12,019
Views
46
CrossRef citations to date
0
Altmetric
Reviews

Drugs, doses, and durations of intraperitoneal chemotherapy: standardising HIPEC and EPIC for colorectal, appendiceal, gastric, ovarian peritoneal surface malignancies and peritoneal mesothelioma

, &
Pages 582-592 | Received 20 Nov 2016, Accepted 01 Feb 2017, Published online: 30 Jun 2017

Abstract

Peritoneal surface malignancy (PSM) is a common manifestation of digestive and gynaecologic malignancies alike. At present, patients with isolated PSM are treated with a combination therapy of cytoreductive surgery (CRS) and hyperthermic peroperative intraperitoneal chemotherapy (HIPEC). The combination of CRS and intraperitoneal (IP) chemotherapy should now be considered standard of care for PSM from appendiceal epithelial cancers, colorectal cancer and peritoneal mesothelioma. Although there is a near universal standardisation regarding the CRS, we are still lacking a much-needed standardisation amongst the various IP chemotherapy treatment modalities used today in clinical practice. Pharmacologic evidence should be generated to answer important questions raised by the myriad of variables associated with IP chemotherapy.

Introduction

Peritoneal surface malignancy (PSM) is a common manifestation of digestive and gynaecologic malignancies alike. At present, patients with isolated PSM are treated with a combination therapy of cytoreductive surgery (CRS) and hyperthermic peroperative intraperitoneal chemotherapy (HIPEC) [Citation1]. CRS and HIPEC have evolved over three decades and have demonstrated encouraging clinical results in several phase II and III trials [Citation2–10]. The combined treatment modality should now be considered standard of care for PSM from appendiceal epithelial cancers, colorectal cancer and peritoneal mesothelioma [Citation11–13]. Promising results have also been published for HIPEC in ovarian cancer and gastric cancer [Citation5,Citation9,Citation14]. Although there is now a clearly defined standardisation of CRS, based on the work by Sugarbaker et al. [Citation15,Citation16], no standardised intraperitoneal (IP) chemotherapy treatment modalities exist. Variables to be considered are: normothermic versus hyperthermic IP chemotherapy, open versus closed HIPEC technique, but also the use of HIPEC with or without early postoperative intraperitoneal chemotherapy (EPIC) or the sole administration of EPIC. This also implicates the important pharmacologic variables associated with the chemotherapy agents that are currently available for the administration of HIPEC [Citation17] and EPIC [Citation18]. There is a pressing need to generate pharmacologic data working towards standardisation amongst the myriad of IP treatment protocols currently applied. Pharmacology of IP chemotherapy can be artificially divided between pharmacokinetics and pharmacodynamics. Whereas pharmacokinetics describes what the body does to the drug, pharmacodynamics looks at what the drug does to the body. Pharmacokinetics of IP chemotherapy studies the alterations between the moment of administration of the IP chemotherapy and the cancer chemotherapy drug showing up at the level of the tumour nodule. Important pharmacokinetic variables include drug dose, volume, duration, carrier solution, pressure and molecular weight. The basic way of depicting pharmacokinetic data is by a concentrations × time graph. Pharmacodynamics subsequently looks into the effect of that cancer chemotherapy drug on the tumour, considering tumour nodule size, density, vascularity, interstitial fluid pressure, binding and temperature. Pharmacodynamic data are depicted in a concentrations × effect graph.

In this manuscript, we review current data regarding drugs, doses, and durations of treatments of IP chemotherapy: standardising HIPEC and EPIC for colorectal, appendiceal, gastric, ovarian PSM and peritoneal mesothelioma.

Selection of chemotherapy drugs for IP administration

Perhaps the most crucial aspect of an optimal IP chemotherapy treatment modality is the selection of a chemotherapy drug for use within the peritoneal space. The ideal drug for IP chemotherapy has a high peritoneal tissue concentration; because of direct IP administration, and a high penetration into the cancer nodule. This should occur in conjunction with slow diffusion of the chemotherapy solution through the peritoneal membrane and deep in the subperitoneal space, resulting in low systemic exposure. The area-under-the-curve (AUC) ratio IP/intravenous (IV) is important in that it quantifies the dose intensity expected in the treatment of PSM. summarises the pharmacologic properties of the chemotherapy drugs most frequently selected for IP application [Citation19]. Pharmacologic variables that should be taken into account are the route of administration, either IP only or IP combined with IV administration, (bidirectional intraoperative chemotherapy (BIC)). The use of naked drugs versus nanoparticles and single drugs versus multiple drugs should also be considered. To select a chemotherapy drug one must know the response expected with this drug in patients with metastatic disease. This emphasises the increasing importance of chemosensitivity testing, towards a patients-tailored approach of selecting the ideal drug for IP and/or IV administration. At present several preclinical work has been conducted in this field using a wide variety of in vitro, in vivo and ex vivo assays using several patient-derived tumour cell-lines in combination with several chemotherapy agents [Citation20,Citation21]. However, important to note is that during the in vitro assays, the 3-D structure of the tumours and hence the important pharmacodynamics of the nodules are lost. Moreover, metabolisation which is very important for the cytotoxic effect of several drugs is not taken into account. Ex vivo assays using patient-derived xenografts and orthotopic animal models also present an impaired view of the clinical situation. For example, implantation of tumour cells subcutaneously, due to differences in microenvironment, will result in the formation of one tumour nodule which fails to progress and metastasise. Further research and careful validation of such assays are needed, taking into account the heterogeneity of tumours and the important pharmacodynamic variables. In the present era of omics assays, gene expression profiling has gained increasing interest in clinical applications to predict oncologic outcomes. Levine et al. analysed gene expression profiles of appendiceal and colorectal PSM samples from patients undergoing HIPEC after complete CRS. They reported distinct genomic signatures for colorectal PSM when compared to appendiceal PSM. Three distinct phenotypes, two consisting of predominantly appendiceal samples (low-risk appendiceal and high-risk appendiceal) and the third with predominately primary colorectal samples (high-risk colorectal), were identified. Furthermore, overall survival (120 months) after optimal CRS and HIPEC was significantly different between the low-risk appendiceal and the high-risk colorectal group [Citation22]. Fujishima et al. used immunohistochemistry to evaluate mucin (MUC) protein expression in tumour nodules of patients with peritoneal dissemination from colorectal cancer as the only synchronous distant metastasis, who had received HIPEC. They report that in patients positive for MUC2 expression the 3-year overall survival rate was 0.0%, whereas in patients negative for MUC2 expression, the 3-year overall survival rate was 61.1% [Citation23]. This emphasises the importance of omics assays to help define better candidates for certain therapies and possibly, in the near future, the choice of chemotherapeutic agents.

Table 1. Overview of pharmacologic properties of the chemotherapy drugs most frequently selected for IP application.

Dosimetry of IP chemotherapy

The current dosing regimens of IP chemotherapy can be divided into body surface area (BSA)-based and concentration-based. Most groups use a drug dose based on calculated BSA (mg/m2) in analogy to systemic chemotherapy regimens. These regimens take BSA as a measure for the effective peritoneal contact area, the peritoneal surface area in the Dedrick formula [Citation24]. The Dedrick formula on itself is an application of Fick’s law of diffusion. Rubin et al. [Citation25] however, demonstrated there is an imperfect correlation between actual peritoneal surface area and calculated BSA. There may also be sex differences in peritoneal surface areas, which in turn affect absorption characteristics. BSA-based IP chemotherapy will result in a fixed dose (BSA-based) diluted in varying volumes of perfusate; i.e.; different concentrations depending on substantial differences in the body composition of patients and differences in the HIPEC technique (open versus closed abdomen). From the Dedrick formula we know that peritoneal concentration and not peritoneal dose is the driving diffusion force [Citation24]. The importance of this has been discussed by Elias et al. [Citation26] in a clinical investigation where 2-, 4-, and 6-litres of chemotherapy solution was administered with a constant dose of chemotherapy solution. A more dilute IP chemotherapy concentration retarded the clearance of chemotherapy and resulted in less systemic toxicity [Citation27]. Therefore, it can be assumed that by the diffusion model, less concentrated chemotherapy would penetrate to a lesser extent into the cancer nodules and normal tissues. On the other hand, concentration-based chemotherapy offers a more predictable exposure of the tumour nodules to the IP chemotherapy and thus efficacy [Citation28]. Unfortunately, the prize to be paid for a better prediction of the efficacy of the IP chemotherapy is a high unpredictability of the plasmatic cancer chemotherapy levels and thus toxicity. Indeed, according to the Dedrick formula of transport over the peritoneal membrane, an increase in the volume of concentration-based IP chemotherapy solution will cause an increase in both diffusion surface and the amount of drug transferred from peritoneal space to plasma [Citation29]. Currently, there is an ongoing study at our hospital evaluating both the pharmacology and morbidity of the different dosing regimens; entitled ‘concentration-based versus body surface area-based peroperative intraperitoneal chemotherapy after optimal cytoreductive surgery in colorectal peritoneal carcinomatosis treatment: randomised non-blinded phase III clinical trial (COBOX trial)’ (https://clinicaltrials.gov/ct2/show/NCT03028155?term=NCT03028155&rank=1). In this pilot study, pharmacologic parameters, the AUC ratio IP/IV and the concentrations in the tumour nodules will be correlated with 3-month overall morbidity and mortality, calculated using the Dindo-Clavien classification. Secondary endpoint is the overall 1-year survival.

Hyperthermic intraperitoneal peroperative chemotherapy (HIPEC)

HIPEC is the most widely explored modality that has consistent clinically improved outcomes in many phase II and III trials [Citation2,Citation3,Citation30–39]. The drugs that are used in this setting are non-cell-cycle specific drugs, which make them applicable for single instillation as in HIPEC () [Citation17].

Table 2. Hyperthermic intraperitoneal peroperative chemotherapy (HIPEC) and bidirectional intraoperative chemotherapy (BIC) regimens.

Cisplatin

Cisplatin (cis-diamminedichloroplatinum-III, CDDP) is an alkylating agent that causes apoptotic cell death by formation of DNA adducts [Citation40]. Both normothermic and hyperthermic IP application have been explored in the treatment of ovarian cancer, gastric cancer, and peritoneal mesothelioma [Citation4,Citation17,Citation41–46]. It is eliminated by renal excretion and consequently the main concern with its use is renal toxicity [Citation47]. Urano et al. showed an excellent in vitro and in vivo thermal augmentation of cisplatin [Citation48]. Current applied cisplatin-based HIPEC regimens are the ‘Sugarbaker Regimen’ [Citation49] and the ‘National Cancer Institute Milan Regimen’ [Citation50].

Oxaliplatin

Oxaliplatin (oxalato-1,2-diaminocyclohexane-platinum(II)) is a third generation platinum complex with proven cytotoxicity in colon and appendiceal neoplasms [Citation51]. In a dose escalation and pharmacokinetic study, Elias et al. demonstrated that 460 mg/m2 of oxaliplatin in 2 L/m2 of chemotherapy solution over 30 min was well tolerated [Citation32,Citation52]. The low AUC ratio is compensated by the rapid absorption of the drug into the tissue, being the reason for the short application time. Oxaliplatin is subject to substantial heat augmentation [Citation48,Citation53]. In a phase I trial, Elias et al. evaluated the pharmacokinetics of heated IP oxaliplatin administered in increasingly hypotonic solutions of 5% dextrose [Citation54]. This trial was based on earlier published experimental data that IP hypotonic solutions increase platinum accumulation in tumour cells [Citation55]. They reported that oxaliplatin clearance from the IP cavity was similar regardless of the osmolarity, but that very hypotonic solutions induce high incidence of IP haemorrhage and thrombocytopenia. As a result of high incidence haemorrhagic complications in another prospective multicenter trial organised by Pomel et al., the dose of oxaliplatin was reduced to 350 mg/m2. However, the incidence of the haemorrhagic complications (29%) did not decrease and the trial was closed prematurely [Citation56]. Chalret du Rieu et al. performed a population pharmacokinetics study, including 75 patients, treated with CRS and oxaliplatin-based HIPEC [Citation57]. They report grade 3/4 thrombocytopenia in 14% of treated patients. Moreover, they concluded that the higher the absorbed dose from the peritoneal cavity, highly dependent on the initial oxaliplatin concentration, the deeper the resultant thrombocytopenia. In an analysis of 701 patients treated with CRS and HIPEC with oxaliplatin or other chemotherapeutic agents, Charrier et al. reported that oxaliplatin-based HIPEC increased the risk of haemorrhagic complications compared to other drugs [Citation58]. In contrast to cisplatin and mitomycin, oxaliplatin traditionally is considered not stable in chloride-containing solutions. This necessitates a dextrose-based carrier which may result in serious electrolyte disturbances and hyperglycaemia during the intracavitary therapy [Citation59]. Unknown to most this degradation of oxaliplatin in normal saline only accounts for less than 10% of the total amount at 30 min; as when applied during HIPEC. Moreover, oxaliplatin degradation was associated with the formation of its active drug form [Pt(dach)Cl2] [Citation60,Citation61]. Different oxaliplatin-based HIPEC regimens are used in current clinical practice: ‘Elias High Dose Oxaliplatin Regimen’ [Citation32], ‘Glehen Medium Dose Oxaliplatin Regimen’ and the ‘Wake Forest University Oxaliplatin Regimen’ [Citation51].

Mitomycin C

Mitomycin C is an alkylating tumour antibiotic extracted from Streptomyces species which most important mechanism of action is through DNA cross-linking. Although mitomycin C is not regarded as a prodrug, it is not active against cancerous tissue as the unchanged molecule. The drug is modified as it enters the cell into an active state [Citation62]. It is inactivated by microsomal enzymes in the liver and is metabolised in the spleen and kidneys. Jacquet et al. reported a clear pharmacokinetic advantage after IP administration with an AUC IP/IV ratio of 23.5 [Citation63]. It is used for PC from colorectal cancer, appendiceal cancer, ovarian cancer, gastric cancer and, for diffuse malignant peritoneal mesothelioma both as HIPEC and EPIC [Citation2,Citation3,Citation29,Citation63–66]. Barlogie et al. suggested in vitro thermal enhancement of mitomycin C [Citation67,Citation68]. Our pharmacokinetic data in 145 HIPEC patients suggest that the largest proportion (62%) of the total drug administered remained in the body at 90 min [Citation29]. This is in line with similar findings by Jacquet et al. and Van Ruth et al. [Citation63,Citation69]. The location and chemical state of this large amount of retained mitomycin C remains to be determined. Controversies still exist regarding the proper dosimetry of the chemotherapy solution. Triple dosing regimen may result in more stable peritoneal levels of the drug throughout the time of IP chemotherapy. Current applied HIPEC dosing regimens are the ‘Sugarbaker Regimen’ [Citation29]. The ‘Duth High Dose Mitomcyin C Regimen: Triple Dosing Regimen’ [Citation70] and the ‘American Society of Peritoneal Surface Malignancy Low Dose Mitomycin C Regimen: Concentration-based Regimen’ [Citation71].

Doxorubicin

Doxorubicin or hydroxyldaunorubicin (adriamycin) is an anthracycline antibiotic. Initial research categorised it as a DNA-intercalating drug. It was later demonstrated that the actual mechanism of action is a temperature-dependent interaction of doxorubicin with the cell surface membrane [Citation72–74]. Doxorubicin was considered a candidate for IP application based on its wide in vitro and in vivo activity against a broad range of malignancies, its slow clearance from the peritoneal compartment due to the high molecular weight of the hydrochloride salt, its favourable AUC ratio of IP to IV concentration times of 230 [Citation17,Citation75–79]. The dose-limiting cardiotoxicity, which is the result of repeated dosing, when administered IV can also be avoided. Pilati et al. suggested a mild hyperthermic augmentation based on increased drug uptake and sensitisation of tumour cells (but not normal mucosal cells) to the cytotoxic effects of doxorubicin [Citation80,Citation81]. More recently PEGylated liposomal doxorubicin has generated interest for HIPEC application due to its favourable pharmacokinetics [Citation82,Citation83]. Doxorubicin-based HIPEC has been used in PSM from appendiceal, gastric, ovarian and colon cancer, as well as in peritoneal mesothelioma [Citation4,Citation84–86].

Bidirectional intraoperative chemotherapy (BIC)

By combining intraoperative IV and intraoperative IP cancer chemotherapy, a bidirectional diffusion gradient is created through the intermediate tissue layer containing the cancer nodules (). In 2002, Elias et al. first reported the clinical use of intraoperative IV 5-fluorouracil and leucovorin in conjunction with oxaliplatin-based HIPEC, to potentiate the effect of oxaliplatin (41). We also reported a clear pharmacokinetic advantage for the intraoperative IV administration of 5-fluorouracil [Citation87]. A similar pharmacokinetic advantage and heat targeting of intraoperative IV ifosfamide was demonstrated [Citation49]. Ifosfamide is a prodrug that needs the cytochrome P450 system of liver or red blood cells to be activated to its active metabolite 4-hydroxyifosfamide. Consequently, it requires IV administration rather than IP instillation for its cytotoxic activity. The drug also shows true heat synergy. It may be an ideal systemic drug to increase the cytotoxicity of HIPEC. Most current protocols advocate bidirectional intraoperative chemotherapy (BIC) (). The bidirectional approach offers the possibility of optimising cancer chemotherapy delivery to the target peritoneal tumour nodules. Further pharmacologic studies are needed to clarify the most efficient method of administration (continuous, bolus or, repeated bolus), doses and, choice of cancer chemotherapy drugs for this bidirectional approach.

Figure 1. Pharmacologic concept of bidirectional intravenous and intraperitoneal chemotherapy. (Credit: Fujiware et al. [Citation139], with permission.)

Figure 1. Pharmacologic concept of bidirectional intravenous and intraperitoneal chemotherapy. (Credit: Fujiware et al. [Citation139], with permission.)

Early postoperative intraperitoneal chemotherapy (EPIC)

EPIC has some conceptual advantages. It is administered shortly after CRS at the time of minimal residual tumour burden. Moreover, IP treatments initiated before wound healing occurs can minimise non-uniform drug distribution and eliminate residual cancer cell entrapment in postoperative fibrin deposits. Disadvantages associated with EPIC are the increased risks of infection and postoperative complications [Citation33,Citation88–90]. EPIC does not involve hyperthermia and is administered postoperatively (typically day 1–day 4/5) through both an inflow catheter and outflow drains inserted at the time of CRS and, can be applied with or without HIPEC [Citation18]. Proper selection of chemotherapy agents based on pharmacologic principles suggests the use of cell-cycle specific drugs such as 5-fluorouracil and the taxanes () [Citation17,Citation91]. This implies administrating multiple cycles, each with a dwell time of around 23 h before renewal. This ensures that all the residual tumour cells are susceptible for the cell cycle specific drug.

Table 3. Early postoperative intraperitoneal chemotherapy (EPIC)-regimens.

5-Fluorouracil

The fluorinated pyrimidines have been successfully used for a wide variety of tumours and, are still an essential component of all successful gastrointestinal cancer chemotherapy regimens [Citation92,Citation93]. This thymidylate synthase inhibitor binds covalently with the enzyme and prevents the formation of thymidine monophosphate, the DNA nucleoside precursor. Also, 5-FU by its metabolites 5-fluoro-uridine diphosphate and 5-fluoro-uridine triphosphate gets incorporated in RNA, resulting in a second cytotoxic pathway. The action of 5-fluorouracil is therefore cell cycle specific. These characteristics limit the use of IP 5-fluorouracil to EPIC [Citation18,Citation94–97]. Minor augmentation of 5-fluorouracil by mild hyperthermia is reported [Citation63]. 5-Fluorouracil is not chemically compatible with other drugs in a mixed solution for infusion or instillation. The current regimens for EPIC with 5-fluorouracil are presented in .

Taxanes

Paclitaxel and docetaxel, with their high molecular weight have a remarkable high AUC ratio of respectively 853 and 861 [Citation19]. The taxanes stabilise the microtubule against depolymerisation; thereby disrupting normal microtubule dynamics [Citation98]. There is evidence supporting additional mechanisms of action [Citation99]. They exert cytotoxic activity against a broad range of tumours. This translates itself into a clear pharmacokinetic advantage for IP administration [Citation100]. The data regarding possible thermal augmentation of taxanes are conflicting [Citation99]. Taxanes have been used in a neoadjuvant intraperitoneal (NIPS) setting as well as intraoperatively and postoperatively. Their cell-cycle specific mechanism of action makes them a better candidate for repetitive application such as in EPIC, NIPS or normothermic adjuvant postoperative IP chemotherapy. Novel formulations of taxanes aiming at an increased bioavailability are under investigation [Citation101]. The current regimens for EPIC with paclitaxel are presented in .

Monoclonal antibodies and avastin

Angiogenesis, the growth of new blood vessels from pre-existing vessels, is paramount for tumour growth and the formation of metastases. It is induced through the production of angiogenic factors by tumour cells [Citation102,Citation103]. A key player in this process is vascular endothelial growth factor-A (VEGF-A), which binds to its receptors VEGFR-1 and VEGFR-2 and thereby increases endothelial cell survival, proliferation, migration and differentiation [Citation104,Citation105]. At present, several targeted molecular therapies are introduced in the treatment of PSM. One of these therapies includes the use of bevacizumab (avastin), a recombinant humanised monoclonal antibody that blocks the activity of VEGF-A. Preclinical work performed by Gremonprez et al. showed that pre-treatment with bevacizumab leads to a significantly lower interstitial fluid pressure in the tumour nodule, which may allow deeper penetration of the IP administered chemotherapeutic agent and higher drug concentrations in the tumour [Citation106]. Several clinical studies have already evaluated the efficacy of IV bevacizumab combined with different chemotherapeutic agents; such as 5-fluorouracil, capecitabine, irinotecan, oxaliplatin, cisplatin and paclitaxel, for the treatment of colorectal and ovarian peritoneal malignancy/ascites [Citation107–112]. It significantly increases the response rate and overall survival for these patients. A recent study, the BEV-IP trial (https://www.clinicaltrials.gov/ct2/show/NCT02399410?term=BEV-IP&rank=1), initiated by Willaert et al., is the first prospective trial that will assess the safety and efficacy of IV bevacizumab followed by CRS and oxaliplatin-based IP chemotherapy in patients diagnosed with synchronous or metachronous colorectal PSM [Citation113]. Attempts have also been made to investigate the potential use of IP bevacizumab as a curative agent. Passot et al. and Chia et al. investigated the IP levels of VEGF at various time points during and after surgery [Citation114,Citation115]. They report that VEGF is present in the peritoneal cavity of patients with PSM treated with curative intent, and its levels increase after CRS. Neoadjuvant bevacizumab significantly decreased the preoperative IP VEGF levels. However, neoadjuvant IV bevacizumab was associated with increased major morbidity [Citation116]. They concluded that the use of preoperative IP bevacizumab for patients with extensive disease burden should be considered, especially in colorectal PSM. Other targeted molecular therapies include the use of drugs that inhibit the endothelial growth factor receptor (EGFR)-related factors to control tumour cell proliferation and differentiation. These drugs include cetuximab and panitumumab [Citation117,Citation118].

Future directions in IP chemotherapy

Neoadjuvant intraperitoneal and systemic chemotherapy (NIPS)

Neoadjuvant bidirectional chemotherapy uses both the IP and IV routes of chemotherapy administration prior to the CRS. It has been suggested as an option for reducing dissemination to the extra-abdominal space, testing the tumour biology and, for reducing the extent of small PC nodules. Theoretically this approach, called neoadjuvant intraperitoneal and systemic chemotherapy (NIPS), may facilitate definitive CRS after initial exploratory laparoscopy or laparotomy [Citation119]. Radiological and clinical responses with NIPS have been reported by several groups [Citation119–122]. However, although NIPS may reduce the tumour load to be addressed by CRS, it has several disadvantages. Adhesions from prior surgical interventions may interfere with adequate IP drug distribution and, as complete responses are unusual, further cytoreduction-chemotherapy is necessary if the approach is to be curative. NIPS is reported to add to morbidity and mortality of further surgical treatment [Citation123]. Furthermore, extensive fibrosis, as a response to chemotherapy, may occur and render judgements concerning the extent of PC difficult or impossible.

Pressurized intraperitoneal aerosol chemotherapy (PIPAC)

Pressurized intraperitoneal aerosol chemotherapy (PIPAC) is a novel approach to deliver IP chemotherapy to patients diagnosed with PSM [Citation124]. During PIPAC, a normothermic capnoperitoneum (pressure of 12 mmHg) is established through a laparoscopic access in an operating room equipped with a laminar airflow. A cytotoxic solution is nebulised into the abdominal cavity during 30 min and thereafter removed through a closed suction system [Citation125]. The hypothesis underlying this technique is that intraabdominal application of chemotherapy under pressure will enhance tumour drug uptake and aerosolizing and spraying chemotherapy will enhance the area of peritoneal surface covered by the drug.

Several experimental and clinical studies have been conducted to test the above-mentioned hypothesis [Citation125–129]. Solass et al. performed PIPAC with cisplatin and doxorubicin in 3 end-staged patients with advanced PC of gastric, appendiceal and ovarian origin. They report that PIPAC required only 1/10 of the doxorubicin dose to achieve higher tumour concentrations as compared to HIPEC. High tissue concentrations of doxorubicin were reported. Moreover, fluorescence microscopy showed nuclear presence of doxorubicin throughout the whole peritoneal layer and up to deeply into the retroperitoneal fatty tissue. They concluded that PIPAC was well tolerated with excellent local exposure and low-systemic exposure [Citation127]. Moreover, PIPAC appeared to be associated with very limited hepatic and renal toxicity even after repeated PIPAC [Citation130,Citation131]. On the other hand, Khosrawipour et al. reported that the depth of doxorubicin penetration was significantly higher in tissues directly exposed to the aerosol jet when compared to the side wall, in an ex vivo PIPAC model [Citation132]. In a phase II study conducted by Tempfer et al., 64 patients with recurrent ovarian, fallopian or peritoneal cancer with PSM were treated with three courses doxorubicin and cisplatin-based PIPAC. PIPAC was well tolerated, easy to perform and associated with a better quality of life as compared to systemic chemotherapy, with the absence of grade 4 toxicities [Citation129]. Demtröder et al. performed a retrospective analysis including 17 patients with pre-treated (surgery alone or combined with systemic chemotherapy) colorectal peritoneal metastases, who had received up to 6 cycles of oxaliplatin based PIPAC. Repeated PIPAC with oxaliplatin could induce regression of the peritoneal metastases, with low toxicities [Citation133]. However, it should be taken into account that patients included in these trials are highly selected and often have had extensive surgery and were already heavily pre-treated with several lines of systemic chemotherapy. The potential limited access of the aerosolized chemotherapy due to the presence of adhesions is not taken into account. Moreover, incomplete responses warrants further cytoreduction. However, it has been reported that PIPAC should not be combined with CRS due to the potential of increased local toxicity [Citation134]. Recently, Kakchekeeva et al. introduced electrostatic PIPAC (ePIPAC), hypothesising that electrostatic charging the aerosol particles may further enhance the pharmacologic properties of PIPAC [Citation135]. They performed a comparative study of PIPAC and ePIPAC assessing the pharmacologic properties using an in vivo porcine model. They reported that ePIPAC has the potential to allow more efficient drug uptake, further dose reduction, a significant shortening of the time required for PIPAC application, further improving health and safety measures.

Today, there are no phase III trial data available for PIPAC emphasising that this is still an experimental treatment, which should be further investigated within the context of controlled clinical trials. These data will be important in identifying the role of PIPAC in the treatment of PSM patients. Today, PIPAC can play a role as a new palliative treatment option in highly selected patients with PSM.

Drug delivery systems

As was previously mentioned, the ideal drug for IP chemotherapy should have a high-peritoneal tissue concentration and this should occur in conjunction with slow diffusion of the chemotherapy solution through the peritoneal membrane and deep in the subperitoneal space. However, today there are no drugs specifically designed for IP use. Therefore, over the past years, a lot of research has been focussing on the use of drug delivery systems to optimise IP drug delivery and to prolong the residence time of the drug in the peritoneal cavity with minimal systemic toxicity. These delivery systems include microspheres, nanoparticles, liposomes, micelles, injectable systems and implantable systems [Citation101,Citation136]. In a preclinical study, De Smet et al. reported the development of a stable nanocrystalline paclitaxel formulation which was of interest for the treatment of ovarian PSM via HIPEC [Citation137]. Xu et al. designed a thermosensitive injectable drug delivery hydrogel assembled by paclitaxel-incorporated nanoparticles with an improved bioavailability and induced effective antitumor efficacy in a colorectal PSM mouse model [Citation138]. Thermosensitive hydrogels can transfer from free-flowing sol to a gel at physiological temperature and are interesting candidates for sustained drug delivery.

Conclusion

The combination of CRS and IP chemotherapy should now be considered standard of care for PSM from appendiceal epithelial cancers, colorectal cancer and peritoneal mesothelioma. Although there is a near universal standardisation regarding the CRS, there is still a much-needed standardisation amongst the various IP chemotherapy treatment modalities used today in clinical practice. Although today, trends in the IP protocols, concerning the reduced dosing of oxaliplatin and the triple dosing regimen of mitomycin C are observed; pharmacologic evidence should be provided to answer important questions raised by the myriad of variables associated with IP chemotherapy. Furthermore, new and innovative IP chemotherapy concepts, like PIPAC, should be investigated in well-designed and adequately powered phase III clinical trials.

Disclosure statement

The authors report no conflicts of interest. The authors alone are responsible for the content and writing of the article.

References

  • Mirnezami R, Mehta AM, Chandrakumaran K, et al. (2014). Cytoreductive surgery in combination with hyperthermic intraperitoneal chemotherapy improves survival in patients with colorectal peritoneal metastases compared with systemic chemotherapy alone. Br J Cancer 111:1500–8.
  • Verwaal VJ, Bruin S, Boot H, et al. (2008). 8-year follow-up of randomized trial: cytoreduction and hyperthermic intraperitoneal chemotherapy versus systemic chemotherapy in patients with peritoneal carcinomatosis of colorectal cancer. Ann Surg Oncol 2008;15:2426–32.
  • Verwaal VJ, van Ruth S, de Bree E, et al. (2003). 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 21:3737–43.
  • Baratti D, Kusamura S, Cabras AD, et al. (2013). Diffuse malignant peritoneal mesothelioma: long-term survival with complete cytoreductive surgery followed by hyperthermic intraperitoneal chemotherapy (HIPEC). Eur J Cancer 49:3140–8.
  • Spiliotis J, Halkia E, Lianos E, et al. (2015) Cytoreductive surgery and HIPEC in recurrent epithelial ovarian cancer: a prospective randomized phase III study. Ann Surg Oncol 22:1570–5.
  • Marcotte E, Dube P, Drolet P, et al. (2014). Hyperthermic intraperitoneal chemotherapy with oxaliplatin as treatment for peritoneal carcinomatosis arising from the appendix and pseudomyxoma peritonei: a survival analysis. World J Surg Oncol 12:332.
  • Sideris L, Mitchell A, Drolet P, et al. (2009). Surgical cytoreduction and intraperitoneal chemotherapy for peritoneal carcinomatosis arising from the appendix. Can J Surg 52:135–41.
  • Yan TD, Deraco M, Baratti D, et al. (2009). Cytoreductive surgery and hyperthermic intraperitoneal chemotherapy for malignant peritoneal mesothelioma: multi-institutional experience. J Clin Oncol 27:6237–42.
  • Glehen O, Gilly FN, Arvieux C, et al. (2010). Peritoneal carcinomatosis from gastric cancer: a multi-institutional study of 159 patients treated by cytoreductive surgery combined with perioperative intraperitoneal chemotherapy. Ann Surg Oncol 17:2370–7.
  • Gonzalez Bayon L, Steiner MA, Vasquez Jimenez W, et al. (2013). Cytoreductive surgery and hyperthermic intraperitoneal chemotherapy for the treatment of advanced epithelial ovarian carcinoma: upfront therapy, at first recurrence, or later? Eur J Surg Oncol 39:1109–15.
  • Glehen O, Gilly FN, Boutitie F, et al. (2010). Toward curative treatment of peritoneal carcinomatosis from nonovarian origin by cytoreductive surgery combined with perioperative intraperitoneal chemotherapy: a multi-institutional study of 1,290 patients. Cancer 116:5608–18.
  • Sugarbaker PH. (2006). New standard of care for appendiceal epithelial neoplasms and pseudomyxoma peritonei syndrome? Lancet Oncol 7:69–76.
  • Esquivel J, Piso P, Verwaal V, et al. (2014). American Society of Peritoneal Surface Malignancies opinion statement on defining expectations from cytoreductive surgery and hyperthermic intraperitoneal chemotherapy in patients with colorectal cancer. J Surg Oncol 110:777–8.
  • Bakrin N, Bereder JM, Decullier E, et al. (2013). Peritoneal carcinomatosis treated with cytoreductive surgery and Hyperthermic Intraperitoneal Chemotherapy (HIPEC) for advanced ovarian carcinoma: a French multicentre retrospective cohort study of 566 patients. Eur J Surg Oncol 39:1435–43.
  • Sugarbaker PH. (2012). Parietal peritonectomy. Ann Surg Oncol 19:1250
  • Sugarbaker P. (2012). Cytoreductive surgery & perioperative chemotherapy for peritoneal surface malignancy: textbook and video atlas. Woodbury, CT: Ciné-Med Publishing. p. 214.
  • Van der Speeten K, Stuart OA, Sugarbaker PH. (2012). Pharmacology of perioperative intraperitoneal and intravenous chemotherapy in patients with peritoneal surface malignancy. Surg Oncol Clin N Am 21:577–97.
  • Sugarbaker PH, Graves T, DeBruijn EA, et al. (1990). Early postoperative intraperitoneal chemotherapy as an adjuvant therapy to surgery for peritoneal carcinomatosis from gastrointestinal cancer: pharmacological studies. Cancer Res 50:5790–4.
  • Sugarbaker PH, Mora JT, Carmignani P, et al. (2005). Update on chemotherapeutic agents utilized for perioperative intraperitoneal chemotherapy. Oncologist 10:112–22.
  • Mahteme H, von Heideman A, Grundmark B, et al. (2008). Heterogeneous activity of cytotoxic drugs in patient samples of peritoneal carcinomatosis. Eur J Surg Oncol 34:547–52.
  • Cashin PH, Mahteme H, Graf W, et al. (2013). Activity ex vivo of cytotoxic drugs in patient samples of peritoneal carcinomatosis with special focus on colorectal cancer. BMC Cancer 13:435.
  • Levine EA, Blazer DG III, Kim MK, et al. (2012). Gene expression profiling of peritoneal metastases from appendiceal and colon cancer demonstrates unique biologic signatures and predicts patient outcomes. J Am Coll Surg 214:599–606. Discussion 7.
  • Fujishima Y, Goi T, Kimura Y, et al. (2012). MUC2 protein expression status is useful in assessing the effects of hyperthermic intraperitoneal chemotherapy for peritoneal dissemination of colon cancer. Int J Oncol 40:960–4.
  • Dedrick RL, Myers CE, Bungay PM, DeVita VT Jr. (1978). Pharmacokinetic rationale for peritoneal drug administration in the treatment of ovarian cancer. Cancer Treat Rep 62:1–11.
  • Rubin J, Clawson M, Planch A, Jones Q. (1988). Measurements of peritoneal surface area in man and rat. Am J Med Sci 295:453–8.
  • Elias DM, Sideris L. (2003). Pharmacokinetics of heated intraoperative intraperitoneal oxaliplatin after complete resection of peritoneal carcinomatosis. Surg Oncol Clin N Am 12:755–69.
  • Sugarbaker PH, Stuart OA, Carmignani CP. (2006) Pharmacokinetic changes induced by the volume of chemotherapy solution in patients treated with hyperthermic intraperitoneal mitomycin C. Cancer Chemother Pharmacol 57:703–8.
  • Mas-Fuster MI, Ramon-Lopez A, Nalda-Molina R. (2013). Importance of standardizing the dose in hyperthermic intraperitoneal chemotherapy (HIPEC): a pharmacodynamic point of view. Cancer Chemother Pharmacol 72:273–4.
  • Van der Speeten K, Stuart OA, Chang D, et al. (2011). Changes induced by surgical and clinical factors in the pharmacology of intraperitoneal mitomycin C in 145 patients with peritoneal carcinomatosis. Cancer Chemother Pharmacol 68:147–56.
  • Elias D, Lefevre JH, Chevalier J, et al. (2009). Complete cytoreductive surgery plus intraperitoneal chemohyperthermia with oxaliplatin for peritoneal carcinomatosis of colorectal origin. J Clin Oncol 27:681–5.
  • Franko J, Ibrahim Z, Gusani NJ, et al. (2010). Cytoreductive surgery and hyperthermic intraperitoneal chemoperfusion versus systemic chemotherapy alone for colorectal peritoneal carcinomatosis. Cancer 116:3756–62.
  • Elias D, Bonnay M, Puizillou JM, et al. (2002). Heated intra-operative intraperitoneal oxaliplatin after complete resection of peritoneal carcinomatosis: pharmacokinetics and tissue distribution. Ann Oncol 13:267–72.
  • Glehen O, Kwiatkowski F, Sugarbaker PH, et al. (2004). Cytoreductive surgery combined with perioperative intraperitoneal chemotherapy for the management of peritoneal carcinomatosis from colorectal cancer: a multi-institutional study. J Clin Oncol 22:3284–92.
  • Yan TD, Black D, Savady R, Sugarbaker PH. (2006). Systematic review on the efficacy of cytoreductive surgery combined with perioperative intraperitoneal chemotherapy for peritoneal carcinomatosis from colorectal carcinoma. J Clin Oncol 24:4011–19.
  • Yan TD, Welch L, Black D, Sugarbaker PH. (2007). A systematic review on the efficacy of cytoreductive surgery combined with perioperative intraperitoneal chemotherapy for diffuse malignancy peritoneal mesothelioma. Ann Oncol 18:827–34.
  • Yan TD, Black D, Sugarbaker PH, et al. (2007). A systematic review and meta-analysis of the randomized controlled trials on adjuvant intraperitoneal chemotherapy for resectable gastric cancer. Ann Surg Oncol 14:2702–13.
  • Bijelic L, Jonson A, Sugarbaker PH. (2007). Systematic review of cytoreductive surgery and heated intraoperative intraperitoneal chemotherapy for treatment of peritoneal carcinomatosis in primary and recurrent ovarian cancer. Ann Oncol 18:1943–50.
  • Helm CW, Richard SD, Pan J, et al. (2010). Hyperthermic intraperitoneal chemotherapy in ovarian cancer: first report of the HYPER-O registry. Int J Gynecol Cancer 20:61–9.
  • Klaver CE, Musters GD, Bemelman WA, et al. (2015). Adjuvant hyperthermic intraperitoneal chemotherapy (HIPEC) in patients with colon cancer at high risk of peritoneal carcinomatosis; the COLOPEC randomized multicentre trial. BMC Cancer 15:428.
  • Los G, Mutsaers PH, van der Vijgh WJ, et al. (1989). Direct diffusion of cis-diamminedichloroplatinum(II) in intraperitoneal rat tumors after intraperitoneal chemotherapy: a comparison with systemic chemotherapy. Cancer Res 49:3380–4.
  • Conti M, De Giorgi U, Tazzari V, et al. (2004). Clinical pharmacology of intraperitoneal cisplatin-based chemotherapy. J Chemother 16(Suppl 5):23–5.
  • Gladieff L, Chatelut E, Dalenc F, Ferron G. (2009). [Pharmacological bases of intraperitoneal chemotherapy]. Bull Cancer 96:1235–42.
  • Coccolini F, Cotte E, Glehen O, et al. (2014). Intraperitoneal chemotherapy in advanced gastric cancer. Meta-analysis of randomized trials. Eur J Surg Oncol 40:12–26.
  • Alberts DS, Liu PY, Hannigan EV, et al. (1996). Intraperitoneal cisplatin plus intravenous cyclophosphamide versus intravenous cisplatin plus intravenous cyclophosphamide for stage III ovarian cancer. N Engl J Med 335:1950–5.
  • Armstrong DK, Bundy B, Wenzel L, et al. (2006). Intraperitoneal cisplatin and paclitaxel in ovarian cancer. N Engl J Med 354:34–43.
  • Zivanovic O, Abramian A, Kullmann M, et al. (2014). HIPEC ROC I: A phase i study of cisplatin administered as hyperthermic intraoperative intraperitoneal chemoperfusion followed by postoperative intravenous platinum-based chemotherapy in patients with platinum-sensitive recurrent epithelial ovarian cancer. Int J Cancer 136:699–708.
  • Hakeam HA, Breakiet M, Azzam A, et al. (2014). The incidence of cisplatin nephrotoxicity post hyperthermic intraperitoneal chemotherapy (HIPEC) and cytoreductive surgery. Ren Fail 36:1486–91.
  • Urano M, Kuroda M, Nishimura Y. (1999). For the clinical application of thermochemotherapy given at mild temperatures. Int J Hyperthermia 15:79–107.
  • Van der Speeten K, Stuart OA, Mahteme H, Sugarbaker PH. (2011). Pharmacokinetic study of perioperative intravenous Ifosfamide. Int J Surg Oncol 2011:185092.
  • Deraco M, Baratti D, Cabras AD, et al. (2010). Experience with peritoneal mesothelioma at the Milan National Cancer Institute. World J Gastrointest Oncol 2:76–84.
  • Stewart JHt, Shen P, Russell G, et al. (2008). A phase I trial of oxaliplatin for intraperitoneal hyperthermic chemoperfusion for the treatment of peritoneal surface dissemination from colorectal and appendiceal cancers. Ann Surg Oncol 15:2137–45.
  • Elias D, Raynard B, Bonnay M, Pocard M. (2006). Heated intra-operative intraperitoneal oxaliplatin alone and in combination with intraperitoneal irinotecan: Pharmacologic studies. Eur J Surg Oncol 32:607–13.
  • Piche N, Leblond FA, Sideris L, et al. (2011). Rationale for heating oxaliplatin for the intraperitoneal treatment of peritoneal carcinomatosis: a study of the effect of heat on intraperitoneal oxaliplatin using a murine model. Ann Surg 254:138–44.
  • Elias D, El Otmany A, Bonnay M, et al. (2002). Human pharmacokinetic study of heated intraperitoneal oxaliplatin in increasingly hypotonic solutions after complete resection of peritoneal carcinomatosis. Oncology 63:346–52.
  • Smith E, Brock AP. (1989). The effect of reduced osmolarity on platinum drug toxicity. Br J Cancer 59:873–5.
  • Pomel C, Ferron G, Lorimier G, et al. (2010). Hyperthermic intra-peritoneal chemotherapy using oxaliplatin as consolidation therapy for advanced epithelial ovarian carcinoma. Results of a phase II prospective multicentre trial. CHIPOVAC study. Eur J Surg Oncol 36:589–93.
  • Chalret du Rieu Q, White-Koning M, Picaud L, et al. (2014). Population pharmacokinetics of peritoneal, plasma ultrafiltrated and protein-bound oxaliplatin concentrations in patients with disseminated peritoneal cancer after intraperitoneal hyperthermic chemoperfusion of oxaliplatin following cytoreductive surgery: correlation between oxaliplatin exposure and thrombocytopenia. Cancer Chemother Pharmacol 74:571–82.
  • Charrier T, Passot G, Peron J, et al. (2016). Cytoreductive surgery combined with hyperthermic intraperitoneal chemotherapy with oxaliplatin increases the risk of postoperative hemorrhagic complications: analysis of predictive factors. Ann Surg Oncol 23:2315–22.
  • De Somer F, Ceelen W, Delanghe J, et al. (2008). Severe hyponatremia, hyperglycemia, and hyperlactatemia are associated with intraoperative hyperthermic intraperitoneal chemoperfusion with oxaliplatin. Perit Dial Int 28:61–6.
  • Jerremalm E, Hedeland M, Wallin I, et al. (2004). Oxaliplatin degradation in the presence of chloride: identification and cytotoxicity of the monochloro monooxalato complex. Pharm Res 21:891–4.
  • Mehta AM, Van den Hoven JM, Rosing H, et al. (2015). Stability of oxaliplatin in chloride-containing carrier solutions used in hyperthermic intraperitoneal chemotherapy. Int J Pharm 479:23–7.
  • Bachur NR, Gordon SL, Gee MV, Kon H. (1979). NADPH cytochrome P-450 reductase activation of quinone anticancer agents to free radicals. Proc Natl Acad Sci USA 76:954–7.
  • Jacquet P, Averbach A, Stephens AD, et al. (1998). Heated intraoperative intraperitoneal mitomycin C and early postoperative intraperitoneal 5-fluorouracil: pharmacokinetic studies. Oncology 55:130–8.
  • Fujita T, Tamura T, Yamada H, et al. (1997). Pharmacokinetics of mitomycin C (MMC) after intraperitoneal administration of MMC-gelatin gel and its anti-tumor effects against sarcoma-180 bearing mice. J Drug Target 4:289–96.
  • Mohamed F, Cecil T, Moran B, Sugarbaker P. (2011). A new standard of care for the management of peritoneal surface malignancy. Curr Oncol 18:e84–96.
  • Levine EA, Stewart JHt, Shen P, et al. (2014). Intraperitoneal chemotherapy for peritoneal surface malignancy: experience with 1,000 patients. J Am Coll Surg 218:573–85.
  • Barlogie B, Corry PM, Drewinko B. (1980). In vitro thermochemotherapy of human colon cancer cells with cis-dichlorodiammineplatinum(II) and mitomycin C. Cancer Res 40:1165–8.
  • de Bree E, Tsiftsis DD. (2007). Experimental and pharmacokinetic studies in intraperitoneal chemotherapy: from laboratory bench to bedside recent results. Cancer Res 169:53–73.
  • van Ruth S, Verwaal VJ, Zoetmulder FA. (2003). Pharmacokinetics of intraperitoneal mitomycin C. Surg Oncol Clin N Am 12:771–80.
  • Witkamp A. (1998). Dose finding study of hyperthermic intraperitoneal chemotherapy with mitomycin C in patients with carcinosis of colorectal origin. Eur J Surg Oncol 24:214.
  • Turaga K, Levine E, Barone R, et al. (2014). Consensus guidelines from The American Society of Peritoneal Surface Malignancies on standardizing the delivery of hyperthermic intraperitoneal chemotherapy (HIPEC) in colorectal cancer patients in the United States. Ann Surg Oncol 21:1501–5.
  • Tritton TR. (1991). Cell surface actions of adriamycin. Pharmacol Ther 49:293–309.
  • Triton TR, Yee G. (1982). The anticancer agent adriamycin can be actively cytotoxic without entering cells. Science 217:248–50.
  • Lane P, Vichi P, Bain DL, Tritton TR. (1987). Temperature dependence studies of adriamycin uptake and cytotoxicity. Cancer Res 47:4038–42.
  • Van der Speeten K, Stuart OA, Mahteme H, Sugarbaker PH. (2009). A pharmacologic analysis of intraoperative intracavitary cancer chemotherapy with doxorubicin. Cancer Chemother Pharmacol 63:799–805.
  • Ozols RF, Young RC, Speyer JL, et al. (1982). Phase I and pharmacological studies of adriamycin administered intraperitoneally to patients with ovarian cancer. Cancer Res 42:4265–9.
  • Ozols RF, Locker GY, Doroshow JH, et al. (1979). Pharmacokinetics of adriamycin and tissue penetration in murine ovarian cancer. Cancer Res 39:3209–14.
  • Ozols RF, Locker GY, Doroshow JH, et al. (1979). Chemotherapy for murine ovarian cancer: a rationale for ip therapy with adriamycin. Cancer Treat Rep 63:269–73.
  • Nagai K, Nogami S, Egusa H, Konishi H. (2014). Pharmacokinetic evaluation of intraperitoneal doxorubicin in rats. Pharmazie 69:125–7.
  • Pilati P, Mocellin S, Rossi CR, et al. (2003). Doxorubicin activity is enhanced by hyperthermia in a model of ex vivo vascular perfusion of human colon carcinoma. World J Surg 27:640–6.
  • Rossi CR, Foletto M, Mocellin S, et al. (2002). Hyperthermic intraoperative intraperitoneal chemotherapy with cisplatin and doxorubicin in patients who undergo cytoreductive surgery for peritoneal carcinomatosis and sarcomatosis: phase I study. Cancer 94:492–9.
  • Harrison LE, Bryan M, Pliner L, Saunders T. (2008). Phase I trial of pegylated liposomal doxorubicin with hyperthermic intraperitoneal chemotherapy in patients undergoing cytoreduction for advanced intra-abdominal malignancy. Ann Surg Oncol 15:1407–13.
  • Salvatorelli E, De Tursi M, Menna P, et al. (2012). Pharmacokinetics of pegylated liposomal doxorubicin administered by intraoperative hyperthermic intraperitoneal chemotherapy to patients with advanced ovarian cancer and peritoneal carcinomatosis. Drug Metab Dispos 40:2365–73.
  • Moran B, Cecil T, Chandrakumaran K, et al. (2015). The results of cytoreductive surgery and hyperthermic intraperitoneal chemotherapy in 1200 patients with peritoneal malignancy. Colorectal Dis 17:772–8.
  • Sugarbaker PH, Van der Speeten K, Anthony Stuart O, Chang D. (2011). Impact of surgical and clinical factors on the pharmacology of intraperitoneal doxorubicin in 145 patients with peritoneal carcinomatosis. Eur J Surg Oncol 37:719–26.
  • Ansaloni L, Agnoletti V, Amadori A, et al. (2012). Evaluation of extensive cytoreductive surgery and hyperthermic intraperitoneal chemotherapy (HIPEC) in patients with advanced epithelial ovarian cancer. Int J Gynecol Cancer 22:778–85.
  • Van der Speeten K, Stuart OA, Mahteme H, Sugarbaker PH. (2010). Pharmacology of perioperative 5-fluorouracil. J Surg Oncol 102:730–5.
  • Elias D, Benizri E, Di Pietrantonio D, et al. (2007). Comparison of two kinds of intraperitoneal chemotherapy following complete cytoreductive surgery of colorectal peritoneal carcinomatosis. Ann Surg Oncol 14:509–14.
  • McConnell YJ, Mack LA, Francis WP, et al. (2013). HIPEC + EPIC versus HIPEC-alone: differences in major complications following cytoreduction surgery for peritoneal malignancy. J Surg Oncol;107:591–6.
  • Tan GH, Ong WS, Chia CS, et al. (2016). Does early post-operative intraperitoneal chemotherapy (EPIC) for patients treated with cytoreductive surgery and hyperthermic intraperitoneal chemotherapy (HIPEC) make a difference? Int J Hyperthermia 32:281–8.
  • Sugarbaker PH, Welch LS, Mohamed F, Glehen O. (2003). A review of peritoneal mesothelioma at the Washington Cancer Institute. Surg Oncol Clin N Am 12:605–21.
  • Heidelberger C, Chaudhuri NK, Danneberg P, et al. (1957). Fluorinated pyrimidines, a new class of tumour-inhibitory compounds. Nature 179:663–6.
  • Muggia FM, Peters GJ, Landolph JR. Jr. (2009). XIII International Charles Heidelberger Symposium and 50 Years of Fluoropyrimidines in Cancer Therapy Held on september 6 to 8, 2007 at New York University Cancer Institute, Smilow Conference Center. Mol Cancer Ther 8:992–9.
  • Wagner PL, Jones D, Aronova A, et al. (2012). Early postoperative intraperitoneal chemotherapy following cytoreductive surgery for appendiceal mucinous neoplasms with isolated peritoneal metastasis. Dis Colon Rectum 55:407–15.
  • Yu W, Whang I, Chung HY, et al. (2001). Indications for early postoperative intraperitoneal chemotherapy of advanced gastric cancer: results of a prospective randomized trial. World J Surg 25:985–90.
  • Kwon OK, Chung HY, Yu W. (2014). Early postoperative intraperitoneal chemotherapy for macroscopically serosa-invading gastric cancer patients. Cancer Res Treat 46:270–9.
  • Passot G, Vaudoyer D, Villeneuve L, et al. (2016). What made hyperthermic intraperitoneal chemotherapy an effective curative treatment for peritoneal surface malignancy: a 25-year experience with 1,125 procedures. J Surg Oncol 113:796–803.
  • Rohena CC, Mooberry SL. (2014). Recent progress with microtubule stabilizers: new compounds, binding modes and cellular activities. Nat Prod Rep 31:335–55.
  • de Bree E, Theodoropoulos PA, Rosing H, et al. (2006). Treatment of ovarian cancer using intraperitoneal chemotherapy with taxanes: from laboratory bench to bedside. Cancer Treat Res 32:471–82.
  • Mohamed F, Sugarbaker PH. (2003). Intraperitoneal taxanes. Surg Oncol Clin N Am 12:825–33.
  • De Smet L, Ceelen W, Remon JP, Vervaet C. (2013). Optimization of drug delivery systems for intraperitoneal therapy to extend the residence time of the chemotherapeutic agent. Scientific World J 2013:720858.
  • Jacquet P, Sugarbaker PH. (1996). Peritoneal-plasma barrier. Cancer Treat Res 82:53–63.
  • Hanahan D, Weinberg RA. (2011). Hallmarks of cancer: the next generation. Cell 144:646–74.
  • Lee S, Chen TT, Barber CL, et al. (2007). Autocrine VEGF signaling is required for vascular homeostasis. Cell 130:691–703.
  • De Falco S, Gigante B, Persico MG. (2002). Structure and function of placental growth factor. Trends Cardiovasc Med 12:241–6.
  • Gremonprez F, Descamps B, Izmer A, et al. (2015). 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 6:29889–900.
  • Kesterson JP, Mhawech-Fauceglia P, Lele S. (2008). The use of bevacizumab in refractory ovarian granulosa-cell carcinoma with symptomatic relief of ascites: a case report. Gynecol Oncol 111:527–9.
  • Hamilton CA, Maxwell GL, Chernofsky MR, et al. (2008). Intraperitoneal bevacizumab for the palliation of malignant ascites in refractory ovarian cancer. Gynecol Oncol 111:530–2.
  • Choe JH, Overman MJ, Fournier KF, et al. (2015). Improved survival with anti-VEGF therapy in the treatment of unresectable appendiceal epithelial neoplasms. Ann Surg Oncol 22:2578–84.
  • Botrel TE, Clark LG, Paladini L, Clark OA. (2016). Efficacy and safety of bevacizumab plus chemotherapy compared to chemotherapy alone in previously untreated advanced or metastatic colorectal cancer: a systematic review and meta-analysis. BMC Cancer 16:677.
  • Oliva P, Decio A, Castiglioni V, et al. (2012). Cisplatin plus paclitaxel and maintenance of bevacizumab on tumour progression, dissemination, and survival of ovarian carcinoma xenograft models. Br J Cancer 107:360–9.
  • Saltz LB, Clarke S, Diaz-Rubio E, et al. (2008). Bevacizumab in combination with oxaliplatin-based chemotherapy as first-line therapy in metastatic colorectal cancer: a randomized phase III study. J Clin Oncol 26:2013–19.
  • Willaert W, Van Der Speeten K, Liberale G, Ceelen W. (2015). BEV-IP: Perioperative chemotherapy with bevacizumab in patients undergoing cytoreduction and intraperitoneal chemoperfusion for colorectal carcinomatosis. BMC Cancer 15:980.
  • Passot G, Bakrin N, Garnier L, et al. (2014). Intraperitoneal vascular endothelial growth factor burden in peritoneal surface malignancies treated with curative intent: the first step before intraperitoneal anti-vascular endothelial growth factor treatment? Eur J Cancer 50:722–30.
  • Chia CS, Glehen O, Bakrin N, et al. (2015). Intraperitoneal vascular endothelial growth factor: a prognostic factor and the potential for intraperitoneal bevacizumab use in peritoneal surface malignancies. Ann Surg Oncol 22(Suppl 3):S880–7.
  • Eveno C, Passot G, Goere D, et al. (2014). Bevacizumab doubles the early postoperative complication rate after cytoreductive surgery with hyperthermic intraperitoneal chemotherapy (HIPEC) for peritoneal carcinomatosis of colorectal origin. Ann Surg Oncol 21:1792–800.
  • Abdel-Rahman O, Fouad M. (2015). Correlation of cetuximab-induced skin rash and outcomes of solid tumor patients treated with cetuximab: a systematic review and meta-analysis. Crit Rev Oncol Hematol 93:127–35.
  • Pettigrew M, Kavan P, Surprenant L, Lim HJ. (2016). Comparative net cost impact of the utilization of panitumumab versus cetuximab for the treatment of patients with metastatic colorectal cancer in Canada. J Med Econ 19:135–47.
  • Yonemura Y, Bandou E, Sawa T, et al. (2006). Neoadjuvant treatment of gastric cancer with peritoneal dissemination. Eur J Surg Oncol 32:661–5.
  • Sugarbaker PH. (1996). Treatment of peritoneal carcinomatosis from colon or appendiceal cancer with induction intraperitoneal chemotherapy. Cancer Treat Res 82:317–25.
  • Zylberberg B, Dormont D, Janklewicz S, et al. (2001). Response to neo-adjuvant intraperitoneal and intravenous immunochemotherapy followed by interval secondary cytoreduction in stage IIIc ovarian cancer. Eur J Gynaecol Oncol 22:40–5.
  • Yonemura Y, Elnemr A, Endou Y, et al. (2012). Effects of neoadjuvant intraperitoneal/systemic chemotherapy (bidirectional chemotherapy) for the treatment of patients with peritoneal metastasis from gastric cancer. Int J Surg Oncol 2012:148420.
  • Esquivel J, Vidal-Jove J, Steves MA, Sugarbaker PH. (1993). Morbidity and mortality of cytoreductive surgery and intraperitoneal chemotherapy. Surgery 113:631–6.
  • Tempfer CB. (2015). Pressurized intraperitoneal aerosol chemotherapy as an innovative approach to treat peritoneal carcinomatosis. Med Hypotheses 85:480–4.
  • Solass W, Hetzel A, Nadiradze G, et al. (2012). Description of a novel approach for intraperitoneal drug delivery and the related device. Surg Endosc 26:1849–55.
  • Solass W, Herbette A, Schwarz T, et al. (2012). Therapeutic approach of human peritoneal carcinomatosis with Dbait in combination with capnoperitoneum: proof of concept. Surg Endosc 26:847–52.
  • Solass W, Kerb R, Murdter T, et al. (2014). Intraperitoneal chemotherapy of peritoneal carcinomatosis using pressurized aerosol as an alternative to liquid solution: first evidence for efficacy. Ann Surg Oncol 21:553–9.
  • Solass W, Giger-Pabst U, Zieren J, Reymond MA. (2013). Pressurized intraperitoneal aerosol chemotherapy (PIPAC): occupational health and safety aspects. Ann Surg Oncol 20:3504–11.
  • Tempfer CB, Winnekendonk G, Solass W, et al. (2015). Pressurized intraperitoneal aerosol chemotherapy in women with recurrent ovarian cancer: a phase 2 study. Gynecol Oncol 137:223–8.
  • Blanco A, Giger-Pabst U, Solass W, et al. (2013). Renal and hepatic toxicities after pressurized intraperitoneal aerosol chemotherapy (PIPAC). Ann Surg Oncol 20:2311–16.
  • Robella M, Vaira M, De Simone M. (2016). Safety and feasibility of pressurized intraperitoneal aerosol chemotherapy (PIPAC) associated with systemic chemotherapy: an innovative approach to treat peritoneal carcinomatosis. World J Surg Oncol 14:128.
  • Khosrawipour V, Khosrawipour T, Diaz-Carballo D, et al. (2016). Exploring the spatial drug distribution Pattern of Pressurized Intraperitoneal Aerosol Chemotherapy (PIPAC). Ann Surg Oncol 23:1220–4.
  • Demtroder C, Solass W, Zieren J, et al. (2016). Pressurized intraperitoneal aerosol chemotherapy with oxaliplatin in colorectal peritoneal metastasis. Colorectal Dis 18:364–71.
  • Tempfer CB, Solass W, Buerkle B, Reymond MA. (2014). Pressurized intraperitoneal aerosol chemotherapy (PIPAC) with cisplatin and doxorubicin in a woman with pseudomyxoma peritonei: A case report. Gynecol Oncol Rep 10:32–5.
  • Kakchekeeva T, Demtroder C, Herath NI, et al. (2016). In vivo feasibility of electrostatic precipitation as an adjunct to Pressurized Intraperitoneal Aerosol Chemotherapy (ePIPAC). Ann Surg Oncol 23:S592–8.
  • Dakwar GR, Shariati M, Willaert W, et al. (2016). Nanomedicine-based intraperitoneal therapy for the treatment of peritoneal carcinomatosis – mission possible? Adv Drug Deliv Rev 108:13–24.
  • De Smet L, Colin P, Ceelen W, et al. (2012). Development of a nanocrystalline Paclitaxel formulation for HIPEC treatment. Pharm Res 29:2398–406.
  • Xu S, Fan H, Yin L, et al. (2016). Thermosensitive hydrogel system assembled by PTX-loaded copolymer nanoparticles for sustained intraperitoneal chemotherapy of peritoneal carcinomatosis. Eur J Pharm Biopharm 104:251–9.
  • Fujiware K, Armstrong D, Morgan M, et al. (2007). Principles and practice of intraperitoneal chemotherapy for ovarian cancer. Int J Gynecol Cancer 17:1–20.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.