyte with a scan speed at 0.05 V/s. Pt wire, glassy carbon discs, and Ag/AgCl were used as the counter, working electrode, and reference electrodes, respectively. To estimate whether the synthesized polymer (P1-P4) can generate reactive oxygen species (ROS), the energy levels of HOMO and LUMO were calculated according to CV tests.
Synthesis of amino-modified Val-Cit-PAB-Niclo (NH2-Val-Cit-PAB-Niclo)
Val-Cit-PAB-OH (200 mg, 0.53 mmol), 4-pentynoic acid (77 mg, 0.78 mmol), HOBT (105.3 mg, 0.78 mmol), HBTU (295.8 mg, 0.78 mmol) and DIPEA (185 μL, 1.06 mmol) were dissolved in anhydrous DMF (15 mL) and stirred under room temperature for 4 h to get Alk-Val-Cit-PAB-OH. Then, the Alk-Val-Cit-PAB-OH was purified by HPLC. 1H NMR (400 MHz, DMSO-d6) δ 9.89 (s, 1H), 8.12–8.10 (d, 1H), 7.95–7.93 (d, 1H), 7.56–7.54 (d, 2H), 7.25–7.23 (d, 2H), 4.43 (s, 2H), 4.39–4.35 (t, 1H), 4.25–4.21 (t, 1H), 3.01–2.96 (m, 2H), 2.75 (s, 1H), 2.38–2.36 (m, 4H), 2.00–1.95 (m, 1H), 1.73–1.70 (m, 1H), 1.62–1.58 (m, 1H), 1.44–1.36 (m, 2H), 0.87–0.84 (m, 6H). 13C NMR (400 MHz, DMSO-d6) δ 171.58, 171.07, 170.83, 159.37, 137.95, 137.86, 127.39, 119.32, 84.25, 71.70, 63.05, 58.14, 53.59, 39.08, 34.48, 30.93, 29.76, 27.26, 19.66, 18.61, 14.77. LCMS (ESI) m/z calcd for C23H33N5O5 [M + H]+ 460.25, found 460.30.
Then, Alk-Val-Cit-PAB-OH (36 mg, 0.078 mmol) was dissolved in 10 mL anhydrous THF. Then, phosphorus tribromide (37 μL, 0.39 mmol) was added dropwise into the reaction mixture under ice bath. Afterward, the reaction system was stirred for 3 h under ice bath and argon protection. The THF was removed by rotary evaporation, and saturated NaHCO3 (40 mL) was added to neutralize phosphorus tribromide. Subsequently, the product was extracted with ethyl acetate (EA, 150 mL, three times) and washed three times with NaHCO3 (30 mL). After drying by anhydrous sodium sulfate, the EA was removed by rotary evaporation. Then, Niclosamide (STAT3 inhibitor) (4.238 mg, 0.013 mmol), K2CO3 (1.1 mg, 0.0078 mmol), Cs2CO3 (2.5 mg, 0.0078 mmol) were added into the system. Anhydrous acetonitrile (ACN, 10 mL) was used to dissolve reactants, and the reaction was stirred under room temperature for 3 h. Alk-Val-Cit-PAB-Niclo (9.2 mg, 0.012 mmol) and 2-azidoethan-1-amine (5.2 mg, 0.06 mmol) were dissolved in dimethyl sulfoxide (DMSO, 2 mL). Then, CuSO4·5H2O (3 mg, 0.012 mmol) and L-ascorbic acid sodium (4.8 mg, 0.024 mmol) were dissolved in water (200 μL) and added into the DMSO solution quickly. After stirring at room temperature for 3 h, the NH2-Val-Cit-PAB-Niclo was purified by HPLC. 1H NMR (400 MHz, DMSO-d6) δ 10.70 (s, 1H), 10.00 (s, 1H), 8.72 (d, 1H), 8.35 (d, 1H), 8.30–8.27 (m, 1H), 8.19 (d, 1H), 8.02 (d, 1H), 7.88 (s, 1H), 7.69–7.63 (m, 3H), 7.48–7.45 (d, 3H), 6.03 (s, 1H), 5.44 (s, 2H), 4.56–4.53 (m, 2H), 4.42–4.37 (m, 1H), 4.23–4.19 (m, 1H), 3.05–3.02 (m, 3H), 2.99 (m, 1H), 2.89–2.85 (m, 2H), 2.01–1.96 (m, 1H), 1.75–1.70 (m, 1H), 1.66–1.61 (m, 2H), 1.58–1.54 (m, 1H), 1.49–1.45 (m, 1H), 1.41–1.36 (m, 1H), 0.84 (m, 6H). 13C NMR (400 MHz, DMSO-d6) δ 171.14, 171.09, 159.39, 155.64, 141.61, 141.20, 134.52, 134.21, 131.16, 130.53, 130.35, 129.88, 125.87, 125.28, 125.10, 124.36, 124.27, 124.09, 123.52, 122.93, 122.87, 121.97, 121.30, 119.92, 119.64, 119.55, 116.96, 84.24, 71.80, 71.69, 58.13, 53.62, 40.45, 39.60, 34.47, 30.93, 29.63, 27.28, 19.65, 18.61, 14.76. LCMS (ESI) m/z calcd for C36H39Cl2N7O8 [M + H]+ 768.22, found 768.07.
Synthesis of alkynyl-modified PEG2k prodrug
NH2-Val-Cit-PAB-Niclo (9.383 mg, 0.011 mmol) and Alk-PEG2k-COOH (22 mg, 0.011 mmol) were dissolved in 6 mL anhydrous DMF. Then, HOBT (2.2 mg, 0.0165 mmol), HBTU (6.3 mg, 0.0165 mmol) and DIPEA (4 μL, 0.022 mmol) were added into the DMF solution. The mixture was stirred at room temperature for 12 h. Then, PEG2k prodrug was purified using dialysis bag (molecular weight cutoff 2000). After lyophilization, PEG2k prodrug powder was collected and stored at 4 °C until use. 1H NMR (400 MHz, DMSO-d6) δ 10.71 (s, 1H), 9.98 (d, 1H), 8.72 (d, 1H), 8.34 (s, 1H), 8.29–8.26 (m, 1H), 8.18–8.13 (m, 1H), 8.02 (d, 1H), 7.96–7.93 (m, 1H), 7.89–7.86 (m, 1H), 7.78 (s, 1H), 7.69–7.62 (m, 3H), 7.48–7.45 (m, 3H), 6.00–5.97 (m, 1H), 5.42 (d, 4H), 4.40–4.37 (m, 2H), 4.25–4.17 (m, 1H), 4.15 (d, 1H), 3.86 (s, 2H), 3.51 (m, 212H), 3.06–3.01 (m, 1H), 2.98–2.91 (m, 1H), 2.86–2.73 (m, 2H), 2.39–2.34 (m, 1H), 2.03–1.96 (m, 1H), 1.74–1.69 (m, 1H), 1.63–1.56 (m, 1H), 1.49–1.36 (m, 4H), 0.88–0.81 (m, 6H).
Synthesis of P1-DPP co-polymer
BTP-2Br-OD (50.00 mg, 0.027 mmol), 2,5-Bis(6-bromohexyl)-3,6-bis(5-bromothiophen-2-yl)pyrrolo[3,4-c]pyrrole-1,4(2H, 5H)-dione (20.9 mg, 0.027 mmol), 4,7-bis(5-trimethylstannyl-2-thienyl)-2,1,3-benzothiadiazole (33.40 mg, 0.054 mmol), Tris(dibenzylideneacetone)dipalladium(0) (5 mg, 0.005 mmol) and Tri(o-tolyl)phosphine (10 mg, 0.0329 mmol) were dissolved in anhydrous chlorobenzene. The reaction mixture was stirred at 120 °C after being degassed by freeze-thaw cycles three times. The reaction was monitored by UV-Vis absorption. The product was precipitated in cold methanol, collected through centrifugation at 1000 × g, washed thrice with methanol and dried under vacuum to obtain P1-DPP co-polymer.
Niclosamide modified hypoxia-tolerance polymeric Type-I PDT photosensitizer (HTPSNiclo)
P1-DPP co-polymer modification with niclosamide: P1-DPP co-polymer (5 mg, 0.007 mmol), CuBr (5 mg, 0.035 mmol), N,N,N′,N′′,N′′-pentamethyldiethylenetriamine (36 μL, 0.17 mmol) were dissolved in 6 mL THF. Then, alkynyl modified mPEG2k (30 mg, 0.015 mmol) or alkynyl modified PEG2k prodrug (43.8 mg, 0.015 mmol) was added into the mixture. The mixture was stirred at room temperature for 48 h. Then, dialysis bag (molecular weight cutoff 3500) was used to purify the HTPS and HTPSNiclo.
Nilclosamide release measurement
The niclosamide release profile of PEG2k-Niclo and was measured after incubation with CatB for different times. For PEG2k-Niclo, the released profile was measured by HPLC directly. For HTPSNiclo, the samples were centrifuged at 1000 × g for 50 min using 3500 MWCO centrifugal concentrator (ThermoFisher). The released niclosamide in the bottom tubes was measured by HPLC.
Cell culture and in vitro cellular uptake assay
4T1 breast cancer cell line purchased from ATCC (American Type Culture Collection) were cultured in Dulbecco’s Modified Eagle Medium (DMEM) with 10% fetal bovine serum (FBS) and 1% antibiotics (penicillin and streptomycin) in an atmosphere containing 5% CO2 and 95% humidified air at 37 °C. 4T1 cancer cells were seeded into confocal cell culture dish at a density of 5 × 104 cells/dish and cultured overnight. Then the cells were treated with HTPS or HTPSNiclo ([BTP-OD] = 1 μg/mL) for 24 h for cellular uptake. The treated cells were washed with PBS and stained with Hoechst 33342 (NucBlue Live ReadyProbesReagent, Thermo Fisher) for nuclei staining and lysosome tracker green. Fluorescence images of the live cells were acquired on Laser Scanning Microscope LSM800 (Zeiss).
In vitro cytotoxicity assay
4T1 cancer cells were seeded in 96-well plates (6000 cells in 200 μL DMEM per well) for 24 h, and then 1 × PBS (pH = 7.4), HTPS or HTPSNiclo solutions of different concentrations were added to the culture medium. After incubation for 12 h, the cells were irradiated under 808 nm laser for 5 min (0.2 W/cm2). After incubation for another 24 h, the culture medium was removed, and cells were gently washed by fresh PBS. Fresh supplemented DMEM (100 μL per well) mixed with MTS (0.1 mg/mL, 20 μL per well) was then added to cells. After incubation for 3 h, the absorbance of the culture medium at 490 nm was recorded by SpectraMax M5 microplate/cuvette reader. Cell viability was calculated as the ratio of absorbance in different groups to that of control cells. Cells in the hypoxic group were cultured in hypoxic environments for 6 h and then treated with PDT treatment as introduced above.
Intracellular ROS detection
4T1 cancer cells were seeded in confocal cell culture dishes at a density of 1 × 105 cells/dish and incubated with HTPS or HTPSNiclo ([BTP-OD] = 5 μg/mL) for 24 h after adherence. H2DCFDA probe or DHR123 was then added to the cells and co-cultured for 30 min. The cells were irradiated with NIR irradiation for 5 min (808 nm, 0.2 W/cm2). Then the treated cells were washed with PBS for confocal imaging. Cells in the hypoxic group were cultured in hypoxic environments for 6 h and then treated with PDT treatment as introduced above.
In vitro DC maturation and M2 macrophages downregulation
BMDCs and BMDMs were isolated from the bone marrow of BALB/c mice according to established protocols. 4T1 cancer cells were first seeded in confocal cell culture dishes at a density of 5 × 104 cells/dish and then cultured overnight for cell adherence. Then the cells were incubated with HTPS or HTPSNiclo ([BTP-OD] = 5 μg/mL) for 24 h. Afterward, the cells were treated with NIR irradiation (808 nm, 0.2 W/cm) for 5 min. After incubation for another 24 h, the cell supernatants after different treatments were collected and added to the BMDCs or BMDMs culture dishes. After incubation for another 24 h, BMDCs or BMDMs was blocked with CD16/32 and stained with anti-CD11c, anti-CD80, and anti-CD86 antibodies or anti-F4/80 and anti-iNOS, for 30 min. After washing three times, the cells were analyzed by Fortessa X20 (BD Biosciences).
Western blotting
BMDCs were incubated with supernatants from 4T1 tumor cells for different groups, followed by gentle washing with cold PBS three times. The cells were then lysed using a cell lysis buffer, and the resulting supernatant was collected by centrifugation at 12,000 × g for 10 min at 4 °C. The protein in the supernatant was quantified using an enhanced BCA protein quantification kit. Equal volumes of the supernatant and loading buffer (1:1) were mixed and incubated in a 95 °C water bath for 30 min. The samples were then subjected to SDS-PAGE gel electrophoresis. After two hours, proteins were transferred from the gel to a PVDF membrane. After that, the membrane was incubated with 5% (w/v) non-fat milk in TBST for 2 h. The membrane was washed three times with TBST in a shaker after each step. Primary antibodies against p-STAT3, STAT3, or actin were added and incubated overnight at 4 °C. The membrane was subsequently incubated with horseradish peroxidase-conjugated secondary antibodies at 37 °C for 2 h. After three washes with TBST, western blotting images were captured using an ECL detection system. Uncropped and unprocessed scans of the western blot data are provided in the supporting information and source data.
Animal ethnics statement
Animal experiments were conducted in accordance with the guidelines for the care and use of laboratory animals set by the Nanyang Technological University-Institutional Animal Care and Use Committee (NTU-IACUC) and were approved by the Institutional Animal Care and Use Committee (IACUC) of Singapore, under protocol number A21078. Mice were housed in groups in ventilated, transparent plastic cages, maintained at a controlled ambient temperature (~22 °C) and humidity (50%), with a standard 12-h light/dark cycle. Sex was not a factor in the study design.
Establishment of mouse tumor model
Five-week-old female BALB/c mice were purchased from InVivos (Singapore). For tumor inoculation, 4T1 cancer cell suspension in PBS were subcutaneously inoculated at the flank region of right hind legs of each mouse (1 × 106 cells). The mice were used for imaging and therapy when tumor size reached about 100 mm3 unless otherwise stated. The experimental endpoints were defined when the tumor size exceeded 1500 mm3.
In vivo fluorescence imaging of 4T1 tumor-bearing mice
Real-time in vivo fluorescence imaging was conducted in 4T1 tumor-bearing mice at t = 0, 1, 4, 12, 24, 36, 48, 72, 96 h after intravenous administration of HTPS or HTPSNiclo (200 µL, [BTP-OD] = 200 μg/mL). Fluorescence images of the mice were conducted on an IVIS spectrum CT system with excitation at 570 nm and emission at 780 nm. At 96 h post-injection time point, the mice for in vivo fluorescence imaging were euthanized and the resected organs were imaged for ex vivo biodistribution analysis. Fluorescence intensity of the tumor and the resected organs from the mice were quantified by the ROI analysis using Living Image 4.3 software.
In vivo evaluation for HTPSNiclo-mediated photo-immunotherapy
4T1 cancer cell suspension in PBS were subcutaneously inoculated at flank region on -7 d (1 × 106 cells, primary tumor) and -1 d (2 × 105 cells, distant tumor). The mice were systematically administrated with 200 µL PBS, HTPS or HTPSNiclo (200 µL, [BTP-OD] = 200 μg/mL). At 1 d, 3 d, and 5 d post-injection, primary tumors in HTPS or HTPSNiclo with NIR irradiation group were irradiated for 5 min (808 nm, 0.2 W/cm2). The body weights and tumor sizes were monitored every 2 days for 20 days. The equation for tumor size calculation is volume = a × b2/2, where a represents tumor length and b represents tumor width. After 20 days, the mice in each group were monitored for 40 days to record survival rates.
In vivo generation of O2 •-
Mice with established tumors (~100 mm3) were intravenously administered with 200 µL PBS, HTPS or HTPSNiclo (200 µL, [BTP-OD] = 200 μg/mL). At 24 h post-injection, 30 µL DHR123 (50 µM) was intratumorally injected into the tumors of the mice in each group. After 30 min, the tumors of the mice were irradiated with NIR for 5 min (808 nm, 0.8 W/cm2). The mice were then euthanized, and the tumors were collected, fixed in 4% paraformaldehyde (PFA), and cryo-sectioned. The tumor sections were stained with Hoechst 33342 and imaged using LSM800 confocal laser scanning microscope. The fluorescent intensity of DHR123 was quantified to evaluate in vivo O2•- generation in each group using the image processing software ImageJ.
DC maturation and cytokine production
Three days after different treatments, the mice in each group were euthanized and their blood and tumor draining lymph nodes (TDLNs) were collected. Single cell suspension was obtained by directly grounding and straining TDLNs using a 70-µm cell strainer. Cells were then incubated with anti-CD16/32 antibody and stained with anti-CD11c, anti-CD80, and anti-CD86 antibodies for flow cytometric analysis. The blood serum was collected from the supernatant of the blood samples for cytokine detection (IL-12, IFN-γ, TNF-α, and IFN-β) by ELISA according to manufacturer’s protocols.
Flow cytometric analysis of immune infiltrate phenotyping
Seven days after different treatments, the mice in each group were euthanized, and tumors and spleens were collected. Single cell suspensions from tumor and spleens were first blocked with anti-CD16/32 antibody, followed by stained with anti-CD45, anti-CD3, anti-CD4, and anti-CD8a antibodies. For flow cytometric analysis of Tregs, the cells were stained with anti-CD45, anti-CD3, anti-CD4, followed by fixation in 0.5 mL/tube Fixation and Perm Buffer and incubated with PE anti-mouse FOXP3. After washing three times, the cells were analyzed by Fortessa X20 (BD Biosciences).
Immunofluorescence staining of tumor sections
According to the standard protocol, tumors were fixed in 4% PFA and cryo-sectioned with a thickness of 10 µm. The tumor sections were incubated with anti-caspase 3, HIF-1 and CD31 at 4 °C followed by staining the nuclei with Hoechst 33342. The fluorescent images were captured using LSM800 confocal laser scanning microscope. The quantification of the fluorescent intensities was performed using software ImageJ.
Histological studies and metastasis analysis
The resected organs from the 4T1 tumor-bearing mice in each group were fixed with 4% PFA, dehydrated with 30% sucrose and sectioned into thin slices (10 μm).
Rechallenged tumor model
Mice bearing 4T1 tumors were systematically administrated with HTPS or HTPSNiclo (200 µL, [BTP-OD] = 200 μg/mL). At 1 d, 3 d, and 5 d post-injection, tumors in HTPS or HTPSNiclo + NIR group were irradiated with NIR for 5 min (808 nm, 0.8 W/cm2). On day 33, mice were re-inoculated with 1 × 106 4T1 cells in PBS on the left flank without any further treatment. Tumor size was monitored every two days for two weeks. Spleen from 4T1 tumor-bearing mice at 57 d were collected for population analysis of memory T cells using flow cytometry. TDLNs were directly ground and strained with a 70-µm cell strainer to obtain single cell suspension. Then, the cells were blocked with anti-CD16/32 antibody for 10 min, followed by stained with FITC anti-CD3, PE anti-CD8, APC anti-CD62L, and BV605 anti-CD44 antibodies. After washed thrice with cell staining buffer, the cells were analyzed by Fortessa X20 (BD Biosciences).
Statistical analysis
Technical replicates were used in all experiments, and data were calculated and expressed as mean ± standard deviation (S.D.) unless stated otherwise. Investigators were blinded to group allocation during experiments. The specific statistical methods are indicated in the figure legends. Statistical differences between two groups were tested with a two-tailed student’s t-test and more than three groups were determined by one-way analysis of variance followed by Tukey’s post hoc test. The p values of survival curves were analyzed using the Kaplan–Meier method via the log-rank test. For all tests, P values less than 0.05 were considered statistically significant. *p < 0.05, ** p < 0.01, *** p < 0.001 and **** p < 0.0001. All statistical calculations were performed using GraphPad Prism 8.0, including assumptions of tests used.
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Data availability
All data needed to evaluate the conclusions in the paper are presented in the paper and/or the Supplementary Materials. Source data are provided with this paper. All other data are available from the corresponding authors upon request. Source data are provided with this paper.
References
Nam, J. et al. Cancer nanomedicine for combination cancer immunotherapy. Nat. Rev. Mater. 4, 398–414 (2019).
Anderson, N. M. & Simon, M. C. The tumor microenvironment. Curr. Biol. 30, R921–R925 (2020).
Li, X., Lovell, J. F., Yoon, J. & Chen, X. Clinical development and potential of photothermal and photodynamic therapies for cancer. Nat. Rev. Clin. Oncol. 17, 657–674 (2020).
Dolmans, D., Fukumura, D. & Jain, R. K. Photodynamic therapy for cancer. Nat. Rev. Cancer 3, 380–387 (2003).
Agostinis, P. et al. Photodynamic therapy of cancer: an update. C. A. Cancer J. Clin. 61, 250–281 (2011).
Yu, J. et al. Synergistic enhancement of immunological responses triggered by hyperthermia sensitive Pt NPs via NIR laser to inhibit cancer relapse and metastasis. Bioact. Mater. 7, 389–400 (2022).
Galluzzi, L., Buque, A., Kepp, O., Zitvogel, L. & Kroemer, G. Immunogenic cell death in cancer and infectious disease. Nat. Rev. Immunol. 17, 97–111 (2017).
Obeid, M. et al. Calreticulin exposure dictates the immunogenicity of cancer cell death. Nat. Med. 13, 54–61 (2007).
Yue, J. et al. Light-triggered multifunctional nanoplatform for efficient cancer photo-immunotherapy. J. Nanobiotechnol. 20, 181 (2022).
Shen, L. et al. Recent progress in tumor photodynamic immunotherapy. Chin. Chem. Lett. 31, 1709–1716 (2020).
Guo, R. et al. Engineered nanomaterials for synergistic photo-immunotherapy. Biomaterials 282, 121425 (2022).
Ferrari, M. Cancer nanotechnology: opportunities and challenges. Nat. Rev. Cancer 5, 161–171 (2005).
Wang, D. et al. Acid-activatable versatile micelleplexes for PD-L1 Blockade-enhanced cancer photodynamic immunotherapy. Nano Lett. 16, 5503–5513 (2016).
Yi, Y. et al. Vaccine-like nanomedicine for cancer immunotherapy. J. Controlled Release 355, 760–778 (2023).
Hashimoto, M. et al. PD-1 combination therapy with IL-2 modifies CD8+ T cell exhaustion program. Nature 610, 173–181 (2022).
Warszyńska, M., Repetowski, P. & Dąbrowski, J. M. Photodynamic therapy combined with immunotherapy: Recent advances and future research directions. Coord. Chem. Rev. 495, 215350 (2023).
He, S. et al. A semiconducting iron-chelating nano-immunomodulator for specific and sensitized sono-metallo-immunotherapy of cancer. Angew. Chem. Int. Ed. 62, e202310178 (2023).
Lucky, S. S., Soo, K. C. & Zhang, Y. Nanoparticles in photodynamic therapy. Chem. Rev. 115, 1990–2042 (2015).
Xie, Z. et al. Emerging combination strategies with phototherapy in cancer nanomedicine. Chem. Soc. Rev. 49, 8065–8087 (2020).
Zhou, Z., Song, J., Nie, L. & Chen, X. Reactive oxygen species generating systems meeting challenges of photodynamic cancer therapy. Chem. Soc. Rev. 45, 6597–6626 (2016).
Chen, D. et al. A highly-efficient type I photosensitizer with robust vascular-disruption activity for hypoxic-and-metastatic tumor specific photodynamic therapy. Small 16, e2001059 (2020).
Zhang, M., Zhao, Y., Ma, H., Sun, Y. & Cao, J. How to improve photodynamic therapy-induced antitumor immunity for cancer treatment? Theranostics 12, 4629–4655 (2022).
Mroz, P. & Hamblin, M. R. The immunosuppressive side of PDT. Photochem. Photobiol. Sci. 10, 751–758 (2011).
Perillo, B. et al. ROS in cancer therapy: the bright side of the moon. Exp. Mol. Med. 52, 192–203 (2020).
Cui, D. et al. A semiconducting polymer nano-prodrug for hypoxia-activated photodynamic cancer therapy. Angew. Chem. Int. Ed. 58, 5920–5924 (2019).
Taylor, C. T. & Scholz, C. C. The effect of HIF on metabolism and immunity. Nat. Rev. Nephrol. 18, 573–587 (2022).
Anderson, K. G., Stromnes, I. M. & Greenberg, P. D. Obstacles posed by the tumor microenvironment to T cell activity: a case for synergistic therapies. Cancer Cell 31, 311–325 (2017).
Chen, D. et al. Type I photosensitizers revitalizing photodynamic oncotherapy. Small 17, e2006742 (2021).
Li, L. et al. An NIR-II-Emissive photosensitizer for hypoxia-tolerant photodynamic theranostics. Adv. Mater. 32, e2003471 (2020).
Li, M. et al. Unimolecular Photodynamic O(2)-Economizer to overcome hypoxia resistance in phototherapeutics. J. Am. Chem. Soc. 142, 5380–5388 (2020).
Li, Y., Zhang, W., Niu, J. & Chen, Y. Mechanism of photogenerated reactive oxygen species and correlation with the antibacterial properties of engineered metal-oxide nanoparticles. ACS Nano 6, 5164–5173 (2012).
Li, J., Rao, J. & Pu, K. Recent progress on semiconducting polymer nanoparticles for molecular imaging and cancer phototherapy. Biomaterials 155, 217–235 (2018).
Jiang, Y. & Pu, K. Multimodal biophotonics of semiconducting polymer nanoparticles. Acc. Chem. Res. 51, 1840–1849 (2018).
Ding, L. et al. Polymer semiconductors: synthesis, processing, and applications. Chem. Rev. 123, 7421–7497 (2023).
Tang, D. et al. NIR-II light accelerated prodrug reduction of Pt(IV)-Incorporating Pseudo Semiconducting Polymers for robust degradation and maximized photothermal/chemo-immunotherapy. Adv. Mater. 35, 2300048 (2023).
Lyu, Y. et al. A photolabile semiconducting polymer nanotransducer for near-infrared regulation of CRISPR/Cas9 Gene Editing. Angew. Chem. Int. Ed. 58, 18197–18201 (2019).
Zhang C. et al. Eosinophil-activating semiconducting polymer nanoparticles for cancer photo-immunotherapy. Angew. Chem. Int. Ed. e202405358. https://doi.org/10.1002/anie.202405358 (2024).
Zhang, C. et al. Checkpoint nano-PROTACs for activatable cancer photo-immunotherapy. Adv. Mater. 35, e2208553 (2023).
Qi, J. et al. Semiconducting polymer nanoparticles with surface-mimicking protein secondary structure as lysosome-targeting chimaeras for self-synergistic cancer immunotherapy. Adv. Mater. 34, 2203309 (2022).
Zhang, C. et al. Catalytical nano-immunocomplexes for remote-controlled sono-metabolic checkpoint trimodal cancer therapy. Nat. Commun. 13, 3468 (2022).
Yu, J. et al. Polymeric STING pro-agonists for tumor-specific sonodynamic immunotherapy. Angew. Chem. Int. Ed. 62, e202307272 (2023).
Li, J. et al. Photoactivatable organic semiconducting pro-nanoenzymes. J. Am. Chem. Soc. 141, 4073–4079 (2019).
Zeng, Z. et al. Activatable polymer nanoenzymes for photodynamic immunometabolic cancer therapy. Adv. Mater. 33, 2007247 (2021).
He, S., Liu, J., Zhang, C., Wang, J. & Pu, K. Semiconducting polymer nano-regulators with cascading activation for photodynamic cancer immunotherapy. Angew. Chem. Int. Ed. 61, e202116669 (2022).
Ding, M. et al. Augmenting immunogenic cell death and alleviating myeloid-derived suppressor cells by sono-activatable semiconducting polymer nanopartners for immunotherapy. Adv. Mater. 35, 2302508 (2023).
Cui, C. et al. Enhancing electron transfer of a semiconducting polymer for type I photodynamic and photothermal synergistic therapy. Front. Bioeng. Biotechnol. 10, 1004921 (2022).
Wen, K. et al. Achieving Efficient NIR-II Type-I photosensitizers for photodynamic/photothermal therapy upon regulating chalcogen elements. Adv. Mater. 34, e2108146 (2022).
Niu, G. et al. Signal transducer and activator of transcription 3 is required for hypoxia-inducible factor-1alpha RNA expression in both tumor cells and tumor-associated myeloid cells. Mol. Cancer Res. 6, 1099–1105 (2008).
Xu, C. et al. Nanoparticles with ultrasound-induced afterglow luminescence for tumour-specific theranostics. Nat. Biomed. Eng. 7, 298–312 (2023).
Yu, L. et al. Photocatalytic superoxide radical generator that induces pyroptosis in cancer cells. J. Am. Chem. Soc. 144, 11326–11337 (2022).
Zhang, Z. et al. Cathepsin B mediates the lysosomal-mitochondrial apoptosis pathway in arsenic-induced microglial cell injury. Hum. Exp. Toxicol. 42, 9603271231172724 (2023).
Zhang, C. et al. A polymeric extracellular matrix nanoremodeler for activatable cancer photo-immunotherapy. Angew. Chem. Int. Ed. 2, e202217339 (2023).
Chen, Y. et al. Achieving high-performance ternary organic solar cells through tuning acceptor alloy. Adv. Mater. 29, 1603154 (2017).
Acknowledgements
Y.Z. thanks National Natural Science Foundation of China (22322406) for financial support. K.P. thanks Singapore National Research Foundation (NRF) (NRF-NRFI07-2021-0005) and the Singapore Ministry of Education Academic Research Fund Tier 2 (MOE-T2EP30220-0010, MOE-T2EP30221-0004) for financial support. This article was also supported by china postdoctoral science foundation under grant number GZC20240525 (J.Y.).
Author information
Authors and Affiliations
Contributions
K.P. and J.Y. conceived the concept and designed the research. J.Y. performed material preparation and in vivo experiments. K.P., Y.Z., and J.Y. contributed to the interpretation of the results. K.P., Y.Z., J.Y., J.W., J.H., C.X., M.X., and C.K. contributed to the writing of this manuscript.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Dan Ding, Quanyin Hu and Jingchao LI for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Yu, J., Wu, J., Huang, J. et al. Hypoxia-tolerant polymeric photosensitizer prodrug for cancer photo-immunotherapy. Nat Commun 16, 153 (2025). https://doi.org/10.1038/s41467-024-55529-8
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-024-55529-8
This article is cited by
-
A Bromobenzoyl Pyridine Photosensitizer with Suppressed Phosphorescence for Efficient ROS-Based Photodynamic Therapy
Journal of Fluorescence (2025)