Direct and indirect anticancer effects of hyperthermic intraperitoneal chemotherapy on peritoneal malignancies (Review)

  • Authors:
    • Ying Zhang
    • Yumin Wu
    • Jun Wu
    • Chen Wu
  • View Affiliations

  • Published online on: February 10, 2021     https://doi.org/10.3892/or.2021.7974
  • Article Number: 23
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

The successful application of hyperthermic intraperitoneal chemotherapy (HIPEC) illustrates its antitumor activity against primary malignances and peritoneal metastases. Although the specific underlying molecular mechanisms remain unclear, increasing evidence suggest that HIPEC directly and indirectly inhibits tumor growth, and prolongs overall survival in both hyperthermic and chemotherapeutic manners. To demonstrate the superiority and limitations of such a therapeutic regimen, the present review focuses on the biological and immunological anticancer mechanisms of HIPEC. In addition, the potential combination of HIPEC with other therapies is discussed, as well as its potential to prolong the overall survival time of patients with peritoneal malignancies.

Introduction

According to statistics, peritoneal metastasis of common malignancies is considered unresectable and poses a great challenge in cancer treatment (1). Previously, traditional intravenous chemotherapy was the recommended option for patients with peritoneal metastasis, the efficacy of which is largely limited by myelotoxicity (2,3). The peritoneal plasma barrier and high interstitial pressure of tumor tissues limit the accumulation of traditional intravenous agents in the peritoneal cavity (4). Several clinical trials have confirmed that intraperitoneal chemotherapy can overcome these limitations and extend patient survival time (5). Intraperitoneal drug delivery under hyperthermia conditions, a procedure known as hyperthermic intraperitoneal chemotherapy (HIPEC), has proven to be a more effective interventional therapy for metastatic tumors of the abdominal cavity (6,7). Since the standard nomenclature for HIPEC was devised at the Fourth International Workshop on Peritoneal Surface Malignancy in 2004 (8), its clinical application has drawn great interest. However, as demonstrated in the PRODIGE7 (9) and PROPHYLOCHIP (10) trials, adjuvant HIPEC does not appear to significantly improve patient survival. As an adjuvant therapeutic method, the underlying molecular mechanisms of HIPEC remain largely unknown, which may affect its clinical application. The present review focuses on the direct and indirect effects of HIPEC, particularly at the immunological level.

Clinical applications of HIPEC

HIPEC is an important adjuvant treatment for malignant tumors, malignant ascites and intraperitoneal metastases, the aim of which is to perfuse chemotherapeutic agents into the abdominal cavity at a constant temperature and for a specific period (7,11). Giovanella et al (12) revealed that exposure to temperatures at 42.5–43.0°C has a significantly greater lethal effect on neoplastic cells compared with non-neoplastic human cells. There are two clinical methods of HIPEC application, namely the open-abdomen and closed-abdomen techniques (13). The open method is usually performed in the operating room directly after surgery and has exhibited significant heat loss in porcine models (14). A meta-analysis revealed that closed HIPEC is used more frequently than the open method, and that the choice of HIPEC method has no impact on the overall recurrence-free survival rate of patients, though this finding requires further verification (15). CO2-HIPEC is a recently developed technology that promises improved heat delivery into the ‘closed’ abdomen (14). The closed CO2-HIPEC system (PRS-1.0 Combat) uses a CO2 recirculation system to distribute the perfusate and increase pressure within the peritoneal cavity, ensuring intra-abdominal thermal homogeneity, in addition to optimal solution distribution and drug penetration (16). Pressurized Intraperitoneal Aerosol Chemotherapy (PIPAC) employs the CO2 recirculation method, using a PIPAC micropump to nebulize the liquid cytotoxic drug into droplets of ~25-µm, generating a polydisperse aerosol in the abdominal cavity that is more effectively absorbed (17,18).

HIPEC has been used to treat colorectal cancer (10,19,20), gastric cancer (21,22), ovarian cancer (7,23), appendiceal neoplasms (24) and other peritoneal metastatic carcinomas (25). van Driel et al (7) discovered that among patients with stage III epithelial ovarian cancer undergoing interval debulking surgery, the median overall survival (OS) time of the intravenous- and IP-therapy groups was 49.7 and 65.6 months, respectively (P=0.03). However, in 2003, Verwaal et al (19) reported the results of a randomized trial, suggesting that cytoreduction and HIPEC improve the survival times of patients with peritoneal carcinomatosis of colorectal origin. In the 2019 COLOPEC randomized multi-center trial, Klaver et al (20) indicated that adjuvant HIPEC did not improve the peritoneal metastasis-free survival times of patients with T4 or perforated colon cancer 18 months post-treatment. In addition, the PRODIGE 7 trial was unable to demonstrate improved OS or relapse-free survival time in patients with peritoneal metastases of colorectal cancer who had accepted oxaliplatin-HIPEC treatment following cytoreductive surgery (CRS) (9). In the PRODIGE 7 trial, median OS time in the HIPEC and CRS-only arms was 41.7 and 41.2 months, respectively. In addition, the rate of grade ≥3 complications in the HIPEC arm was higher compared with the non-HIPEC arm (24 vs. 14%) at 60 days. The PROPHYLOCHIP-PRODIGE15 trial also revealed that adjuvant oxaliplatin-HIPEC did not improve patient disease-free survival time compared with standard surveillance in 2020 (10).

The clinical effects of HIPEC remain a contradiction. Although HIPEC has been demonstrated to increase intraperitoneal hydrostatic pressure to enhance drug delivery to the peritoneal surface, and theoretically, to eliminate residual microscopic peritoneal disease, the practice remains controversial (26,27). Thus, investigations on the basic molecular mechanisms of HIPEC may prove beneficial. HIPEC is also known to exert direct and indirect antitumor effects by promoting albuminous degeneration, inducing apoptosis and inhibiting angiogenesis, which may indicate a synergistic effect between hyperthermia and chemotherapy (28,29).

Direct hyperthermic injury to tumor cells

Cell membrane and cytoskeleton

During hyperthermia, supranormal temperatures can induce a distinct decrease in membrane and cytoplasmatic proteins, resulting in tumor cell death (30,31). For example, multidrug resistance-associated protein 1 and 3 are responsible for the failure of several oncological treatments, such as chemotherapy of taxanes, vinca alkaloids and anthracyclines, while the expression of these proteins can be reduced by hyperthermia (32). The expression levels of membrane-bound cytoskeletal proteins, such as actin, are also reduced by hyperthermia, which results in the subsequent loss of integrin CD11a (also known as leukocyte function-associated antigen-1) from the cell surface (33,34). CD11a deficiency or blockade subsequently abrogates the aggregation of Treg cells, enhancing the efficacy of anticancer treatments (35). Furthermore, defects in the tumor cell membrane result in enhanced membrane fluidity (36). As a result, hyperthermia-induced membrane remodeling and the release of lipid signals can upregulate cellular thermosensitivity (37). In addition, activation of the purinergic receptor, P2X7, can be potentiated by the associated changes in membrane fluidity, which ultimately promotes tumor cell death (38). Apoptosis is one of the mechanisms underlying hyperthermia-associated cell death, which occurs alongside downregulation of p53 and Bcl-2 expression, as well as upregulation of Bax expression and mitotic catastrophe (39,40).

DNA damage

Early trials have reported that hyperthermia can result in nuclear DNA double-strand breaks (DSBs), single strand breaks and chromosomal aberrations in tumor cells (41,42). Through the enhancement of ataxia-telangiectasia mutated protein (ATM) kinase activity, and an increase in cellular ATM autophosphorylation, foci formation of phosphorylated H2AX (at serine 139; γ-H2AX) can be induced by hyperthermia, which acts as a specific indicator for the occurrence of DNA DSBs (43,44). While DSBs are primarily repaired via non-homologous end joining, hyperthermia increases the probability of incorrect DSB reconnections (45). However, hyperthermia appears to delay the repair of DNA DSBs caused by cisplatin, doxorubicin or radiotherapy, rather than by direct proteasomal DNA damage (46,47). Mild hyperthermia (41.0–42.5°C) has also been reported to induce BRCA2 degradation and inhibit homologous recombination, thus impeding DNA damage repair (41,48). Cisplatin-induced adduct formation is also increased in vivo, which is an important determinant of toxicity (49).

Indirect effects of hyperthermia

Enhancement of chemotherapeutic effects

In 1994, researchers determined that hyperthermia enhances the effects of intraperitoneal chemotherapy by increasing the uptake of chemotherapeutic agents (50). As a result, the clinically effective dose could be reduced to decrease the incidence of severe side effects (51). Hyperthermia has been demonstrated to enhance the sensitivity of tumors to chemotherapy by the impairing DNA repair (39), and to decrease the number of proliferating tumor cells when combined with chemotherapeutic drugs (52). However, Cesna et al (28) suggested that without cisplatin, hyperthermia does not synergistically effect cancer cell viability, indicating that the interaction between chemotherapy and hyperthermia requires further investigation.

Modulation of inflammation and the immune response

Due to the clinical similarity between exogenously induced hyperthermia and natural fever, it has been speculated that immune cells, including antigen presenting cells (APCs) and natural killer (NK) cells, can be stimulated by hyperthermia to enhance the antitumor effects of chemotherapeutic agents in the abdominal cavity (53). The premise of these immunological influences is that peptides released from dead tumor cells can be internalized by APCs, including dendritic cells (DCs) and macrophages, which subsequently activate cytotoxic T lymphocytes (CTLs) (54,55). In this regard, hyperthermia can substantially enhance the phagocytic potential of DCs and increase DC infiltration by regulating the (C-C) receptor (CCR)7-(C-X-C motif) ligand 21 axis (56). The viability of NK cells is substantially reduced between 41–42°C, while thermal stress activates NK cell cytotoxicity by activating receptor NKG2D and its ligand, MHC class I-related chain A (57).

The peritoneum is a visceral tissue with unique immune functions, in which the largest cell fraction are the peritoneum mesothelial cells (HPMCs) (58). The HPMCs comprise fibroblasts, endothelial cells, macrophages and lymphocytes (59). Tumor-associated macrophages are the primary immune cell population in the tumor peritoneal cavity and can be stimulated to differentiate into two distinct subtypes: Anti-tumorigenic M1 macrophages and pro-tumorigenic M2 macrophages, of which M2 macrophages represent the majority (60,61). M2 macrophages inactivate CD8+ T cells through increased expression of programmed cell death 1 ligand 1 and cytotoxic T lymphocyte antigen 4 (62). In addition, M2 macrophages accelerate tumor cell proliferation through signal transducer and activator of transcription 3 (STAT3) activation via transforming growth factor β (TGF-β), interleukin (IL)-6 and IL-10 (63,65).

A study suggested that fever-like mild heat (~43°C) can synergistically induce macrophage polarization from the M2 to M1 phenotype during low-temperature photothermal therapy with a lipid nanocomposite (66). However, to the best of our knowledge, no studies have reported changes in the immunological competency of the peritoneum in patients undergoing HIPEC, which may indicate a determining factor for HIPEC resistance, and thus requires further investigation. In addition, hyperthermia may impair the functions and decrease the number of immune cells (67,68), thus future studies should also focus on the dual effects of hyperthermia.

Fever is commonly considered to be a protective response following inflammation (69). Artificial HIPEC-induced fever appears to evoke an inflammatory response with the release of cytokines, such as IL-1, IL-6 and TGF-β (70). IL-1 is expressed by malignant and infiltrating cell, and promotes tumor progression and invasiveness (71). By activating STAT3, classical IL-6 signaling blocks DC maturation, inhibits T-cell activation and enhances tumor cell proliferation (72). In addition, IL-6 and TGF-β are considered to promote the differentiation of Th17 cells, which support tumor progression by secreting immunosuppressive IL-17, thus facilitating immune escape (73). Conflictingly, hyperthermia phospho-regulates gp130, which is anchored to the endothelial cell membrane of tumor microvessels (74). Soluble IL-6 receptor α binds IL-6 and gp130 to activate IL-6 trans-signaling, which regulates high endothelial venule adhesion and promotes the trafficking and recruitment of CD8+ T cell into the tumor site (75).

Regulation of the heat shock response

The expression of molecular chaperone heat-shock proteins (HSPs) in the abdominal cavity (including HSP-70, HSP-72 and HSP-92) can be upregulated by hyperthermic antineoplastic agents, which protect the organs from heat-induced stress (76). Similar to other multidomain proteins, HSPs recognize and bind unfolded/disordered sequences to facilitate the folding/refolding of these sequences, or simply to present themselves to the proteasome for destruction, protecting intracellular proteins from stress-induced cellular damage (77,78). The upregulation of HSP machinery in cancerous cells may prevent the misfolding and degradation of mutated and overexpressed oncoproteins (79). The highly conserved molecular mechanisms of HSPs hinder the antiproliferative and apoptotic effects of HIPEC on tumor cells (80). For example, upregulation of HSP27 resulting from HIPEC/CRS combination treatment may promote thermotolerance and chemoresistance (81). However, chemotherapy can also suppress HSP27 expression, rendering cancer cells sensitive to mild hyperthermia (43°C) (82).

Zunion et al (82) have suggested that following HIPEC, HSP90 induces an anticancer immune response at the cell surface. As with the HSP90 inhibitor, 17-allylamino geldanamycin, hyperthermia causes drugs to accumulate in the tumor, leading to potent antitumor activity (83). The increased delivery of HSP-targeted chemotherapeutics can be attributed to the hyperthermia-induced expression of cell surface HSP glucose regulated protein 78 kDa, which targets N-(2-hydroxypropyl) methacrylamide copolymer-drug conjugates (84).

Professional APCs of the innate immune response can be stimulated by extracellular HSPs, followed by cytokine release and the expression of cell surface molecules (85). In addition, acting as a tumor vaccine, the cross-presentation of HSP-bound peptide antigens to MHC class I molecules can stimulate adaptive immunity via DCs (86,87), leading to the efficient induction of antigen-specific CTLs (88,89). There is also evidence that macrophages can be activated by heat shock factor-1 through upregulation of the inducible nitric oxide synthase gene (90).

Role of chemotherapy in HIPEC

Pharmacokinetics of chemotherapeutics in the peritoneal cavity

The peritoneal cavity is a closed space that consists of the peritoneum, abdominal organs and 50–70 ml peritoneal fluid (91). The peritoneum is an extensive serosal exchange membrane comprised of a single layer of squamous HPMCs (~25 µm in diameter), collagen, adipose tissue, lymphocytes, blood vessels and lymphatics (4). The so-called peritoneal-plasma barrier comprises HPMCs, the subserosal interstitium and the capillary walls (92). As the primary absorption barrier, large molecular drugs in the peritoneal cavity are absorbed slowly into the systemic blood circulation (93). The area under the concentration-time curve (AUC) of drugs from the peritoneal cavity to the plasma demonstrated the pharmacological advantage of intraperitoneal administration (94).

Table I outlines the pharmacokinetic properties of commonly used chemotherapeutics following intraperitoneal administration (95101). As presented in Table I, a high peritoneal AUC reflects high local exposure and potential drug efficacy, while a low plasma AUC indicates low systemic exposure and reduced systemic toxicity (102). The dose of the drug used in HIPEC is another factor affecting intraperitoneal administration (103). There are usually two primary methods of dose determination, including body surface area-based and concentration-based dose calculations (104). Lemoine et al (105) compared these methods using oxaliplatin-based HIPEC and found no differences in pharmacokinetic advantage between the two models. Concentration-based dose determination appeared to result in a higher drug concentration in the tumor nodule, whilst imparting higher toxicity and efficacy (106). However, due to compromised lymphatic function, leaky vasculature and the dense structure of the extracellular matrix, the penetration depth of these drugs into the tumor nodule was limited to several millimeters (100). Previous studies have suggested that the combination of antiangiogenic therapy and intraperitoneal chemotherapy improves drug penetration by decreasing interstitial fluid pressure (107109). Drugs with nano-sized particles, such as paclitaxel, have also been used in intraperitoneal chemotherapy to increase exposure time in the peritoneal cavity, and thus improve intratumoral drug accumulation (110). It is well-known that CO2-HIPEC enhances drug penetration in the tumor (18). Shamsi et al (111) reported a novel magnetically assisted strategy, using drug-coated magnetic nanoparticles and a permanent external magnet to increase the final concentration in tumor nodules (radius, 5–10 mm).

Table I.

Pharmacokinetic characteristics of commonly used chemotherapeutics following hyperthermic intraperitoneal chemotherapy administration.

Table I.

Pharmacokinetic characteristics of commonly used chemotherapeutics following hyperthermic intraperitoneal chemotherapy administration.

DrugMolecular weight, DaltonsIntraperitoneal Dose, mg/m2Carrier solventsMean AUC, peritonealMean AUC, plasmaAUC ratioDrug penetration distanceRefs.
Mitomycin C334.335.0Dialysis fluid630±130, mg • min/l71±24, mg • min/l10.10± 4.60<5 mm(94)
Cisplatin300.150.0Dialysis fluid2.87±0.52, mM • min0.46±0.05, mM • min6.281–3 mm(95)
Oxaliplatin397.3460.05% Dextrose2,412.9±711.4, mg • min/l138.1±33.1, mg • min/l18.601–2 mm(96)
Paclitaxel853.9175.0Normal saline736.8±305.2, mg • h/l2.284±1.185, mg • h/l387.00±260.0080 cell layers(97)
Docetaxel861.975.0Normal saline85.10, mg • h/l0.7634, mg • h/l207.40NA(98)
Mitoxantrone517.428.0Normal saline15,530±2,471, ng • h/l1,036±394, ng • h/l12.505–6 cell layers(99)
Doxorubicin543.560.0–75.0Normal saline372.0±260.0, mg • min/l4.1±6.0, mg • min/l162.00 ±113.004–6 cell layers(100)

[i] AUC, area under the concentration-time curve; NA, not available.

As for the selection of intraperitoneal chemotherapeutics for solid tumors, several factors must be considered in addition to the selection of effective proliferation inhibitors, such as the active form and half-life of agents (112). The effects of peritoneal delivery are based on direct contact between tumor cells and the chemotherapeutic agent (113). As these agents must be delivered in an active form, chemotherapeutics that require activation by hepatic metabolization are considered unsuitable (113).

Immunological effects of chemotherapy

Increasing evidence suggest that several immunological factors can be induced by conventional chemotherapeutic agents, including the composition, phenotype and function of immune cells, as well as alterations in several immune-related parameters (114,115). For example, Latchman et al (116) revealed that the expression of T cell inhibitory molecule programmed death receptor-ligand 2 (PD-L2) on both human DCs and tumor cells is markedly decreased following exposure to platinum-based chemotherapeutics. As a second ligand for PD-1, PD-L2 inhibits T cell activation (117). In addition, Treg cells and circulating myeloid-derived suppressor cells have been demonstrated to be depleted by paclitaxel, gemcitabine and vinorelbine, therapeutically enabling relevant tumor-targeting immune responses (118120). As for the immunological effects of chemotherapeutics, immune effector cells may be stimulated, and Treg cells may be depleted following the release of cytokines and chemokines, while tumor-specific antigens containing tumor cell peptides may be internalized by APCs following chemotherapeutic exposure (121).

Chemotherapy-induced immunological cell death (ICD) (Fig. 1) may enhance the immunological effects of treatment (122,123). Following hyperthermia- and chemotherapy-associated tumor cell death, damage-associated molecular patterns (DAMPs) are released, which include surface-exposed calreticulin, secreted ATP and IFN, and high-mobility group protein B1 (130). By binding to pattern recognition receptors, DAMPs can elicit innate and/or adaptive immune responses (124). During anticancer immune responses, primary effector cells, such as DCs, internalize tumor-associated antigens and mature in the surrounding tumor tissues, resulting in the production of immune-activating cytokines (125). T cells are then recruited, which destroy tumor cells. Notably, this phenomenon was not observed in immunodeficient mice (126). Accordingly, CTLs are one of the most effective means to combat chemotherapeutic resistance, serving a key role in cancer treatment (54). Thus, it was hypothesized that this ‘passive immunotherapy’ may be an important adjunct strategy for overcoming chemotherapeutic resistance (127). In addition, clarifying the molecular mechanisms of ICD and applying them to the development of a cancer-specific vaccines during HIPEC can potentially improve chemotherapeutic resistance and eliminate residual tumor cells (122).

Cyclic GMP-AMP synthase (cGAS)-cyclic GMP-AMP (cGAMP)-stimulator of interferon genes (STING) pathway during HIPEC

During chemotherapy-induced tumor cell death, the release of DNA, a tumor-specific antigen, can enhance antitumor immune responses via the cGAS-cGAMP-STING pathway (Fig. 1). The production of cGAMP can be catalyzed by the interaction between dsDNA and cGAS, after which cGAMP activates the adaptor protein STING, a second messenger in APCs (128,129). The downstream NF-κB and TBK1-IRF3 pathways are then activated to induce the expression of type I IFNs, which induce innate and adaptive immune responses, including various immune cell subsets, such as NK cells, DCs, B cells and effector T cells (130,131). As hyperthermia can promote tumor cell DNA damage (43,44) and delay the repair of DSBs caused by chemotherapy (48,49), it may be a sensitizer for such a process. Thus, HIPEC may exert unexpected antitumor effects following activation of the CGAS-cGAMP-STING pathway.

Conclusions

Following improvements by medical personnel, HIPEC appears to increase the OS rate of patients more favorably with peritoneal metastatic carcinoma compared with systemic chemotherapy. However, the therapeutic effects of HIPEC have been debated since the publication of the PRODIGE7 and PROPHYLOCHIP trial results. Thus, additional clinical trials are required to definitively establish the role of HIPEC in abdominal metastatic adenocarcinoma.

HIPEC exhibits several noteworthy side effects, such as visceral hemorrhage, fatigue and gastrointestinal or neurotoxic effects (132,133). Intra-abdominal infection, peritoneal recurrence and small bowel obstruction are also frequently observed (134). Antibiotic-induced intestinal dysbiosis can result in the failure of cancer immunotherapy (135). Thus, detrimental changes in the gut microbiome of patients undergoing HIPEC may also occur (136,137). It was hypothesized that temperate chemotherapy delivered into the abdominal cavity may also lead to intestinal dysbiosis; however, further clarification is required.

The present review focuses on the dual actions of chemotherapeutic drugs and hyperthermia using HIPEC to clarify the molecular mechanisms underlying the enhanced efficacy of HIPEC, and to identify other therapies for its combinatory use. However, the precise molecular indexes, not just the retrospective indexes, require further investigation to predict patient prognosis.

Acknowledgements

Not applicable.

Funding

The present review was financially supported by the National Natural Science Foundation of China (grant nos. 31700792 and 81801568) and the Maternal and Child Health Association Foundation of Jiangsu (grant no. FYX202017).

Availability of data and materials

Not applicable.

Authors' contributions

YZ and YW performed the literature review and drafted the initial manuscript. JW and CW revised the manuscript for important intellectual content and confirmed the authenticity of all the raw data. All authors have read and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Lengyel E: Ovarian cancer development and metastasis. Am J Pathol. 177:1053–1064. 2010. View Article : Google Scholar : PubMed/NCBI

2 

Polyzos A, Tsavaris N, Kosmas C, Giannikos L, Katsikas M, Kalahanis N, Karatzas G, Christodoulou K, Giannakopoulos K, Stamatiadis D and Katsilambros N: A comparative study of intraperitoneal carboplatin versus intravenous carboplatin with intravenous cyclophosphamide in both arms as initial chemotherapy for stage III ovarian cancer. Oncology. 56:291–296. 1999. View Article : Google Scholar : PubMed/NCBI

3 

Arshad U, Ploylearmsaeng SA, Karlsson MO, Doroshyenko O, Langer D, Schömig E, Kunze S, Güner SA, Skripnichenko R, Ullah S, et al: Prediction of exposure-driven myelotoxicity of continuous infusion 5-fluorouracil by a semi-physiological pharmacokinetic-pharmacodynamic model in gastrointestinal cancer patients. Cancer Chemother Pharmacol. 85:711–722. 2020. View Article : Google Scholar : PubMed/NCBI

4 

Flessner MF: The transport barrier in intraperitoneal therapy. Am J Physiol Renal Physiol. 288:F433–F442. 2005. View Article : Google Scholar : PubMed/NCBI

5 

Barlin JN, Dao F, Bou Zgheib N, Ferguson SE, Sabbatini PJ, Hensley ML, Bell-McGuinn KM, Konner J, Tew WP, Aghajanian C and Chi DS: Progression-free and overall survival of a modified outpatient regimen of primary intravenous/intraperitoneal paclitaxel and intraperitoneal cisplatin in ovarian, fallopian tube, and primary peritoneal cancer. Gynecol Oncol. 125:621–624. 2012. View Article : Google Scholar : PubMed/NCBI

6 

Wust P, Hildebrandt B, Sreenivasa G, Rau B, Gellermann J, Riess H, Felix R and Schlag PM: Hyperthermia in combined treatment of cancer. Lancet Oncol. 3:487–497. 2002. View Article : Google Scholar : PubMed/NCBI

7 

van Driel WJ, Koole SN, Sikorska K, Schagen van Leeuwen JH, Schreuder HWR, Hermans RHM, de Hingh IHJT, van der Velden J, Arts HJ, Massuger LFAG, et al: Hyperthermic intraperitoneal chemotherapy in ovarian cancer. N Engl J Med. 378:230–240. 2018. View Article : Google Scholar : PubMed/NCBI

8 

Gonzalez-Moreno S: Peritoneal surface oncology: A PROGRESS REPort. Eur J Surg Oncol. 32:593–596. 2006. View Article : Google Scholar : PubMed/NCBI

9 

Quénet F, Elias D, Roca L, Goéré D, Ghouti L, Pocard M, Facy O, Arvieux C, Lorimier G, Pezet D, et al: Cytoreductive surgery plus hyperthermic intraperitoneal chemotherapy versus cytoreductive surgery alone for colorectal peritoneal metastases (PRODIGE 7): a multicentre, randomised, open-label, phase 3 trial. Lancet Oncol. 22:256–266. 2021. View Article : Google Scholar : PubMed/NCBI

10 

Goéré D, Glehen O, Quenet F, Guilloit JM, Bereder JM, Lorimier G, Thibaudeau E, Ghouti L, Pinto A, Tuech JJ, et al: Second-look surgery plus hyperthermic intraperitoneal chemotherapy versus surveillance in patients at high risk of developing colorectal peritoneal metastases (PROPHYLOCHIP-PRODIGE 15): A randomised, phase 3 study. Lancet Oncol. 21:1147–1154. 2020. View Article : Google Scholar : PubMed/NCBI

11 

González-Moreno S, González-Bayón LA and Ortega-Pérez G: Hyperthermic intraperitoneal chemotherapy: Rationale and technique. World J Gastrointest Oncol. 2:68–75. 2010. View Article : Google Scholar : PubMed/NCBI

12 

Giovanella BC, Stehlin JS and Morgan AC: Selective lethal effect of supranormal temperatures on human neoplastic cells. Cancer Res. 36:3944–3950. 1976.PubMed/NCBI

13 

Glehen O, Cotte E, Kusamura S, Deraco M, Baratti D, Passot G, Beaujard AC and Noel GF: Hyperthermic intraperitoneal chemotherapy: Nomenclature and modalities of perfusion. J Surg Oncol. 98:242–246. 2008. View Article : Google Scholar : PubMed/NCBI

14 

Sánchez-García S, Padilla-Valverde D, Villarejo-Campos P, Martín-Fernández J, García-Rojo M and Rodríguez-Martínez M: Experimental development of an intra-abdominal chemohyperthermia model using a closed abdomen technique and a PRS-1.0 Combat CO2 recirculation system. Surgery. 155:719–725. 2014. View Article : Google Scholar : PubMed/NCBI

15 

Leiting JL, Cloyd JM, Ahmed A, Fournier K, Lee AJ, Dessureault S, Felder S, Veerapong J, Baumgartner JM, Clarke C, et al: Comparison of open and closed hyperthermic intraperitoneal chemotherapy: Results from the United States hyperthermic intraperitoneal chemotherapy collaborative. World J Gastrointest Oncol. 12:756–767. 2020. View Article : Google Scholar : PubMed/NCBI

16 

Sánchez-García S, Villarejo-Campos P, Padilla-Valverde D, Amo-Salas M and Martín-Fernández J: Intraperitoneal chemotherapy hyperthermia (HIPEC) for peritoneal carcinomatosis of ovarian cancer origin by fluid and CO2 recirculation using the closed abdomen technique (PRS-1.0 Combat): A clinical pilot study. Int J Hyperthermia. 32:488–495. 2016. View Article : Google Scholar : PubMed/NCBI

17 

Khosrawipour V, Khosrawipour T, Diaz-Carballo D, Förster E, Zieren J and Giger-Pabst U: Exploring the spatial drug distribution pattern of pressurized intraperitoneal aerosol chemotherapy (PIPAC). Ann Surg Oncol. 23:1220–1224. 2016. View Article : Google Scholar : PubMed/NCBI

18 

Nadiradze G, Horvath P, Sautkin Y, Archid R, Weinreich FJ, Königsrainer A and Reymond MA: Overcoming drug resistance by taking advantage of physical principles: Pressurized intraperitoneal aerosol chemotherapy (PIPAC). Cancers (Basel). 12:342019. View Article : Google Scholar

19 

Verwaal VJ, van Ruth S, de Bree E, van Sloothen GW, van Tinteren H, Boot H and Zoetmulder FA: Randomized trial of cytoreduction and hyperthermic intraperitoneal chemotherapy versus systemic chemotherapy and palliative surgery in patients with peritoneal carcinomatosis of colorectal cancer. J Clin Oncol. 21:3737–3743. 2003. View Article : Google Scholar : PubMed/NCBI

20 

Klaver CE, Musters GD, Bemelman WA, Punt CJ, Verwaal VJ, Dijkgraaf MG, Aalbers AG, van der Bilt JD, Boerma D, Bremers AJ, et al: Adjuvant hyperthermic intraperitoneal chemotherapy (HIPEC) in patients with colon cancer at high risk of peritoneal carcinomatosis; the COLOPEC randomized multicentre trial. BMC Cancer. 15:4282015. View Article : Google Scholar : PubMed/NCBI

21 

Reutovich MY, Krasko OV and Sukonko OG: Hyperthermic intraperitoneal chemotherapy in serosa-invasive gastric cancer patients. Eur J Surg Oncol. 45:2405–2411. 2019. View Article : Google Scholar : PubMed/NCBI

22 

Yang XJ, Huang CQ, Suo T, Mei LJ, Yang GL, Cheng FL, Zhou YF, Xiong B, Yonemura Y and Li Y: Cytoreductive surgery and hyperthermic intraperitoneal chemotherapy improves survival of patients with peritoneal carcinomatosis from gastric cancer: Final results of a phase III randomized clinical trial. Ann Surg Oncol. 18:1575–1581. 2011. View Article : Google Scholar : PubMed/NCBI

23 

Spiliotis J, Halkia E, Lianos E, Kalantzi N, Grivas A, Efstathiou E and Giassas S: Cytoreductive surgery and HIPEC in recurrent epithelial ovarian cancer: A prospective randomized phase III study. Ann Surg Oncol. 22:1570–1575. 2015. View Article : Google Scholar : PubMed/NCBI

24 

Glockzin G, Rochon J, Arnold D, Lang SA, Klebl F, Zeman F, Koller M, Schlitt HJ and Piso P: A prospective multicenter phase II study evaluating multimodality treatment of patients with peritoneal carcinomatosis arising from appendiceal and colorectal cancer: The COMBATAC trial. BMC Cancer. 13:672013. View Article : Google Scholar : PubMed/NCBI

25 

van Leeuwen BL, Graf W, Pahlman L and Mahteme H: Swedish experience with peritonectomy and HIPEC. HIPEC in peritoneal carcinomatosis. Ann Surg Oncol. 15:745–753. 2008. View Article : Google Scholar : PubMed/NCBI

26 

Zivanovic O, Chi DS, Filippova O, Randall LM, Bristow RE and O'Cearbhaill RE: It's time to warm up to hyperthermic intraperitoneal chemotherapy for patients with ovarian cancer. Gynecol Oncol. 151:555–561. 2018. View Article : Google Scholar : PubMed/NCBI

27 

Kusamura S, Azmi N, Fumagalli L, Baratti D, Guaglio M, Cavalleri A, Garrone G, Battaglia L, Barretta F and Deraco M: Phase II randomized study on tissue distribution and pharmacokinetics of cisplatin according to different levels of intra-abdominal pressure (IAP) during HIPEC (NCT02949791). Eur J Surg Oncol. 47:82–88. 2021. View Article : Google Scholar : PubMed/NCBI

28 

Cesna V, Sukovas A, Jasukaitiene A, Naginiene R, Barauskas G, Dambrauskas Z, Paskauskas S and Gulbinas A: Narrow line between benefit and harm: Additivity of hyperthermia to cisplatin cytotoxicity in different gastrointestinal cancer cells. World J Gastroenterol. 24:1072–1083. 2018. View Article : Google Scholar : PubMed/NCBI

29 

Ceelen W, Braet H, van Ramshorst G, Willaert W and Remaut K: Intraperitoneal chemotherapy for peritoneal metastases: An expert opinion. Expert Opin Drug Deliv. 17:511–522. 2020. View Article : Google Scholar : PubMed/NCBI

30 

Zhang Y and Calderwood SK: Autophagy, protein aggregation and hyperthermia: A mini-review. Int J Hyperthermia. 27:409–414. 2011. View Article : Google Scholar : PubMed/NCBI

31 

Ahmed K, Zaidi SF, Mati-Ur-Rehman, Rehman R and Kondo T: Hyperthermia and protein homeostasis: Cytoprotection and cell death. J Therm Biol. 91:1026152020. View Article : Google Scholar : PubMed/NCBI

32 

Franke K, Kettering M, Lange K, Kaiser WA and Hilger I: The exposure of cancer cells to hyperthermia, iron oxide nanoparticles, and mitomycin C influences membrane multidrug resistance protein expression levels. Int J Nanomedicine. 8:351–363. 2013.PubMed/NCBI

33 

Luchetti F, Mannello F, Canonico B, Battistelli M, Burattini S and Falcieri E: Integrin and cytoskeleton behaviour in human neuroblastoma cells during hyperthermia-related apoptosis. Apoptosis. 9:635–648. 2004. View Article : Google Scholar : PubMed/NCBI

34 

Luchetti F, Canonico B, Della Felice M, Burattini S, Battistelli M, Papa S and Falcieri E: Hyperthermia triggers apoptosis and affects cell adhesiveness in human neuroblastoma cells. Histol Histopathol. 18:1041–1052. 2003.PubMed/NCBI

35 

Onishi Y, Fehervari Z, Yamaguchi T and Sakaguchi S: Foxp3+ natural regulatory T cells preferentially form aggregates on dendritic cells in vitro and actively inhibit their maturation. Proc Natl Acad Sci USA. 105:10113–10118. 2008. View Article : Google Scholar : PubMed/NCBI

36 

Alvarez-Berríos MP, Castillo A, Mendéz J, Soto O, Rinaldi C and Torres-Lugo M: Hyperthermic potentiation of cisplatin by magnetic nanoparticle heaters is correlated with an increase in cell membrane fluidity. Int J Nanomedicine. 8:1003–1013. 2013.PubMed/NCBI

37 

Csoboz B, Balogh GE, Kusz E, Gombos I, Peter M, Crul T, Gungor B, Haracska L, Bogdanovics G, Torok Z, et al: Membrane fluidity matters: Hyperthermia from the aspects of lipids and membranes. Int J Hyperthermia. 29:491–499. 2013. View Article : Google Scholar : PubMed/NCBI

38 

de Andrade Mello P, Bian S, Savio LEB, Zhang H, Zhang J, Junger W, Wink MR, Lenz G, Buffon A, Wu Y and Robson SC: Hyperthermia and associated changes in membrane fluidity potentiate P2X7 activation to promote tumor cell death. Oncotarget. 8:67254–67268. 2017. View Article : Google Scholar : PubMed/NCBI

39 

van Oorschot B, Granata G, Di Franco S, Ten Cate R, Rodermond HM, Todaro M, Medema JP and Franken NA: Targeting DNA double strand break repair with hyperthermia and DNA-PKcs inhibition to enhance the effect of radiation treatment. Oncotarget. 7:65504–65513. 2016. View Article : Google Scholar : PubMed/NCBI

40 

Mehta IS, Kulashreshtha M, Chakraborty S, Kolthur-Seetharam U and Rao BJ: Chromosome territories reposition during DNA damage-repair response. Genome Biol. 14:1352013. View Article : Google Scholar : PubMed/NCBI

41 

Warters RL and Henle KJ: DNA degradation in chinese hamster ovary cells after exposure to hyperthermia. Cancer Res. 42:4427–4432. 1982.PubMed/NCBI

42 

Takahashi A, Matsumoto H, Nagayama K, Kitano M, Hirose S, Tanaka H, Mori E, Yamakawa N, Yasumoto J, Yuki K, et al: Evidence for the involvement of double-strand breaks in heat-induced cell killing. Cancer Res. 64:8839–8845. 2004. View Article : Google Scholar : PubMed/NCBI

43 

Takahashi A, Mori E, Somakos GI, Ohnishi K and Ohnishi T: Heat induces gammaH2AX foci formation in mammalian cells. Mutat Res. 656:88–92. 2008. View Article : Google Scholar : PubMed/NCBI

44 

Hunt CR, Pandita RK, Laszlo A, Higashikubo R, Agarwal M, Kitamura T, Gupta A, Rief N, Horikoshi N, Baskaran R, et al: Hyperthermia activates a subset of ataxia-telangiectasia mutated effectors independent of DNA strand breaks and heat shock protein 70 status. Cancer Res. 67:3010–3017. 2007. View Article : Google Scholar : PubMed/NCBI

45 

El-Awady RA, Dikomey E and Dahm-Daphi J: Heat effects on DNA repair after ionising radiation: Hyperthermia commonly increases the number of non-repaired double-strand breaks and structural rearrangements. Nucleic Acids Res. 29:1960–1966. 2001. View Article : Google Scholar : PubMed/NCBI

46 

Muenyi CS, States VA, Masters JH, Fan TW, Helm CW and States JC: Sodium arsenite and hyperthermia modulate cisplatin-DNA damage responses and enhance platinum accumulation in murine metastatic ovarian cancer xenograft after hyperthermic intraperitoneal chemotherapy (HIPEC). J Ovarian Res. 4:92011. View Article : Google Scholar : PubMed/NCBI

47 

Oei AL, Vriend LE, Crezee J, Franken NA and Krawczyk PM: Effects of hyperthermia on DNA repair pathways: One treatment to inhibit them all. Radiat Oncol. 10:1652015. View Article : Google Scholar : PubMed/NCBI

48 

Krawczyk PM, Eppink B, Essers J, Stap J, Rodermond H, Odijk H, Zelensky A, van Bree C, Stalpers LJ, Buist MR, et al: Mild hyperthermia inhibits homologous recombination, induces BRCA2 degradation, and sensitizes cancer cells to poly (ADP-ribose) polymerase-1 inhibition. Proc Natl Acad Sci USA. 108:9851–9856. 2011. View Article : Google Scholar : PubMed/NCBI

49 

Zhang J, Zhao B, Chen S, Wang Y, Zhang Y, Wang Y, Wei D, Zhang L, Rong G and Weng Y: Near-infrared light irradiation induced Mild hyperthermia enhances glutathione depletion and DNA interstrand cross-link formation for efficient chemotherapy. ACS Nano. 14:14831–14845. 2020. View Article : Google Scholar : PubMed/NCBI

50 

Ohno S, Siddik ZH, Kido Y, Zwelling LA and Bull JM: Thermal enhancement of drug uptake and DNA adducts as a possible mechanism for the effect of sequencing hyperthermia on cisplatin-induced cytotoxicity in L1210 cells. Cancer Chemother Pharmacol. 34:302–306. 1994. View Article : Google Scholar : PubMed/NCBI

51 

Sato I, Umemura M, Mitsudo K, Kioi M, Nakashima H, Iwai T, Feng X, Oda K, Miyajima A, Makino A, et al: Hyperthermia generated with ferucarbotran (Resovist®) in an alternating magnetic field enhances cisplatin-induced apoptosis of cultured human oral cancer cells. J Physiol Sci. 64:177–183. 2014. View Article : Google Scholar : PubMed/NCBI

52 

Clavel CM, Nowak-Sliwinska P, Păunescu E, Griffioen AW and Dyson PJ: In vivo evaluation of small-molecule thermoresponsive anticancer drugs potentiated by hyperthermia. Chem Sci. 6:2795–2801. 2015. View Article : Google Scholar : PubMed/NCBI

53 

Peer AJ, Grimm MJ, Zynda ER and Repasky EA: Diverse immune mechanisms may contribute to the survival benefit seen in cancer patients receiving hyperthermia. Immunol Res. 46:137–154. 2010. View Article : Google Scholar : PubMed/NCBI

54 

Borst J, Ahrends T, Bąbała N, Melief CJM and Kastenmüller W: CD4+T cell help in cancer immunology and immunotherapy. Nat Rev Immunol. 18:635–647. 2018. View Article : Google Scholar : PubMed/NCBI

55 

Schuijs MJ, Hammad H and Lambrecht BN: Professional and ‘Amateur’ Antigen-Presenting Cells In Type 2 Immunity. Trends Immunol. 40:22–34. 2019. View Article : Google Scholar : PubMed/NCBI

56 

Evans SS, Repasky EA and Fisher DT: Fever and the thermal regulation of immunity: The immune system feels the heat. Nat Rev Immunol. 15:335–349. 2015. View Article : Google Scholar : PubMed/NCBI

57 

Ostberg JR, Dayanc BE, Yuan M, Oflazoglu E and Repasky EA: Enhancement of natural killer (NK) cell cytotoxicity by fever-range thermal stress is dependent on NKG2D function and is associated with plasma membrane NKG2D clustering and increased expression of MICA on target cells. J Leukoc Biol. 82:1322–1331. 2007. View Article : Google Scholar : PubMed/NCBI

58 

Zhang L, Liu F, Peng Y, Sun L and Chen G: Changes in expression of four molecular marker proteins and one microRNA in mesothelial cells of the peritoneal dialysate effluent fluid of peritoneal dialysis patients. Exp Ther Med. 6:1189–1193. 2013. View Article : Google Scholar : PubMed/NCBI

59 

Rynne-Vidal A, Au-Yeung CL, Jiménez-Heffernan JA, Pérez-Lozano ML, Cremades-Jimeno L, Bárcena C, Cristóbal-García I, Fernández-Chacón C, Yeung TL, Mok SC, et al: Mesothelial-to-mesenchymal transition as a possible therapeutic target in peritoneal metastasis of ovarian cancer. J Pathol. 242:140–151. 2017. View Article : Google Scholar : PubMed/NCBI

60 

Pathria P, Louis TL and Varner JA: Targeting Tumor-associated macrophages in cancer. Trends Immunol. 40:310–327. 2019. View Article : Google Scholar : PubMed/NCBI

61 

Li X, Liu R, Su X, Pan Y, Han X, Shao C and Shi Y: Harnessing tumor-associated macrophages as aids for cancer immunotherapy. Mol Cancer. 18:1772019. View Article : Google Scholar : PubMed/NCBI

62 

Farhood B, Najafi M and Mortezaee K: CD8(+) cytotoxic T lymphocytes in cancer immunotherapy: A review. J Cell Physiol. 234:8509–8521. 2019. View Article : Google Scholar : PubMed/NCBI

63 

Yin Z, Ma T, Lin Y, Lu X, Zhang C, Chen S and Jian Z: IL-6/STAT3 pathway intermediates M1/M2 macrophage polarization during the development of hepatocellular carcinoma. J Cell Biochem. 119:9419–9432. 2018. View Article : Google Scholar : PubMed/NCBI

64 

Sica A and Mantovani A: Macrophage plasticity and polarization: In vivo veritas. J Clin Invest. 122:787–795. 2012. View Article : Google Scholar : PubMed/NCBI

65 

Qu D, Qin Y, Liu Y, Liu T, Liu C, Han T, Chen Y, Ma C and Li X: Fever-inducible lipid nanocomposite for boosting cancer therapy through synergistic engineering of a tumor microenvironment. ACS Appl Mater Interfaces. 12:32301–32311. 2020. View Article : Google Scholar : PubMed/NCBI

66 

Frey B, Weiss EM, Rubner Y, Wunderlich R, Ott OJ, Sauer R, Fietkau R and Gaipl US: Old and new facts about hyperthermia-induced modulations of the immune system. Int J Hyperthermia. 28:528–542. 2012. View Article : Google Scholar : PubMed/NCBI

67 

Agarwal SS, Katz EJ and Loeb LA: Effect of hyperthermia on the survival of normal human peripheral blood mononuclear cells. Cancer Res. 43:3124–3126. 1983.PubMed/NCBI

68 

Harden LM, Kent S, Pittman QJ and Roth J: Fever and sickness behavior: Friend or foe? Brain Behav Immun. 50:322–333. 2015. View Article : Google Scholar : PubMed/NCBI

69 

Zhang L, Zhang Y, Xue Y, Wu Y, Wang Q, Xue L, Su Z and Zhang C: Transforming weakness into strength: Photothermal-therapy-induced inflammation enhanced cytopharmaceutical chemotherapy as a combination anticancer treatment. Adv Mater. 31:e18059362019.PubMed/NCBI

70 

Mantovani A, Barajon I and Garlanda C: IL-1 and IL-1 regulatory pathways in cancer progression and therapy. Immunol Rev. 281:57–61. 2018. View Article : Google Scholar : PubMed/NCBI

71 

Kitamura H, Ohno Y, Toyoshima Y, Ohtake J, Homma S, Kawamura H, Takahashi N and Taketomi A: Interleukin-6/STAT3 signaling as a promising target to improve the efficacy of cancer immunotherapy. Cancer Sci. 108:1947–1952. 2017. View Article : Google Scholar : PubMed/NCBI

72 

Korn T, Bettelli E, Oukka M and Kuchroo VK: IL-17 and Th17 cells. Annu Rev Immunol. 27:485–517. 2009. View Article : Google Scholar : PubMed/NCBI

73 

Appenheimer MM, Chen Q, Girard RA, Wang WC and Evans SS: Impact of fever-range thermal stress on lymphocyte-endothelial adhesion and lymphocyte trafficking. Immunol Invest. 34:295–323. 2005. View Article : Google Scholar : PubMed/NCBI

74 

Chonov DC, Ignatova MMK, Ananiev JR and Gulubova MV: IL-6 Activities in the Tumour Microenvironment. Part 1. Open Access Maced J Med Sci. 7:2391–2398. 2019. View Article : Google Scholar : PubMed/NCBI

75 

Wagner AC, Weber H, Jonas L, Nizze H, Strowski M, Fiedler F, Printz H, Steffen H and Göke B: Hyperthermia induces heat shock protein expression and protection against cerulein-induced pancreatitis in rats. Gastroenterology. 111:1333–1342. 1996. View Article : Google Scholar : PubMed/NCBI

76 

Burd R, Dziedzic TS, Xu Y, Caligiuri MA, Subjeck JR and Repasky EA: Tumor cell apoptosis, lymphocyte recruitment and tumor vascular changes are induced by low temperature, long duration (fever-like) whole body hyperthermia. J Cell Physiol. 177:137–147. 1998. View Article : Google Scholar : PubMed/NCBI

77 

Hartl FU and Hayer-Hartl M: Molecular chaperones in the cytosol: From nascent chain to folded protein. Science. 295:1852–1858. 2002. View Article : Google Scholar : PubMed/NCBI

78 

Calderwood SK and Gong J: Heat shock proteins promote cancer: It's a protection racket. Trends Biochem Sci. 41:311–323. 2016. View Article : Google Scholar : PubMed/NCBI

79 

Pelz JO, Vetterlein M, Grimmig T, Kerscher AG, Moll E, Lazariotou M, Matthes N, Faber M, Germer CT, Waaga-Gasser AM and Gasser M: Hyperthermic intraperitoneal chemotherapy in patients with peritoneal carcinomatosis: Role of heat shock proteins and dissecting effects of hyperthermia. Ann Surg Oncol. 20:1105–1113. 2013. View Article : Google Scholar : PubMed/NCBI

80 

Kepenekian V, Aloy MT, Magné N, Passot G, Armandy E, Decullier E, Sayag-Beaujard A, Gilly FN, Glehen O and Rodriguez-Lafrasse C: Impact of hyperthermic intraperitoneal chemotherapy on Hsp27 protein expression in serum of patients with peritoneal carcinomatosis. Cell Stress Chaperones. 18:623–630. 2013. View Article : Google Scholar : PubMed/NCBI

81 

Mu C, Wu X, Zhou X, Wolfram J, Shen J, Zhang D, Mai J, Xia X, Holder AM, Ferrari M, et al: Chemotherapy sensitizes therapy-resistant cells to Mild hyperthermia by suppressing heat shock protein 27 expression in triple-negative breast cancer. Clin Cancer Res. 24:4900–4912. 2018. View Article : Google Scholar : PubMed/NCBI

82 

Zunino B, Rubio-Patiño C, Villa E, Meynet O, Proics E, Cornille A, Pommier S, Mondragón L, Chiche J, Bereder JM, et al: Hyperthermic intraperitoneal chemotherapy leads to an anticancer immune response via exposure of cell surface heat shock protein 90. Oncogene. 35:261–268. 2016. View Article : Google Scholar : PubMed/NCBI

83 

Isambert N, Delord JP, Soria JC, Hollebecque A, Gomez-Roca C, Purcea D, Rouits E, Belli R and Fumoleau P: Debio0932, a second-generation oral heat shock protein (HSP) inhibitor, in patients with advanced cancer-results of a first-in-man dose-escalation study with a fixed-dose extension phase. Ann Oncol. 26:1005–1011. 2015. View Article : Google Scholar : PubMed/NCBI

84 

Larson N, Gormley A, Frazier N and Ghandehari H: Synergistic enhancement of cancer therapy using a combination of heat shock protein targeted HPMA copolymer-drug conjugates and gold nanorod induced hyperthermia. J Control Release. 170:41–50. 2013. View Article : Google Scholar : PubMed/NCBI

85 

Taha EA, Ono K and Eguchi T: Roles of extracellular HSPs as biomarkers in immune surveillance and immune evasion. Int J Mol Sci. 20:45882019. View Article : Google Scholar

86 

Mukhopadhaya A, Mendecki J, Dong X, Liu L, Kalnicki S, Garg M, Alfieri A and Guha C: Localized hyperthermia combined with intratumoral dendritic cells induces systemic antitumor immunity. Cancer Res. 67:7798–7806. 2007. View Article : Google Scholar : PubMed/NCBI

87 

Zhang HG, Mehta K, Cohen P and Guha C: Hyperthermia on immune regulation: A temperature's story. Cancer Lett. 271:191–204. 2008. View Article : Google Scholar : PubMed/NCBI

88 

Torigoe T, Tamura Y and Sato N: Heat shock proteins and immunity: Application of hyperthermia for immunomodulation. Int J Hyperthermia. 25:610–616. 2009. View Article : Google Scholar : PubMed/NCBI

89 

Calderwood SK, Theriault JR and Gong J: How is the immune response affected by hyperthermia and heat shock proteins? Int J Hyperthermia. 21:713–716. 2005. View Article : Google Scholar : PubMed/NCBI

90 

Lee S, Son B, Park G, Kim H, Kang H, Jeon J, Youn H and Youn B: Immunogenic effect of hyperthermia on enhancing radiotherapeutic efficacy. Int J Mol Sci. 19:27952018. View Article : Google Scholar

91 

van Baal JO, Van de Vijver KK, Nieuwland R, van Noorden CJ, van Driel WJ, Sturk A, Kenter GG, Rikkert LG and Lok CA: The histophysiology and pathophysiology of the peritoneum. Tissue Cell. 49:95–105. 2017. View Article : Google Scholar : PubMed/NCBI

92 

Kastelein AW, Vos LMC, de Jong KH, van Baal JOAM, Nieuwland R, van Noorden CJF, Roovers JWR and Lok CAR: Embryology, anatomy, physiology and pathophysiology of the peritoneum and the peritoneal vasculature. Semin Cell Dev Biol. 92:27–36. 2019. View Article : Google Scholar : PubMed/NCBI

93 

de Bree E, Michelakis D, Stamatiou D, Romanos J and Zoras O: Pharmacological principles of intraperitoneal and bidirectional chemotherapy. Pleura Peritoneum. 2:47–62. 2017. View Article : Google Scholar : PubMed/NCBI

94 

Ceelen WP and Flessner MF: Intraperitoneal therapy for peritoneal tumors: Biophysics and clinical evidence. Nat Rev Clin Oncol. 7:108–115. 2010. View Article : Google Scholar : PubMed/NCBI

95 

van Ruth S, Mathôt RA, Sparidans RW, Beijnen JH, Verwaal VJ and Zoetmulder FA: Population pharmacokinetics and pharmacodynamics of mitomycin during intraoperative hyperthermic intraperitoneal chemotherapy. Clinical Pharmacokinet. 43:131–143. 2004. View Article : Google Scholar

96 

Cashin PH, Ehrsson H, Wallin I, Nygren P and Mahteme H: Pharmacokinetics of cisplatin during hyperthermic intraperitoneal treatment of peritoneal carcinomatosis. Eur J Clin Pharmacol. 69:533–540. 2013. View Article : Google Scholar : PubMed/NCBI

97 

Leinwand JC, Bates GE, Allendorf JD, Chabot JA, Lewin SN and Taub RN: Body surface area predicts plasma oxaliplatin and pharmacokinetic advantage in hyperthermic intraoperative intraperitoneal chemotherapy. Ann Surg Oncol. 20:1101–1104. 2013. View Article : Google Scholar : PubMed/NCBI

98 

de Bree E, Rosing H, Filis D, Romanos J, Melisssourgaki M, Daskalakis M, Pilatou M, Sanidas E, Taflampas P, Kalbakis K, et al: Cytoreductive surgery and intraoperative hyperthermic intraperitoneal chemotherapy with paclitaxel: A clinical and pharmacokinetic study. Ann Surg Oncol. 15:1183–1192. 2008. View Article : Google Scholar : PubMed/NCBI

99 

de Bree E, Rosing H, Beijnen JH, Romanos J, Michalakis J, Georgoulias V and Tsiftsis DD: Pharmacokinetic study of docetaxel in intraoperative hyperthermic i.p. chemotherapy for ovarian cancer. Anticancer Drugs. 14:103–110. 2003. View Article : Google Scholar : PubMed/NCBI

100 

Nicoletto MO, Padrini R, Galeotti F, Ferrazzi E, Cartei G, Riddi F, Palumbo M, De Paoli M and Corsini A: Pharmacokinetics of intraperitoneal hyperthermic perfusion with mitoxantrone in ovarian cancer. Cancer Chemother Pharmacol. 45:457–462. 2000. View Article : Google Scholar : PubMed/NCBI

101 

Rossi CR, Mocellin S, Pilati P, Foletto M, Quintieri L, Palatini P and Lise M: Pharmacokinetics of intraperitoneal cisplatin and doxorubicin. Surg Oncol Clin N Am. 12:781–794. 2003. View Article : Google Scholar : PubMed/NCBI

102 

Choi YH: Interpretation of drug interaction using systemic and local tissue exposure changes. Pharmaceutics. 12:4172020. View Article : Google Scholar

103 

Tentes AA, Kyziridis D, Kakolyris S, Pallas N, Zorbas G, Korakianitis O, Mavroudis C, Courcoutsakis N and Prasopoulos P: Preliminary results of hyperthermic intraperitoneal intraoperative chemotherapy as an adjuvant in resectable pancreatic cancer. Gastroenterol Res Pract. 2012:5065712012. View Article : Google Scholar : PubMed/NCBI

104 

Lemoine L, Thijssen E, Carleer R, Cops J, Lemmens V, Eyken PV, Sugarbaker P and der Speeten KV: Body surface area-based versus concentration-based intraperitoneal perioperative chemotherapy in a rat model of colorectal peritoneal surface malignancy: Pharmacologic guidance towards standardization. Oncotarget. 10:1407–1424. 2019. View Article : Google Scholar : PubMed/NCBI

105 

Lemoine L, Thijssen E, Carleer R, Geboers K, Sugarbaker P and van der Speeten K: Body surface area-based vs concentration-based perioperative intraperitoneal chemotherapy after optimal cytoreductive surgery in colorectal peritoneal surface malignancy treatment: COBOX trial. J Surg Oncol. 119:999–1010. 2019. View Article : Google Scholar : PubMed/NCBI

106 

Liesenfeld LF, Hillebrecht HC, Klose J, Schmidt T and Schneider M: Impact of perfusate concentration on hyperthermic intraperitoneal chemotherapy efficacy and toxicity in a rodent model. J Surg Res. 253:262–271. 2020. View Article : Google Scholar : PubMed/NCBI

107 

Shah DK, Shin BS, Veith J, Tóth K, Bernacki RJ and Balthasar JP: Use of an anti-vascular endothelial growth factor antibody in a pharmacokinetic strategy to increase the efficacy of intraperitoneal chemotherapy. J Pharmacol Exp Ther. 329:580–591. 2009. View Article : Google Scholar : PubMed/NCBI

108 

Gremonprez F, Descamps B, Izmer A, Vanhove C, Vanhaecke F, De Wever O and Ceelen W: Pretreatment with VEGF(R)-inhibitors reduces interstitial fluid pressure, increases intraperitoneal chemotherapy drug penetration, and impedes tumor growth in a mouse colorectal carcinomatosis model. Oncotarget. 6:29889–29900. 2015. View Article : Google Scholar : PubMed/NCBI

109 

Li H, Mao X, Liu K, Sun J, Li B, Malyar RM, Liu D, Pan C, Gan F and Liu Y: A pilot study of combination intraperitoneal recombinant human endostatin and chemotherapy for refractory malignant ascites secondary to ovarian cancer. Med Oncol. 31:9302014. View Article : Google Scholar : PubMed/NCBI

110 

Cristea MC, Frankel P, Synold T, Rivkin S, Lim D, Chung V, Chao J, Wakabayashi M, Paz B, Han E, et al: A phase I trial of intraperitoneal nab-paclitaxel in the treatment of advanced malignancies primarily confined to the peritoneal cavity. Cancer Chemother Pharmaco. 83:589–598. 2019. View Article : Google Scholar

111 

Shamsi M, Sedaghatkish A, Dejam M, Saghafian M, Mohammadi M and Sanati-Nezhad A: Magnetically assisted intraperitoneal drug delivery for cancer chemotherapy. Drug Deliv. 25:846–861. 2018. View Article : Google Scholar : PubMed/NCBI

112 

Sugarbaker PH and Van der Speeten K: Surgical technology and pharmacology of hyperthermic perioperative chemotherapy. J Gastrointest Oncol. 7:29–44. 2016.PubMed/NCBI

113 

Dakwar GR, Shariati M, Willaert W, Ceelen W, De Smedt SC and Remaut K: Nanomedicine-based intraperitoneal therapy for the treatment of peritoneal carcinomatosis-Mission possible? Adv Drug Deliv Rev. 108:13–24. 2017. View Article : Google Scholar : PubMed/NCBI

114 

Galluzzi L, Buqué A, Kepp O, Zitvogel L and Kroemer G: Immunological effects of conventional chemotherapy and targeted anticancer Agents. Cancer Cell. 28:690–714. 2015. View Article : Google Scholar : PubMed/NCBI

115 

Coffelt SB and de Visser KE: Immune-mediated mechanisms influencing the efficacy of anticancer therapies. Trends Immunol. 36:198–216. 2015. View Article : Google Scholar : PubMed/NCBI

116 

Latchman Y, Wood CR, Chernova T, Chaudhary D, Borde M, Chernova I, Iwai Y, Long AJ, Brown JA, Nunes R, et al: PD-L2 is a second ligand for PD-1 and inhibits T cell activation. Nat Immunol. 2:261–268. 2001. View Article : Google Scholar : PubMed/NCBI

117 

Pfistershammer K, Klauser C, Pickl WF, Stöckl J, Leitner J, Zlabinger G, Majdic O and Steinberger P: No evidence for dualism in function and receptors: PD-L2/B7-DC is an inhibitory regulator of human T cell activation. Eur J Immunol. 36:1104–1113. 2006. View Article : Google Scholar : PubMed/NCBI

118 

Schiavoni G, Sistigu A, Valentini M, Mattei F, Sestili P, Spadaro F, Sanchez M, Lorenzi S, D'Urso MT, Belardelli F, et al: Cyclophosphamide synergizes with type I interferons through systemic dendritic cell reactivation and induction of immunogenic tumor apoptosis. Cancer Res. 71:768–778. 2011. View Article : Google Scholar : PubMed/NCBI

119 

Wu J and Waxman DJ: Metronomic cyclophosphamide eradicates large implanted GL261 gliomas by activating antitumor Cd8 T-cell responses and immune memory. Oncoimmunology. 4:e10055212015. View Article : Google Scholar : PubMed/NCBI

120 

Chen C, Chen Z, Chen D, Zhang B, Wang Z and Le H: Suppressive effects of gemcitabine plus cisplatin chemotherapy on regulatory T cells in nonsmall-cell lung cancer. J Int Med Res. 43:180–187. 2015. View Article : Google Scholar : PubMed/NCBI

121 

Zitvogel L, Galluzzi L, Smyth MJ and Kroemer G: Mechanism of action of conventional and targeted anticancer therapies: Reinstating immunosurveillance. Immunity. 39:74–88. 2013. View Article : Google Scholar : PubMed/NCBI

122 

Kepp O, Senovilla L, Vitale I, Vacchelli E, Adjemian S, Agostinis P, Apetoh L, Aranda F, Barnaba V, Bloy N, et al: Consensus guidelines for the detection of immunogenic cell death. Oncoimmunology. 3:e9556912014. View Article : Google Scholar : PubMed/NCBI

123 

Garg AD and Agostinis P: Cell death and immunity in cancer: From danger signals to mimicry of pathogen defense responses. Immunol Rev. 280:126–148. 2017. View Article : Google Scholar : PubMed/NCBI

124 

Krysko DV, Garg AD, Kaczmarek A, Krysko O, Agostinis P and Vandenabeele P: Immunogenic cell death and DAMPs in cancer therapy. Nat Rev Cancer. 12:860–875. 2012. View Article : Google Scholar : PubMed/NCBI

125 

Banchereau J, Briere F, Caux C, Davoust J, Lebecque S, Liu YJ, Pulendran B and Palucka K: Immunobiology of dendritic cells. Annu Rev Immunol. 18:767–811. 2000. View Article : Google Scholar : PubMed/NCBI

126 

Buqué A and Galluzzi L: Modeling tumor immunology and immunotherapy in Mice. Trends Cancer. 4:599–601. 2018. View Article : Google Scholar : PubMed/NCBI

127 

Curiel TJ: Immunotherapy: A useful strategy to help combat multidrug resistance. Drug Resist Updat. 15:106–113. 2012. View Article : Google Scholar : PubMed/NCBI

128 

Chen Q, Sun L and Chen ZJ: Regulation and function of the cGAS-STING pathway of cytosolic DNA sensing. Nat Immunol. 17:1142–1149. 2016. View Article : Google Scholar : PubMed/NCBI

129 

Li A, Yi M, Qin S, Song Y, Chu Q and Wu K: Activating cGAS-STING pathway for the optimal effect of cancer immunotherapy. J Hematol Oncol. 12:352019. View Article : Google Scholar : PubMed/NCBI

130 

Le Bon A, Thompson C, Kamphuis E, Durand V, Rossmann C, Kalinke U and Tough DF: Cutting edge: Enhancement of antibody responses through direct stimulation of B and T cells by type I IFN. J Immunol. 176:2074–2078. 2006. View Article : Google Scholar : PubMed/NCBI

131 

Fuertes MB, Woo SR, Burnett B, Fu YX and Gajewski TF: Type I interferon response and innate immune sensing of cancer. Trends Immunol. 34:67–73. 2013. View Article : Google Scholar : PubMed/NCBI

132 

Bhagwandin SB, Naffouje S and Salti G: Delayed presentation of major complications in patients undergoing cytoreductive surgery plus hyperthermic intraperitoneal chemotherapy following hospital discharge. J Surg Oncol. 111:324–327. 2015. View Article : Google Scholar : PubMed/NCBI

133 

Blaj S, Nedelcut S, Mayr M, Leebmann H, Leucuta D, Glockzin G and Piso P: Re-operations for early postoperative complications after CRS and HIPEC: Indication, timing, procedure, and outcome. Langenbecks Arch Surg. 404:541–546. 2019. View Article : Google Scholar : PubMed/NCBI

134 

Dreznik Y, Hoffman A, Hamburger T, Ben-Yaacov A, Dux Y, Jacoby H, Berger Y, Nissan A and Gutman M: Hospital readmission rates and risk factors for readmission following cytoreductive surgery (CRS) and hyperthermic intraperitoneal chemotherapy (HIPEC) for peritoneal surface malignancies. Surgeon. 16:278–282. 2018. View Article : Google Scholar : PubMed/NCBI

135 

Panebianco C, Andriulli A and Pazienza V: Pharmacomicrobiomics: Exploiting the drug-microbiota interactions in anticancer therapies. Microbiome. 6:922018. View Article : Google Scholar : PubMed/NCBI

136 

Routy B, Le Chatelier E, Derosa L, Duong CPM, Alou MT, Daillère R, Fluckiger A, Messaoudene M, Rauber C, Roberti MP, et al: Gut microbiome influences efficacy of PD-1-based immunotherapy against epithelial tumors. Science. 359:91–97. 2018. View Article : Google Scholar : PubMed/NCBI

137 

Routy B, Gopalakrishnan V, Daillère R, Zitvogel L, Wargo JA and Kroemer G: The gut microbiota influences anticancer immunosurveillance and general health. Nat Rev Clin Oncol. 15:382–396. 2018. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

April-2021
Volume 45 Issue 4

Print ISSN: 1021-335X
Online ISSN:1791-2431

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Zhang Y, Wu Y, Wu J and Wu C: Direct and indirect anticancer effects of hyperthermic intraperitoneal chemotherapy on peritoneal malignancies (Review). Oncol Rep 45: 23, 2021
APA
Zhang, Y., Wu, Y., Wu, J., & Wu, C. (2021). Direct and indirect anticancer effects of hyperthermic intraperitoneal chemotherapy on peritoneal malignancies (Review). Oncology Reports, 45, 23. https://doi.org/10.3892/or.2021.7974
MLA
Zhang, Y., Wu, Y., Wu, J., Wu, C."Direct and indirect anticancer effects of hyperthermic intraperitoneal chemotherapy on peritoneal malignancies (Review)". Oncology Reports 45.4 (2021): 23.
Chicago
Zhang, Y., Wu, Y., Wu, J., Wu, C."Direct and indirect anticancer effects of hyperthermic intraperitoneal chemotherapy on peritoneal malignancies (Review)". Oncology Reports 45, no. 4 (2021): 23. https://doi.org/10.3892/or.2021.7974