Next Article in Journal
Biomaterial Drug Delivery Systems for Prominent Ocular Diseases
Next Article in Special Issue
Microfluidic Synthesis of Magnetite Nanoparticles for the Controlled Release of Antibiotics
Previous Article in Journal
Impact of Peptide Structure on Colonic Stability and Tissue Permeability
Previous Article in Special Issue
Recent Advances in the Development of Drug Delivery Applications of Magnetic Nanomaterials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Systematic Review

Nanomedicine and Hyperthermia for the Treatment of Gastrointestinal Cancer: A Systematic Review

1
Institute of Biopathology and Regenerative Medicine (IBIMER), Center of Biomedical Research (CIBM), University of Granada, 18100 Granada, Spain
2
Department of Anatomy and Embryology, Faculty of Medicine, University of Granada, 18071 Granada, Spain
3
Biosanitary Institute of Granada (ibs.GRANADA), SAS-University of Granada, 18014 Granada, Spain
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Pharmaceutics 2023, 15(7), 1958; https://doi.org/10.3390/pharmaceutics15071958
Submission received: 9 June 2023 / Revised: 8 July 2023 / Accepted: 14 July 2023 / Published: 15 July 2023

Abstract

:
The incidence of gastrointestinal cancers has increased in recent years. Current treatments present numerous challenges, including drug resistance, non-specificity, and severe side effects, needing the exploration of new therapeutic strategies. One promising avenue is the use of magnetic nanoparticles, which have gained considerable interest due to their ability to generate heat in tumor regions upon the application of an external alternating magnetic field, a process known as hyperthermia. This review conducted a systematic search of in vitro and in vivo studies published in the last decade that employ hyperthermia therapy mediated by magnetic nanoparticles for treating gastrointestinal cancers. After applying various inclusion and exclusion criteria (studies in the last 10 years where hyperthermia using alternative magnetic field is applied), a total of 40 articles were analyzed. The results revealed that iron oxide is the preferred material for magnetism generation in the nanoparticles, and colorectal cancer is the most studied gastrointestinal cancer. Interestingly, novel therapies employing nanoparticles loaded with chemotherapeutic drugs in combination with magnetic hyperthermia demonstrated an excellent antitumor effect. In conclusion, hyperthermia treatments mediated by magnetic nanoparticles appear to be an effective approach for the treatment of gastrointestinal cancers, offering advantages over traditional therapies.

Graphical Abstract

1. Introduction

The treatment of gastrointestinal cancer poses a significant challenge due to its increasing incidence in the population [1]. For example, colorectal cancer (CRC) ranks third in incidence and second in mortality, while stomach cancer (GC) and esophageal cancer (EC) rank fourth and sixth in mortality, respectively [2].
Current treatments encompass resection surgery along with chemotherapy, radiation therapy, and/or targeted therapies [3,4]. A broad range of drugs are utilized in various types of gastrointestinal cancer. These include 5-fluorouracil (5-FU), oxaliplatin (OXA), and irinotecan (IRI) in CRC; 5-fluorouracil combined with leucovorin, docetaxel, and oxaliplatin (FLOT) in GC [5]; and carboplatin and paclitaxel in EC, among others [6]. However, most drugs and their combinations generate side effects that can lead to treatment failure. For example, neurotoxicity is induced by OXA [7], cardiotoxicity by 5-fluorouracil [8], and severe neutropenia and hypersensitivity reactions follow paclitaxel treatment [9]. Additionally, drug resistance mechanisms induced by cytotoxins lead to poor response to chemotherapy and patient relapse [10]. Therefore, there is an essential need for new strategies to improve the prognosis of these diseases.
In this context, nanomedicine has emerged as a promising approach to cancer treatment and diagnosis. Numerous nanoformulations are available, with sizes ranging from 1 to 100 nm. Their inherent characteristics provide a drug delivery system with several advantages, such as reduced side effects of antitumor agents, improved targeting of the affected region, and increased drug levels in the tumor region, among others [10]. Both organic and inorganic nanoparticles (NPs) [11] have been employed to enhance cancer therapy. In fact, lipid-based nanoparticles have been the first clinically approved therapeutic nanoplatform against cancer by the FDA [12]. A classic example of their application is Doxil, a PEGylated liposome loaded with the drug Doxorubicin (DOXO) [13]. More recently, Onivyde, a liposome encapsulated with irinotecan, was approved by the FDA for the treatment of metastatic pancreatic cancer [14].
In this regard, magnetic NPs, a group of inorganic nanoformulations, have been proposed as an innovative strategy due to their physicochemical properties. They consist of a magnetic core and a polymeric coating, with iron oxide NPs being the most widely utilized. Magnetic NPs have superparamagnetic properties, which means that they are magnetized in the presence of an alternating external magnetic field (AMF), but lose magnetization without it, thereby reducing the potential for aggregation in the body and, consequently, the probability of embolization [15,16]. Among the advantages of these NPs is their ability to diffuse to the tumor region due to the application of a magnetic field near the target tissue [17]. Additionally, magnetic cores are used as contrast agents in various imaging techniques, such as magnetic resonance imaging (MRI), or newer techniques such as magnetic particle imaging. This allows for the tracking of these nanoformulations as they circulate within the organism, which is an interesting approach in terms of establishing targeted therapy [18].
Another major advantage derived from the use of magnetic NPs is their ability to generate high temperatures when an AMF is applied, which is known as hyperthermia. This property is considered one of the most intriguing and promising applications in the field of cancer nanomedicine, since it offers the possibility of applying a combined treatment, integrating the antitumor capacity of the drug loaded in the NPs and the hyperthermia generated [19]. Promising results have been obtained in various types of cancer after applying hyperthermia treatment alone or in combination with chemotherapy [20,21]. Recently, Narayanaswamy et al. (2022) utilized NPs with a MnFe2O4 core and an Fe3O4 shell against human colon and breast cancer cell lines (MDA-MB-231 and HT-29, respectively), increasing cell death by up to 70% [22]. Furthermore, Piehler et al. (2020) and Rego et al. (2020) demonstrated the applicability of hyperthermia in vivo using DOX-functionalized magnetic NPs and aminosilane-coated iron oxide NPs, respectively [23,24]. Dabaghi et al. (2021) developed 5-FU functionalized chitosan-coated magnetic NPs to deliver hyperthermia specifically against CRC-induced mice, showing a significant reduction in tumor volume and tumor vascularization [25]. In sum, many results support the benefits of hyperthermia therapy in cancer treatment.
In the present systematic review, we analyzed the most recently published studies on the application of hyperthermia based on magnetic NPs in gastrointestinal cancers. The review highlights crucial aspects of the emerging advancements in magnetic nanomaterials and provides a brief overview of the challenges and limitations of this therapeutic strategy.

2. Materials and Methods

2.1. Study Eligibility

The purpose of the present systematic review was to analyze the most recent and representative information on studies evaluating the therapeutic efficacy of NP-mediated hyperthermia in the treatment of various gastrointestinal cancers. This review was conducted following the criteria set out in the PRISMA guidelines [26]. To this end, only studies from the last 10 years were considered, deeming older ones obsolete. According to the Burton–Kebler index for obsolescence [27], more than half of the publications on this subject were included.

2.2. Inclusion Criteria

This systematic review included scientific publications between January 2013 and January 2023, with full text available and written in English. We also included works where hyperthermia treatment was applied through the use of an AMF on the NPs of interest as a therapy against any of the known gastrointestinal cancers.

2.3. Exclusion Criteria

Studies were excluded if hyperthermia, defined as an increase in temperature, was applied by any method other than the use of a magnetic field, such as water baths, lasers, ultrasound, etc. Furthermore, reviews, meta-analyses, systematic reviews, book chapters, or editorials were not considered for the review.

2.4. Data Sources

For the bibliographic search, the electronic databases Pubmed, SCOPUS, and Web of Science were used. The first established medical subject heading (MeSH) terms included: “Colorectal Neoplasms”, “Gastrointestinal Neoplasms”, “Esophageal Neoplasms”, “Intestinal Neoplasms”, “Stomach Neoplasms”, “Cecal Neoplasms”, “Duodenal Neoplasms”, “Ileal Neoplasms”, “Jejunal Neoplasms”, “Nanoparticles”, “Liposomes”, and “Hyperthermia”. The final search equation was ((“Colorectal Neoplasms” [MeSH Terms] OR “Gastrointestinal Neoplasms” [MeSH Terms] OR “Esophageal Neoplasms” [MeSH Terms] OR “Intestinal Neoplasms” [MeSH Terms] OR “Stomach Neoplasms” [MeSH Terms] OR “Cecal Neoplasms” [MeSH Terms] OR “Duodenal Neoplasms” [MeSH Terms] OR “Ileal Neoplasms” [MeSH Terms] OR “Jejunal Neoplasms” [MeSH Terms]) OR ((“colon” [Title/Abstract] OR “colorectal” [Title/Abstract] OR “colonic” [Title/Abstract] OR “Gastric*” [Title/Abstract] OR “Gastrointestinal” [Title/Abstract] OR “Esophageal*” [Title/Abstract] OR “Intestinal*” [Title/Abstract] OR “Stomach*” [Title/Abstract] OR “Cecal*” [Title/Abstract] OR “Duodenal*” [Title/Abstract] OR “Ileal*” [Title/Abstract] OR “Jejunal*” [Title/Abstract]) AND (“cancer*” [Title/Abstract] OR “tumor*” [Title/Abstract] OR “tumour*” [Title/Abstract] OR “neoplasm*” [Title/Abstract] OR “carcinoma*” [Title/Abstract]))) AND (“nanoparticles” [MeSH Terms] OR “nanoparticle*” [Title/Abstract] OR “nanoconjugate*” [Title/Abstract] OR “liposomes” [MeSH Terms] OR “liposome*” [Title/Abstract]) AND (“hyperthermia” [MeSH Terms] OR “hyperthermia*” [Title/Abstract]). Some minor modifications were made to adjust the search in the rest of the databases.

2.5. Study Selection

Two of the authors (L.G. and F.Q.) conducted the literature search. Initially, all articles were analyzed by title and abstract, with those meeting the inclusion criteria being selected. Both authors then reviewed all the selected articles through full-text analysis, considering the established inclusion and exclusion criteria.

2.6. Data Extraction

Following the study selection process, the same two authors separately analyzed the selected articles to extract data. The Cohen’s kappa statistical test result exceeded 0.8 (Cohen, 1968), indicating good agreement between the two authors [28]. All discrepancies were resolved by consensus between authors F.Q. and L.G. and, when necessary, two other authors intervened. A specific questionnaire, divided into two evaluation phases, was used to establish the quality of the selected articles; those papers scoring less than 6 points were excluded from the systematic review. Table 1, which presents the data obtained after exhaustive analysis of each article, includes information on the types of nanoformulations used, the antitumor agents transported, the applied magnetic fields, and notable in vitro and in vivo results, in addition to the article references.

3. Results and Discussion

3.1. Study Description

After conducting the bibliographic search in the PubMed, SCOPUS, and Web of Science databases, a total of 672 articles were obtained. Subsequently, 193 duplicate articles were excluded and, once analyzed by title and abstract, another 429 articles were excluded, leaving 50 selected. Likewise, nine of the fifty articles did not meet the inclusion criteria and one of them had low quality values. Therefore, a total of 40 articles were finally included in the present systematic review. All the data concerning the search are represented in the flow diagram in Figure 1.

3.2. Characteristics of Magnetic Nanoformulations

Table 1 shows the different nanoformulations used in each article and some of their main characteristics. Of the 40 articles analyzed, 100% of the nanoformulations were based on the magnetic properties of iron oxide nuclei or derivatives (magnetite or maghemite), indicating that iron was the preferred material for creating magnetic NPs. Furthermore, 36 out of 40 articles utilized an NP-based nanocarrier, while the remaining four articles employed more complex systems, including an exosome-based system [29], chitosan nanofibers [30], microrobots [31], and induced pluripotent stem cells [32]. Interestingly, three manuscripts featured antibody-functionalized NPs, including anti-CD133 [33], anti-HER2 [34], and radioactively labeled anti-CC49 [35]. Two articles used AP-1 [36] and TAT [37] peptides for functionalization. Regardless, the objective was to enhance the capabilities of the different nanoformulations (Figure 2). Approximately 47% of manuscripts combined hyperthermia therapy with drug usage. The most widely used chemotherapeutic agents were Doxorubicin (DOXO) and 5-Fluorouracil (5-FU), featured in eight and six articles, respectively. Other drugs included Oxaliplatin (OXA), Irinotecan (Iri), Cisplatin (CDDP), Bortezomib, and Niclosamide. Most of the selected articles analyzed the magnetic characteristics of the nanoformulations. Specifically, 25 articles highlighted the specific absorption rate (SAR) or magnetic saturation point (Ms), both of which are closely related to heat generation capacity after the application of an AMF. The value of the applied magnetic field and the duration of its application vary depending on the hyperthermia system. In fact, 30 of the 40 articles employed a field frequency within the range of 100 to 650 kHz. Conversely, two articles applied frequencies below 100 kHz, five used high frequencies such as Jahangiri et al. (2021) (13.56 MHz) [38], and three did not specify the frequency used. Ha et al. (2020) and Wang et al. (2021) demonstrated that functionalizations such as quantum dots [39] or the inclusion of NPs in gels [40] could impede temperature rise, suggesting that the choice of NPs is very relevant.
Moreover, and regarding the time of the treatment followed, the application of a 30 min treatment predominated, both in vitro and in vivo (13 of the 40 articles). The rest of the studies applied times ranging from 5 to 60 min. Likewise, in the case of in vivo tests, some authors applied a 30 min treatment continued over time, exposing the experimental animals to cycles of 30 min during the days established in each case [32,35,36]. Transfer to the clinic would involve the application of hyperthermia cycles alone or in combination with other treatments, depending on the approach. Due to the malignancy of the tumors and recurrences, it may be necessary to apply treatment cycles every certain time previously established. This is the case of the clinical trial conducted by Johannsen et al. (2005) for prostate cancer, in which patients were exposed to six weekly 60 min hyperthermia treatments [67]. Finally, concerning cell lines on which the NPs were tested, in vitro or in vivo, 38 articles were conducted on colon lines, notably the CT26 murine colorectal carcinoma line (11 articles) and HT29 human colorectal adenocarcinoma (9 articles). Only three of the forty articles used gastric [32,34] and esophageal [41] cancer lines. Therefore, the scarcity of investigations in some gastrointestinal cancers necessitates new research.

3.3. Biocompatibility of Hyperthermia Assays

Hyperthermia treatment safety is a major limitation in its clinical application. Interestingly, iron oxide was employed in the generation of NPs in all analyzed articles (40 articles) due to its biocompatibility [68], thereby avoiding damage to healthy cells. Fernández-Álvarez et al. (2021) used a non-tumor fibroblastic line and human blood samples to ensure no effect on normal tissue, erythrocytes, coagulation, and the complement system [42].
In addition, clinically accepted values of magnetic fields have been established, indicating that the product of the frequency and amplitude values must not exceed 5 × 109, as higher values can potentially harm DNA [69]. In fact, only 12 out of the 40 articles used magnetic fields within the clinically accepted range [32,35,40,43,44,45,46,47,48,49,50,51]. Conversely, 12 articles did not provide the necessary information to calculate this value [29,30,33,36,38,39,41,52,53,54,55,56]. Some authors have sought alternatives to generate magnetic NPs through a combustion system and varying concentrations of citric acid. These NPs induced high temperatures following an 87 kHz magnetic field, suggesting that a high Fe2+/Fe3+ ratio can enhance the hyperthermic capacity of nanoformulations without the need to increase the field frequency [57]. Ninety-five percent of the selected articles with in vivo tests demonstrated that hyperthermia treatment with NP did not induce damage to healthy tissues. In fact, Shen et al. (2019) generated magnetic solid lipid NPs coated with folic acid (FA) and Dextran and performed biocompatibility assays in CT26 colorectal tumor-bearing mice. Following hyperthermia treatment, they analyzed blood values and potential histological damage, obtaining normal ratios in all cases, thus supporting the apparent safety of these treatments in vivo [56]. Furthermore, Fang et al. (2021) demonstrated that magnetic liposomes functionalized with TAT/CSF1R inhibitor did not cause changes in body weight or histopathological damage following hyperthermia treatment in CT26 tumor-bearing mice [37]. However, new nanoformulation approaches are required that allow for an increase in temperature without increasing the frequency and intensity of the applied AMF. In this regard, it has been described that shortening the distance between NPs can enhance temperature rise [70]. Yang et al. (2020) generated magnetic nanoparticles and assembled and packaged them into a magnetic complex, obtaining higher temperature rises at very low frequencies (1.3–1.8 kHz) compared to single NPs [33]. These results have also been supported by other authors, such as Hu et al. (2023), who developed a controlled intracellular aggregation of NPs in acidic environments, obtaining better overall heating results [71].

3.4. In Vitro Assays

Of the 40 articles analyzed, 26 conducted cell viability tests applying hyperthermia treatment with or without chemotherapy (Table 1). All studies displayed better results with AMF than without it. However, significant differences were observed in induced cell death relative to the nanoformulations and the applied hyperthermia protocol.
Castellanos-Rubio et al. (2020) underscored the importance of selecting an optimal iron concentration for generating hyperthermia. They noted that at a concentration of 0.25 mg/mL, no significant cell death was observed in the colorectal cancer cell line HCT116. Conversely, at 0.5 mg/mL, cell survival decreased drastically [58]. Similarly, some NP functionalizations not only enhance heating capabilities but also cytotoxic effects in vitro. For example, Teo et al. (2017) generated SPIONs functionalized with 3-aminopropyltriethoxysilane (APTS) and/or protamine sulfate (PRO) loaded with TNF-α. They demonstrated that PRO increases NP toxicity in tumor cell lines, such as HepG2 and SW480, after AMF application compared to an APTS coating [59].
In certain cases, the molecular characteristics of the tumor cell line enable the selection of the appropriate NP functionalization, as demonstrated by Kagawa et al. (2021), who used anti-HER2 antibodies for treating the NUGC-4 cell line from gastric cancer, not showing any toxicity in healthy human fibroblasts due to their selectivity for internalization in cells with high HER2 expression, characteristic of gastrointestinal tumors such as gastric and esophageal tumors [34,72]. Among our selected articles, four reported complete cell death derived from the treatment [34,52,60,61]. Interestingly, three of these articles applied NP functionalized with carboxydextran. Both Fernandes et al. (2021) and Álvarez-Berríos et al. (2014) employed hyperthermia therapies in combination with chemotherapy, demonstrating the synergy that can result from applying both therapeutic approaches. Specifically, Fernandes et al. (2021) used polymer-coated iron oxide nanocubes loaded with DOXO, applying hyperthermia treatment from 10 to 90 min (3 cycles of 30 min) (182 kHz AMF) on patient-derived tumor stem cells (CSCs) [52]. After 24 h of exposure to treatment, more than 50% cell death was observed, reaching 100% at 7 days with a significant increase in the percentage of apoptosis and necrosis. These authors demonstrated that the heat generated by AMFs enhanced drug release and stimulated internalization in cells, thereby sensitizing them to chemotherapy. Similarly, five additional articles demonstrated significant drug sensitization, corroborating the enhanced release of some chemotherapeutics, such as 5-FU, OXA, and DOXO, following the use of AMF [36,43,46,47,62].
It has been proposed that the improvement following combined hyperthermia-chemotherapy treatment is due solely to the increase in temperature. However, interestingly, three of the selected articles demonstrated that less cell death was induced when water baths were applied (at the same temperature as those induced by AMF) [43,61]. Specifically, Álvarez-Berríos et al. (2013) used cisplatin-loaded iron oxide NPs and increased the temperature using a water bath or 237 kHz AMF. Hyperthermia generated 50% cell death in the Caco-2 CRC cell line compared to the 40% induced by the water bath. They hypothesized that AMF generates additional cellular stress that enhances membrane fluidity and ultimately results in cell death [49]. Therefore, hyperthermia generated by magnetic NPs appears to be a superior option for improving anticancer therapies compared to other systems.
Finally, hyperthermia associated with chemotherapy was not the only therapeutic approach against gastrointestinal cancer. Mirzaghavami et al. (2021) employed a combined treatment of chemotherapy, hyperthermia, and radiotherapy, inducing a greater decrease in the percentage of cell viability (45%) in the colorectal cancer cell line HT29 compared to individual treatments [50]. Additionally, a significant increase in apoptosis and necrosis was observed in the treated cell lines, increasing the Bax/Bcl2 ratio. Hyperthermia therapies induce more pronounced apoptosis than necrosis (Figure 2). Apoptosis was analyzed in nine of the forty selected articles. In fact, Jahangiri et al. (2021) provided an extensive description of this process, noting the overexpression of proapoptotic factor Bax, cleaved caspase-3, cleaved caspase-9, and PARP after treatment on HT29 and HCT116 colorectal cancer cell lines [38]. A similar increase in cleaved caspase-3 was shown in HCT116 by Ahmad et al. (2020) [55]. Moreover, Wydra et al. (2015) observed an increase in the generation of ROS after the application of AMF [63].

3.5. In Vivo Assays

A total of 50% (20) of the selected manuscripts carried out in vivo experiments, as shown in Table 1. The most commonly used animals were mice (90%), with CT26 tumor-bearing mice being the cancer model most frequently chosen by the authors. The utilization of magnetic NPs yielded beneficial results in all instances. In fact, Beyk and Tavakoli (2019) utilized nanohybrids of iron oxide and gold NPs, applying a magnet to the tumor region in CT26-tumor bearing mice for 3 h prior to hyperthermia treatment [53]. The results exhibited a higher temperature increase (49 °C) in the tumor area compared to treatment without a magnet (46 °C). This temperature increase led to significant inhibition of tumor size (92%). With regard to the generated temperature, most of the experiments achieved temperatures ranging from 41 to 50 °C [32,60]. However, after removal of the organs and application of MHT, it was observed that in mice with tumors induced from gastric lines, a large increase in temperature (up to 60 °C) was produced by applying ex vivo magnetic hyperthermia for 5 min, while in mice without tumors this high increase occurred in the liver, the organ in which the accumulation of these NPs took place [32]. Garanina et al. (2020) examined the impact of different temperatures on treatment. They noticed effective reduction in tumor growth in CT26 tumor-bearing mice at 42–43 °C. However, in 4T1 breast cancer tumor-bearing mice, which have a more resistant cell line, the tumor cells recurred after 20 days. In this case, effective treatment occurred at 46–48 °C. Furthermore, temperatures of 58–60 °C were tested, but these caused weight and motility losses, although recovery was observed over time [48]. These findings underscore the importance of personalizing treatments based on the tumors to be treated, whenever feasible, and avoiding excessively high temperatures that might lead to adverse effects.
Alternatively, magnetic steering was also reported by Wang et al. (2020) following magnet application (1 h) with positive outcomes [64]. Similarly, Beyk and Tavakoli observed MRI targeting due to these NPs’ ability to act as contrast agents, as was shown in seven other articles (Figure 3) [53]. These results hence signify the possibility of externally enhancing in vivo targeting to the tumor thanks to the magnetic capabilities of NPs. Additionally, Kwon et al. introduced the potential of improving targeting by functionalizing NPs. They applied an FA polymer to the shell of their nanoformulations, achieving improved tumor targeting in HT29 colorectal cancer tumor-bearing mice [29]. However, FA receptors are also found in the small intestine. That is why Shen et al. coated their nanoformulations with dextran, circumventing this compound’s recognition, thus directing more NPs to the colon region where the dextran was degraded by the dextranase produced there [56].
Although most in vivo studies were performed in mice or rats, Liu et al. (2013) conducted an in vivo model of esophageal cancer by injecting VX2 cells into the esophageal mucosa. The authors employed two hyperthermia models: one using a magnetic stem introduced into the esophagus of mice and the other by introducing magnetic NPs into the tumor mass. The results showed that, after the application of a 300 kHz field, both treatments showed a proven anti-tumor efficacy, although it was necessary to control both the temperature and the time of exposure to hyperthermia to avoid causing damage to healthy tissue [41].
All analyzed articles demonstrated positive outcomes in terms of tumor volume reduction when applying hyperthermia and NPs together compared to the application of the AMF or the nanoformulation alone. Nonetheless, it is worth noting that Arriortua et al. (2016) displayed highly varied results, as some tumors in the animal models used (CC-531 colon adenocarcinoma tumor-bearing rats) were almost obliterated while in other cases cell death was minor [65]. Moreover, eight of the twenty-one articles including in vivo analysis examined the apoptotic and necrotic effect induced by hyperthermia at the histological or genetic level. In fact, Beyk and Tavakoli (2019) and Kwon et al. (2021) revealed increases in the expression of some genes indicating apoptosis (cleaved PARP, Bax or cleaved Caspase-3) (Figure 3) [29,53]. These results were confirmed at the histological level in three more articles [35,43,65]. Conversely, Dabaghi et al. (2020) did not demonstrate any modulation in apoptosis gene expression, suggesting a cell death mediated by an increase in ROS [66]. Therefore, the mechanism through which tumor cell death is accomplished could be related to the type of NP and the hyperthermia treatment system. Additionally, three articles assessed the decrease in tumor metastases following treatment. Stankovic’ et al. (2020) did not exhibit dissemination of tumor cells after treatment in histological sections, while Matsumi et al. (2021) and Shen et al. (2019) observed a significant decrease in the number of metastatic nodules and ascites in murine models [35,56,60]. Finally, Fang et al. (2021) transplanted tumor cells from a mouse to other regions post hyperthermia treatment, but no tumor recurrence was noticed, implying activation of immune memory [37]. These results were validated by Jiang et al. (2022) using a CRC model surrounded by bacteria. In this case, immune system activation occurred after hyperthermia, resulting in an increase in cytokines, re-polarization of macrophages, and an increase in antigen presentation [54]. Therefore, hyperthermia treatments also have the ability to activate the immune response, which is typically suppressed in cancer (Figure 3). Likewise, the combination of hyperthermia and immunotherapy is another combined treatment option that may have very promising results. One of the selected articles used a magnetic liposomal system possessing the penetrating TAT peptide by which they administered the CSF1R inhibitor, so that it was possible to repolarize M2 macrophages, thus reducing immunosuppression in the tumor region [37]. It has previously been described that hyperthermia and immunotherapy have synergistic effects, giving rise to the possibility of triggering immunogenic cell death or reversing the immunosuppressive environment of tumors [73]. Thus, after tumor ablation by hyperthermia treatment, antigenic remnants would be released into the environment so they could be used as autologous vaccines against cancer. An example of this is presented in the work carried out by Pan et al. (2020), in which they applied magnetic NPs in combination with the programmed death ligand α-PD-L1 against a breast cancer model. Briefly, the generation of cytotoxic T cells against tumor antigens is achieved and α-PD-L1 prevents tumor immunosuppression, ultimately increasing the number of T cells and the immune response [74]. These results have been confirmed in other articles, demonstrating the potential of this therapeutic approach [75,76].
The benefits observed in in vitro and in vivo trials encourage the transfer of these treatments to the clinic. Currently, the application of magnetic hyperthermia as a possible treatment has been tested in clinical trials against prostate cancer (NCT02033447) and glioblastoma (DRKS00005476). In the first case, patients treated with magnetic nanoparticles received six cycles of therapy for 1 h during phase I of the study and showed tolerance and efficacy as antitumor therapy. Nevertheless, this therapy is still in phase II clinical trials. Regarding the hyperthermia treatment itself, its major limitations lie in the control of the local temperature reached in the tissue, which can negatively affect healthy cells, in addition to the heterogeneous distribution of the temperature in the tumor mass [77]. Additionally, one of the major problems in bringing this therapy to the clinic is the biosafety of the nanoformulations, which must have an exhaustive control of the size and components, making their production totally controlled [69]. Furthermore, another problem is the frequent parenteral administration of nanoparticles, as opposed to the simpler traditional oral administration. This fact generates a more expensive treatment, so that the commercial production of these treatments must be justified by greater efficacy or safety (including side effects) compared to the therapy traditionally used. Finally, many of the results obtained in preclinical studies (in vitro and in vivo models) are not subsequently retained in clinical trials, since certain characteristics, such as specific functionalization against a target, do not act in the same way in these models as in the human body [78].
For all the above reasons, the future of this line of research implies the need to expand the current research in order to solve the drawbacks encountered and finally allow the existence of a novel treatment that improves the quality of life of the affected patients.

4. Conclusions

Magnetic NP-driven hyperthermia treatment offers an innovative and promising therapeutic strategy for gastrointestinal cancers. Numerous magnetic NPs, capable of inducing heat and exhibiting varying biological properties, have been developed in recent years. These have been applied to some gastrointestinal cancers, although most assays have been conducted in vitro on CRC. As for the magnetic characteristics of NPs, iron oxide has predominantly been used as the magnetic core, with magnetic fields ranging between 100 and 600 kHz. Nearly half of the tests were conducted using combination therapies with drugs (chemotherapy), with DOXO being notably prominent. The outcomes have been very promising both in vitro and in vivo, reducing metastasis and tumor recurrence in certain cases. However, it has become evident that there is a need to broaden studies to encompass other cancers within the gastrointestinal tract. Further investigation will be necessary to affirm the benefits of hyperthermia application using magnetic NPs in the treatment of gastrointestinal cancer and to overcome barriers to clinical application.

Author Contributions

Conceptualization, J.P., R.O. and L.C.; methodology, L.G. and F.Q.; software, G.P.; validation, C.M. and G.P.; formal analysis, L.G. and F.Q.; investigation, L.G., F.Q. and G.P.; data curation, C.M.; writing—original draft preparation, C G., F.Q. and G.P.; writing—review and editing, C.M. and J.P.; visualization, C.M.; supervision, L.C. and R.O.; funding acquisition, C.M and J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the PI19/01478 and PMPTA22/00136 (Instituto de Salud Carlos III) (FEDER). In addition, it was partially supported by Project P20_00540, A-CTS-666-UGR20, B-CTS-122-UGR20 and PYC20 RE 035 (Proyectos I + D + i Junta de Andalucía 2020) (FEDER). LG acknowledges FP-PRE grant (2021) from the Junta de Andalucia (Spain).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Instrumentation Scientific Center (CIC) from University of Granada for technical assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Arnold, M.; Abnet, C.C.; Neale, R.E.; Vignat, J.; Giovannucci, E.L.; McGlynn, K.A.; Bray, F. Global Burden of 5 Major Types of Gastrointestinal Cancer. Gastroenterology 2020, 159, 335–349.e15. [Google Scholar] [CrossRef]
  2. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  3. Huang, F.L.; Yu, S.J. Esophageal Cancer: Risk Factors, Genetic Association, and Treatment. Asian J. Surg. 2018, 41, 210–215. [Google Scholar] [CrossRef] [PubMed]
  4. Machlowska, J.; Baj, J.; Sitarz, M.; Maciejewski, R.; Sitarz, R. Gastric Cancer: Epidemiology, Risk Factors, Classification, Genomic Characteristics and Treatment Strategies. Int. J. Mol. Sci. 2020, 21, 4012. [Google Scholar] [CrossRef]
  5. Sexton, R.E.; Al Hallak, M.N.; Diab, M.; Azmi, A.S. Gastric Cancer: A Comprehensive Review of Current and Future Treatment Strategies. Cancer Metastasis Rev. 2020, 39, 1179–1203. [Google Scholar] [CrossRef]
  6. Watanabe, M.; Otake, R.; Kozuki, R.; Toihata, T.; Takahashi, K.; Okamura, A.; Imamura, Y. Recent Progress in Multidisciplinary Treatment for Patients with Esophageal Cancer. Surg. Today 2020, 50, 12–20. [Google Scholar] [CrossRef] [Green Version]
  7. Knowlton, C.A.; Mackay, M.K.; Speer, T.W.; Vera, R.B.; Arthur, D.W.; Wazer, D.E.; Lanciano, R.; Brashears, J.H.; Knowlton, C.A.; Mackay, M.K.; et al. Cancer Colon. In Encyclopedia of Radiation Oncology; Springer: Berlin/Heidelberg, Germany, 2013; p. 77. [Google Scholar]
  8. Shiga, T.; Hiraide, M. Cardiotoxicities of 5-Fluorouracil and Other Fluoropyrimidines. Curr. Treat. Options Oncol. 2020, 21, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Al-Mahayri, Z.N.; AlAhmad, M.M.; Ali, B.R. Current Opinion on the Pharmacogenomics of Paclitaxel-Induced Toxicity. Expert. Opin. Drug Metab. Toxicol. 2021, 17, 785–801. [Google Scholar] [CrossRef]
  10. Garbayo, E.; Pascual-Gil, S.; Rodríguez-Nogales, C.; Saludas, L.; Estella-Hermoso de Mendoza, A.; Blanco-Prieto, M.J. Nanomedicine and Drug Delivery Systems in Cancer and Regenerative Medicine. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2020, 12, e1637. [Google Scholar] [CrossRef]
  11. Aghebati-Maleki, A.; Dolati, S.; Ahmadi, M.; Baghbanzhadeh, A.; Asadi, M.; Fotouhi, A.; Yousefi, M.; Aghebati-Maleki, L. Nanoparticles and Cancer Therapy: Perspectives for Application of Nanoparticles in the Treatment of Cancers. J. Cell Physiol. 2020, 235, 1962–1972. [Google Scholar] [CrossRef]
  12. Chaudhuri, A.; Kumar, D.N.; Shaik, R.A.; Eid, B.G.; Abdel-Naim, A.B.; Md, S.; Ahmad, A.; Agrawal, A.K. Lipid-Based Nanoparticles as a Pivotal Delivery Approach in Triple Negative Breast Cancer (TNBC) Therapy. Int. J. Mol. Sci. 2022, 23, 10068. [Google Scholar] [CrossRef] [PubMed]
  13. Gonzalez-Valdivieso, J.; Girotti, A.; Schneider, J.; Arias, F.J. Advanced Nanomedicine and Cancer: Challenges and Opportunities in Clinical Translation. Int. J. Pharm. 2021, 599, 120438. [Google Scholar] [CrossRef] [PubMed]
  14. Milano, G.; Innocenti, F.; Minami, H. Liposomal Irinotecan (Onivyde): Exemplifying the Benefits of Nanotherapeutic Drugs. Cancer Sci. 2022, 113, 2224–2231. [Google Scholar] [CrossRef] [PubMed]
  15. Vurro, F.; Jabalera, Y.; Mannucci, S.; Glorani, G.; Sola-Leyva, A.; Gerosa, M.; Romeo, A.; Romanelli, M.G.; Malatesta, M.; Calderan, L.; et al. Improving the Cellular Uptake of Biomimetic Magnetic Nanoparticles. Nanomaterials 2021, 11, 766. [Google Scholar] [CrossRef]
  16. Wu, K.; Su, D.; Liu, J.; Saha, R.; Wang, J.-P. Magnetic Nanoparticles in Nanomedicine: A Review of Recent Advances. Nanotechnology 2019, 30, 502003. [Google Scholar] [CrossRef] [Green Version]
  17. Farzin, A.; Etesami, S.A.; Quint, J.; Memic, A.; Tamayol, A. Magnetic Nanoparticles in Cancer Therapy and Diagnosis. Adv. Heal. Healthc. Mater. 2020, 9, 1901058. [Google Scholar] [CrossRef] [PubMed]
  18. Li, Y.; Xin, J.; Sun, Y.; Han, T.; Zhang, H.; An, F. Magnetic Resonance Imaging-Guided and Targeted Theranostics of Colorectal Cancer. Cancer Biol. Med. 2020, 17, 307–327. [Google Scholar] [CrossRef]
  19. Jose, J.; Kumar, R.; Harilal, S.; Mathew, G.E.; Parambi, D.G.T.; Prabhu, A.; Uddin, M.S.; Aleya, L.; Kim, H.; Mathew, B. Magnetic Nanoparticles for Hyperthermia in Cancer Treatment: An Emerging Tool. Environ. Sci. Pollut. Res. 2020, 27, 19214–19225. [Google Scholar] [CrossRef]
  20. Acar, M.; Solak, K.; Yildiz, S.; Unver, Y.; Mavi, A. Comparative Heating Efficiency and Cytotoxicity of Magnetic Silica Nanoparticles for Magnetic Hyperthermia Treatment on Human Breast Cancer Cells. 3 Biotech 2022, 12, 313. [Google Scholar] [CrossRef]
  21. Minaei, S.E.; Khoei, S.; Khoee, S.; Mahdavi, S.R. Sensitization of Glioblastoma Cancer Cells to Radiotherapy and Magnetic Hyperthermia by Targeted Temozolomide-Loaded Magnetite Tri-Block Copolymer Nanoparticles as a Nanotheranostic Agent. Life Sci. 2022, 306, 120729. [Google Scholar] [CrossRef]
  22. Narayanaswamy, V.; Jagal, J.; Khurshid, H.; Al-Omari, I.A.; Haider, M.; Kamzin, A.S.; Obaidat, I.M.; Issa, B. Hyperthermia of Magnetically Soft-Soft Core-Shell Ferrite Nanoparticles. Int. J. Mol. Sci. 2022, 23, 14825. [Google Scholar] [CrossRef]
  23. Piehler, S.; Dähring, H.; Grandke, J.; Göring, J.; Couleaud, P.; Aires, A.; Cortajarena, A.L.; Courty, J.; Latorre, A.; Somoza, Á.; et al. Iron Oxide Nanoparticles as Carriers for DOX and Magnetic Hyperthermia after Intratumoral Application into Breast Cancer in Mice: Impact and Future Perspectives. Nanomaterials 2020, 10, 1016. [Google Scholar] [CrossRef]
  24. Rego, G.; Nucci, M.; Mamani, J.; Oliveira, F.; Marti, L.; Filgueiras, I.; Ferreira, J.; Real, C.; Faria, D.; Espinha, P.; et al. Therapeutic Efficiency of Multiple Applications of Magnetic Hyperthermia Technique in Glioblastoma Using Aminosilane Coated Iron Oxide Nanoparticles: In Vitro and In Vivo Study. Int. J. Mol. Sci. 2020, 21, 958. [Google Scholar] [CrossRef] [Green Version]
  25. Dabaghi, M.; Rasa, S.M.M.; Cirri, E.; Ori, A.; Neri, F.; Quaas, R.; Hilger, I. Iron Oxide Nanoparticles Carrying 5-Fluorouracil in Combination with Magnetic Hyperthermia Induce Thrombogenic Collagen Fibers, Cellular Stress, and Immune Responses in Heterotopic Human Colon Cancer in Mice. Pharmaceutics 2021, 13, 1625. [Google Scholar] [CrossRef] [PubMed]
  26. Muka, T.; Glisic, M.; Milic, J.; Verhoog, S.; Bohlius, J.; Bramer, W.; Chowdhury, R.; Franco, O.H. A 24-Step Guide on How to Design, Conduct, and Successfully Publish a Systematic Review and Meta-Analysis in Medical Research. Eur. J. Epidemiol. 2020, 35, 49–60. [Google Scholar] [CrossRef]
  27. Száva-Kováts, E. Unfounded Attribution of the “Half-Life” Index-Number of Literature Obsolescence to Burton and Kebler: A Literature Science Study. J. Am. Soc. Inf. Sci. Technol. 2002, 53, 1098–1105. [Google Scholar] [CrossRef]
  28. Wanden-Berghe, C.; Sanz-Valero, J. Systematic Reviews in Nutrition: Standardized Methodology. Br. J. Nutr. 2012, 107, S3–S7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Kwon, S.-H.; Faruque, H.A.; Kee, H.; Kim, E.; Park, S. Exosome-Based Hybrid Nanostructures for Enhanced Tumor Targeting and Hyperthermia Therapy. Colloids Surf. B Biointerfaces 2021, 205, 111915. [Google Scholar] [CrossRef] [PubMed]
  30. Lin, T.-C.; Lin, F.-H.; Lin, J.-C. In Vitro Characterization of Magnetic Electrospun IDA-Grafted Chitosan Nanofiber Composite for Hyperthermic Tumor Cell Treatment. J. Biomater. Sci. Polym. Ed. 2013, 24, 1152–1163. [Google Scholar] [CrossRef]
  31. Park, J.; Jin, C.; Lee, S.; Kim, J.; Choi, H. Magnetically Actuated Degradable Microrobots for Actively Controlled Drug Release and Hyperthermia Therapy. Adv. Healthc. Mater. 2019, 8, 1900213. [Google Scholar] [CrossRef]
  32. Li, C.; Ruan, J.; Yang, M.; Pan, F.; Gao, G.; Qu, S.; Shen, Y.L.; Dang, Y.J.; Wang, K.; Jin, W.L.; et al. Human Induced Pluripotent Stem Cells Labeled with Fluorescent Magnetic Nanoparticles for Targeted Imaging and Hyperthermia Therapy for Gastric Cancer. Cancer Biol. Med. 2015, 12, 163. [Google Scholar] [CrossRef]
  33. Yang, S.-J.; Tseng, S.-Y.; Wang, C.-H.; Young, T.-H.; Chen, K.-C.; Shieh, M.-J. Magnetic Nanomedicine for CD133-Expressing Cancer Therapy Using Locoregional Hyperthermia Combined with Chemotherapy. Nanomedicine 2020, 15, 2543–2561. [Google Scholar] [CrossRef]
  34. Kagawa, T.; Matsumi, Y.; Aono, H.; Ohara, T.; Tazawa, H.; Shigeyasu, K.; Yano, S.; Takeda, S.; Komatsu, Y.; Hoffman, R.M.; et al. Immuno-Hyperthermia Effected by Antibody-Conjugated Nanoparticles Selectively Targets and Eradicates Individual Cancer Cells. Cell Cycle 2021, 20, 1221–1230. [Google Scholar] [CrossRef]
  35. Stanković, A.; Mihailović, J.; Mirković, M.; Radović, M.; Milanović, Z.; Ognjanović, M.; Janković, D.; Antić, B.; Mijović, M.; Vranješ-Đurić, S.; et al. Aminosilanized Flower-Structured Superparamagnetic Iron Oxide Nanoparticles Coupled to 131I-Labeled CC49 Antibody for Combined Radionuclide and Hyperthermia Therapy of Cancer. Int. J. Pharm. 2020, 587, 119628. [Google Scholar] [CrossRef] [PubMed]
  36. Kuo, C.Y.; Liu, T.Y.; Chan, T.Y.; Tsai, S.C.; Hardiansyah, A.; Huang, L.Y.; Yang, M.C.; Lu, R.H.; Jiang, J.K.; Yang, C.Y.; et al. Magnetically Triggered Nanovehicles for Controlled Drug Release as a Colorectal Cancer Therapy. Colloids Surf. B Biointerfaces 2016, 140, 567–573. [Google Scholar] [CrossRef]
  37. Fang, Y.; He, Y.; Wu, C.; Zhang, M.; Gu, Z.; Zhang, J.; Liu, E.; Xu, Q.; Asrorov, A.M.; Huang, Y. Magnetism-Mediated Targeting Hyperthermia-Immunotherapy in “Cold” Tumor with CSF1R Inhibitor. Theranostics 2021, 11, 6860–6872. [Google Scholar] [CrossRef]
  38. Jahangiri, S.; Khoei, S.; Khoee, S.; Safa, M.; Shirvalilou, S.; Pirhajati Mahabadi, V. Potential Anti-Tumor Activity of 13.56 MHz Alternating Magnetic Hyperthermia and Chemotherapy on the Induction of Apoptosis in Human Colon Cancer Cell Lines HT29 and HCT116 by up-Regulation of Bax, Cleaved Caspase 3&9, and Cleaved PARP Proteins. Cancer Nanotechnol. 2021, 12, 34. [Google Scholar] [CrossRef]
  39. Ha, P.T.; Le, T.T.H.; Ung, T.D.T.; Do, H.D.; Doan, B.T.; Mai, T.T.T.; Pham, H.N.; Hoang, T.M.N.; Phan, K.S.; Bui, T.Q. Properties and Bioeffects of Magneto–near Infrared Nanoparticles on Cancer Diagnosis and Treatment. New J. Chem. 2020, 44, 17277–17288. [Google Scholar] [CrossRef]
  40. Wang, Y.-J.; Lin, P.-Y.; Hsieh, S.-L.; Kirankumar, R.; Lin, H.-Y.; Li, J.-H.; Chen, Y.-T.; Wu, H.-M.; Hsieh, S. Utilizing Edible Agar as a Carrier for Dual Functional Doxorubicin-Fe3O4 Nanotherapy Drugs. Materials 2021, 14, 1824. [Google Scholar] [CrossRef]
  41. Liu, J.; Li, N.; Li, L.; Li, D.; Liu, K.; Zhao, L.; Tang, J.; Li, L. Local Hyperthermia for Esophageal Cancer in a Rabbit Tumor Model: Magnetic Stent Hyperthermia versus Magnetic Fluid Hyperthermia. Oncol. Lett. 2013, 6, 1550–1558. [Google Scholar] [CrossRef] [Green Version]
  42. Fernández-Álvarez, F.; Caro, C.; García-García, G.; García-Martín, M.L.; Arias, J.L. Engineering of Stealth (Maghemite/PLGA)/Chitosan (Core/Shell)/Shell Nanocomposites with Potential Applications for Combined MRI and Hyperthermia against Cancer. J. Mater. Chem. B 2021, 9, 4963–4980. [Google Scholar] [CrossRef] [PubMed]
  43. Thirunavukkarasu, G.K.; Cherukula, K.; Lee, H.; Jeong, Y.Y.; Park, I.-K.; Lee, J.Y. Magnetic Field-Inducible Drug-Eluting Nanoparticles for Image-Guided Thermo-Chemotherapy. Biomaterials 2018, 180, 240–252. [Google Scholar] [CrossRef] [PubMed]
  44. Mannucci, S.; Ghin, L.; Conti, G.; Tambalo, S.; Lascialfari, A.; Orlando, T.; Benati, D.; Bernardi, P.; Betterle, N.; Bassi, R.; et al. Magnetic Nanoparticles from Magnetospirillum Gryphiswaldense Increase the Efficacy of Thermotherapy in a Model of Colon Carcinoma. PLoS ONE 2014, 9, e108959. [Google Scholar] [CrossRef] [PubMed]
  45. Muñoz de Escalona, M.; Sáez-Fernández, E.; Prados, J.C.; Melguizo, C.; Arias, J.L. Magnetic Solid Lipid Nanoparticles in Hyperthermia against Colon Cancer. Int. J. Pharm. 2016, 504, 11–19. [Google Scholar] [CrossRef]
  46. Clares, B.; Biedma-Ortiz, R.A.; Sáez-Fernández, E.; Prados, J.C.; Melguizo, C.; Cabeza, L.; Ortiz, R.; Arias, J.L. Nano-Engineering of 5-Fluorouracil-Loaded Magnetoliposomes for Combined Hyperthermia and Chemotherapy against Colon Cancer. Eur. J. Pharm. Biopharm. 2013, 85, 329–338. [Google Scholar] [CrossRef]
  47. Jabalera, Y.; Garcia-Pinel, B.; Ortiz, R.; Iglesias, G.; Cabeza, L.; Prados, J.; Jimenez-Lopez, C.; Melguizo, C. Oxaliplatin–Biomimetic Magnetic Nanoparticle Assemblies for Colon Cancer-Targeted Chemotherapy: An In Vitro Study. Pharmaceutics 2019, 11, 395. [Google Scholar] [CrossRef] [Green Version]
  48. Garanina, A.S.; Naumenko, V.A.; Nikitin, A.A.; Myrovali, E.; Petukhova, A.Y.; Klimyuk, S.V.; Nalench, Y.A.; Ilyasov, A.R.; Vodopyanov, S.S.; Erofeev, A.S.; et al. Temperature-Controlled Magnetic Nanoparticles Hyperthermia Inhibits Primary Tumor Growth and Metastases Dissemination. Nanomedicine 2020, 25, 102171. [Google Scholar] [CrossRef]
  49. Torres-Lugo, M.; Castillo, A.; Mendez, J.; Rinaldi, C.; Soto, O.; Alvarez-Berrios, M.P. Hyperthermic Potentiation of Cisplatin by Magnetic Nanoparticle Heaters Is Correlated with an Increase in Cell Membrane Fluidity. Int. J. Nanomed. 2013, 8, 1003–1013. [Google Scholar] [CrossRef] [Green Version]
  50. Mirzaghavami, P.S.; Khoei, S.; Khoee, S.; Shirvalilou, S.; Mahdavi, S.R.; Pirhajati Mahabadi, V. Radio-Sensitivity Enhancement in HT29 Cells through Magnetic Hyperthermia in Combination with Targeted Nano-Carrier of 5-Flourouracil. Mater. Sci. Eng. C 2021, 124, 112043. [Google Scholar] [CrossRef]
  51. Pawlik, P.; Blasiak, B.; Depciuch, J.; Pruba, M.; Kitala, D.; Vorobyova, S.; Stec, M.; Bushinsky, M.; Konakov, A.; Baran, J.; et al. Application of Iron-Based Magnetic Nanoparticles Stabilized with Triethanolammonium Oleate for Theranostics. J. Mater. Sci. 2022, 57, 4716–4737. [Google Scholar] [CrossRef]
  52. Fernandes, S.; Fernandez, T.; Metze, S.; Balakrishnan, P.B.; Mai, B.T.; Conteh, J.; De Mei, C.; Turdo, A.; Di Franco, S.; Stassi, G.; et al. Magnetic Nanoparticle-Based Hyperthermia Mediates Drug Delivery and Impairs the Tumorigenic Capacity of Quiescent Colorectal Cancer Stem Cells. ACS Appl. Mater. Interfaces 2021, 13, 15959–15972. [Google Scholar] [CrossRef]
  53. Beyk, J.; Tavakoli, H. Selective Radiofrequency Ablation of Tumor by Magnetically Targeting of Multifunctional Iron Oxide–Gold Nanohybrid. J. Cancer Res. Clin. Oncol. 2019, 145, 2199–2209. [Google Scholar] [CrossRef] [PubMed]
  54. Jiang, H.; Guo, Y.; Yu, Z.; Hu, P.; Shi, J. Nanocatalytic Bacteria Disintegration Reverses Immunosuppression of Colorectal Cancer. Natl. Sci. Rev. 2022, 9, nwac169. [Google Scholar] [CrossRef]
  55. Ahmad, A.; Gupta, A.; Ansari, M.M.; Vyawahare, A.; Jayamurugan, G.; Khan, R. Hyperbranched Polymer-Functionalized Magnetic Nanoparticle-Mediated Hyperthermia and Niclosamide Bimodal Therapy of Colorectal Cancer Cells. ACS Biomater. Sci. Eng. 2020, 6, 1102–1111. [Google Scholar] [CrossRef]
  56. Shen, M.Y.; Liu, T.I.; Yu, T.W.; Kv, R.; Chiang, W.H.; Tsai, Y.C.; Chen, H.H.; Lin, S.C.; Chiu, H.C. Hierarchically Targetable Polysaccharide-Coated Solid Lipid Nanoparticles as an Oral Chemo/Thermotherapy Delivery System for Local Treatment of Colon Cancer. Biomaterials 2019, 197, 86–100. [Google Scholar] [CrossRef] [PubMed]
  57. Ramirez, D.; Oliva, J.; Cordova-Fraga, T.; Basurto-Islas, G.; Benal-Alvarado, J.J.; Mtz-Enriquez, A.I.; Quintana, M.; Gomez-Solis, C. High Heating Efficiency of Magnetite Nanoparticles Synthesized with Citric Acid: Application for Hyperthermia Treatment. J. Electron. Mater. 2022, 51, 4425–4436. [Google Scholar] [CrossRef]
  58. Castellanos-Rubio, I.; Rodrigo, I.; Olazagoitia-Garmendia, A.; Arriortua, O.; Gil de Muro, I.; Garitaonandia, J.S.; Bilbao, J.R.; Fdez-Gubieda, M.L.; Plazaola, F.; Orue, I.; et al. Highly Reproducible Hyperthermia Response in Water, Agar, and Cellular Environment by Discretely PEGylated Magnetite Nanoparticles. ACS Appl. Mater. Interfaces 2020, 12, 27917–27929. [Google Scholar] [CrossRef] [PubMed]
  59. Teo, P.; Wang, X.; Chen, B.; Zhang, H.; Yang, X.; Huang, Y.; Tang, J. Complex of TNF-α and Modified Fe3O4 Nanoparticles Suppresses Tumor Growth by Magnetic Induction Hyperthermia. Cancer Biother. Radiopharm. 2017, 32, 379–386. [Google Scholar] [CrossRef]
  60. Matsumi, Y.; Kagawa, T.; Yano, S.; Tazawa, H.; Shigeyasu, K.; Takeda, S.; Ohara, T.; Aono, H.; Hoffman, R.M.; Fujiwara, T.; et al. Hyperthermia Generated by Magnetic Nanoparticles for Effective Treatment of Disseminated Peritoneal Cancer in an Orthotopic Nude-Mouse Model. Cell Cycle 2021, 20, 1122–1133. [Google Scholar] [CrossRef]
  61. Alvarez-Berríos, M.P.; Castillo, A.; Rinaldi, C.; Torres-Lugo, M. Magnetic Fluid Hyperthermia Enhances Cytotoxicity of Bortezomib in Sensitive and Resistant Cancer Cell Lines. Int. J. Nanomed. 2014, 9, 145–153. [Google Scholar] [CrossRef] [Green Version]
  62. Hardiansyah, A.; Huang, L.Y.; Yang, M.C.; Liu, T.Y.; Tsai, S.C.; Yang, C.Y.; Kuo, C.Y.; Chan, T.Y.; Zou, H.M.; Lian, W.N.; et al. Magnetic Liposomes for Colorectal Cancer Cells Therapy by High-Frequency Magnetic Field Treatment. Nanoscale Res. Lett. 2014, 9, 497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Wydra, R.J.; Rychahou, P.G.; Evers, B.M.; Anderson, K.W.; Dziubla, T.D.; Hilt, J.Z. The Role of ROS Generation from Magnetic Nanoparticles in an Alternating Magnetic Field on Cytotoxicity. Acta Biomater. 2015, 25, 284–290. [Google Scholar] [CrossRef] [Green Version]
  64. Wang, J.T.-W.; Martino, U.; Khan, R.; Bazzar, M.; Southern, P.; Tuncel, D.; Al-Jamal, K.T. Engineering Red-Emitting Multi-Functional Nanocapsules for Magnetic Tumour Targeting and Imaging. Biomater. Sci. 2020, 8, 2590–2599. [Google Scholar] [CrossRef]
  65. Arriortua, O.K.; Garaio, E.; Herrero de la Parte, B.; Insausti, M.; Lezama, L.; Plazaola, F.; García, J.A.; Aizpurua, J.M.; Sagartzazu, M.; Irazola, M.; et al. Antitumor Magnetic Hyperthermia Induced by RGD-Functionalized Fe 3 O 4 Nanoparticles, in an Experimental Model of Colorectal Liver Metastases. Beilstein J. Nanotechnol. 2016, 7, 1532–1542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Dabaghi, M.; Quaas, R.; Hilger, I. The Treatment of Heterotopic Human Colon Xenograft Tumors in Mice with 5-Fluorouracil Attached to Magnetic Nanoparticles in Combination with Magnetic Hyperthermia Is More Efficient than Either Therapy Alone. Cancers 2020, 12, 2562. [Google Scholar] [CrossRef] [PubMed]
  67. Johannsen, M.; Gneveckow, U.; Eckelt, L.; Feussner, A.; Waldöfner, N.; Scholz, R.; Deger, S.; Wust, P.; Loening, S.A.; Jordan, A. Clinical Hyperthermia of Prostate Cancer Using Magnetic Nanoparticles: Presentation of a New Interstitial Technique. Int. J. Hyperth. 2005, 21, 637–647. [Google Scholar] [CrossRef] [Green Version]
  68. Sharifi, I.; Shokrollahi, H.; Amiri, S. Ferrite-Based Magnetic Nanofluids Used in Hyperthermia Applications. J. Magn. Magn. Mater. 2012, 324, 903–915. [Google Scholar] [CrossRef]
  69. Liu, X.; Zhang, Y.; Wang, Y.; Zhu, W.; Li, G.; Ma, X.; Zhang, Y.; Chen, S.; Tiwari, S.; Shi, K.; et al. Comprehensive Understanding of Magnetic Hyperthermia for Improving Antitumor Therapeutic Efficacy. Theranostics 2020, 10, 3793–3815. [Google Scholar] [CrossRef]
  70. Giustini, A.J.; Ivkov, R.; Hoopes, P.J. Magnetic Nanoparticle Biodistribution Following Intratumoral Administration. Nanotechnology 2011, 22, 345101. [Google Scholar] [CrossRef]
  71. Hu, A.; Pu, Y.; Xu, N.; Cai, Z.; Sun, R.; Fu, S.; Jin, R.; Guo, Y.; Ai, H.; Nie, Y.; et al. Controlled Intracellular Aggregation of Magnetic Particles Improves Permeation and Retention for Magnetic Hyperthermia Promotion and Immune Activation. Theranostics 2023, 13, 1454–1469. [Google Scholar] [CrossRef]
  72. Abrahao-Machado, L.F.; Scapulatempo-Neto, C. HER2 Testing in Gastric Cancer: An Update. World J. Gastroenterol. 2016, 22, 4619. [Google Scholar] [CrossRef]
  73. Stephen, Z.R.; Zhang, M. Recent Progress in the Synergistic Combination of Nanoparticle-Mediated Hyperthermia and Immunotherapy for Treatment of Cancer. Adv. Healthc. Mater. 2021, 10, 2001415. [Google Scholar] [CrossRef] [PubMed]
  74. Pan, J.; Hu, P.; Guo, Y.; Hao, J.; Ni, D.; Xu, Y.; Bao, Q.; Yao, H.; Wei, C.; Wu, Q.; et al. Combined Magnetic Hyperthermia and Immune Therapy for Primary and Metastatic Tumor Treatments. ACS Nano 2020, 14, 1033–1044. [Google Scholar] [CrossRef] [PubMed]
  75. Chao, Y.; Chen, G.; Liang, C.; Xu, J.; Dong, Z.; Han, X.; Wang, C.; Liu, Z. Iron Nanoparticles for Low-Power Local Magnetic Hyperthermia in Combination with Immune Checkpoint Blockade for Systemic Antitumor Therapy. Nano Lett. 2019, 19, 4287–4296. [Google Scholar] [CrossRef]
  76. Liu, X.; Zheng, J.; Sun, W.; Zhao, X.; Li, Y.; Gong, N.; Wang, Y.; Ma, X.; Zhang, T.; Zhao, L.-Y.; et al. Ferrimagnetic Vortex Nanoring-Mediated Mild Magnetic Hyperthermia Imparts Potent Immunological Effect for Treating Cancer Metastasis. ACS Nano 2019, 13, 8811–8825. [Google Scholar] [CrossRef] [PubMed]
  77. Hedayatnasab, Z.; Abnisa, F.; Daud, W.M.A.W. Review on Magnetic Nanoparticles for Magnetic Nanofluid Hyperthermia Application. Mater. Des. 2017, 123, 174–196. [Google Scholar] [CrossRef]
  78. Metselaar, J.M.; Lammers, T. Challenges in Nanomedicine Clinical Translation. Drug Deliv. Transl. Res. 2020, 10, 721–725. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Flow diagram that represents the articles included in the systematic review.
Figure 1. Flow diagram that represents the articles included in the systematic review.
Pharmaceutics 15 01958 g001
Figure 2. Use of magnetic nanoparticles in in vitro AMF hyperthermia application experiments. Magnetic nanoparticles, typically composed of iron, can be functionalized with antibodies (against HER2, TAG72, or CD133) or specific peptides to actively target a tumor population. AMF facilitates, in formulations that encapsulate drugs, a greater drug release into the cell and the induction of heightened cellular stress. These factors ultimately result in the death of tumor cells, triggering the activation of PARP and the cleavage of several caspases to activate the apoptotic pathway.
Figure 2. Use of magnetic nanoparticles in in vitro AMF hyperthermia application experiments. Magnetic nanoparticles, typically composed of iron, can be functionalized with antibodies (against HER2, TAG72, or CD133) or specific peptides to actively target a tumor population. AMF facilitates, in formulations that encapsulate drugs, a greater drug release into the cell and the induction of heightened cellular stress. These factors ultimately result in the death of tumor cells, triggering the activation of PARP and the cleavage of several caspases to activate the apoptotic pathway.
Pharmaceutics 15 01958 g002
Figure 3. Use of magnetic hypothermia in in vivo experiments. Typically, the assays involve an intravenous administration of the NPs into the mouse such that once they reach the tumor, they are capable of: (1) generating hyperthermia when exposed to an AMF, serving as an antitumor therapy alone or in combination with chemotherapeutic drugs; or (2) acting as contrast agents, creating negative contrast and potentially being used for tumor diagnosis and monitoring. Following the generation of hyperthermia, tumor cell death can occur through several pathways, with apoptosis, necrosis, or extensive oxidative stress being the most notable.
Figure 3. Use of magnetic hypothermia in in vivo experiments. Typically, the assays involve an intravenous administration of the NPs into the mouse such that once they reach the tumor, they are capable of: (1) generating hyperthermia when exposed to an AMF, serving as an antitumor therapy alone or in combination with chemotherapeutic drugs; or (2) acting as contrast agents, creating negative contrast and potentially being used for tumor diagnosis and monitoring. Following the generation of hyperthermia, tumor cell death can occur through several pathways, with apoptosis, necrosis, or extensive oxidative stress being the most notable.
Pharmaceutics 15 01958 g003
Table 1. Summary of the most relevant characteristics of the selected articles.
Table 1. Summary of the most relevant characteristics of the selected articles.
NanoformulationAntitumor AgentAMFIn Vitro AssayIn Vivo AssayTumor TypeMain ResultsReference
MnFe2O4-Fe3O4 core–shell NPs-384.5 kHz, 27.85 kA/mCytotoxicity assay (HT29)-CRCHigh cytotoxicity effect [22]
Cs MNPs5-FU435 kHz, 15.4 kA/m-HT29 tumor-bearing miceCRCDecrease in tumor size[25]
Exosome-FA-MNPsDOXO310 kHz Cytotoxicity assay (HT29)HT29 tumor-bearing miceCRCHigh cytotoxicity effect and decrease in tumor size[29]
MNPs loaded Cs nanofibers-750–1150 kHzCytotoxicity assay (CT26)-CRCHigh cytotoxicity effect [30]
SPIONs loaded microrobots5-FU430 kHz, 45 kA/mCytotoxicity assay (HCT116)-CRCHigh cytotoxicity effect [31]
Fluorescent MNP labeled iPS -63 kHz, 7 kA/m-MGC803 tumor-bearing miceGCDecrease in tumor size and good MRI results[32]
SPIO-APTES anti-CD133 MNPsIRI1.3–1.8 kHzCytotoxicity (Caco-2, HCT116, DLD1)HCT116 tumor-bearing miceCRCHigh cytotoxicity assay, decrease in tumor size and good MRI results[33]
anti-HER2 carboxydextran and amphiphilic polimer SPIONs-280 kHz, 31 kA/mCytotoxicity assay (NUGC-4)-GCHigh cytotoxicity effect [34]
Anti-131I-labeled CC49 SPIONs-252 kHz, 15.9 kA/m-LS174T tumor-bearing miceCRCDecrease in tumor size[35]
MPVA-AP1 nanovehiclesDOXO50–100 kHzLiberation assay -CRCHigh drug liberation and drug release[36]
TAT/CSF1R inhibitor functionalized magnetic liposomes-288 kHz, 35 kA/m-CT26 tumor-bearing miceCRCDecrease in tumor size and increased magnetic targeting[37]
PEG-PBA-PEG coated SPIONs5-FU13,560 kHzCytotoxicity assay (HT29, HCT116)-CRCHigh cytotoxicity effect [38]
Alginate coated MPNPs and QDsDOXO4–6.3 kA/m-CT26 tumor-bearing miceCRCGood MRI results[39]
Agar encapsulated MNPs DOXO400 kHz, 0.45 kA/mCytotoxicity assay (HT29)-CRCHigh cytotoxicity effect [40]
APTES coated MNPs-300 kHz-VX2 tumor-bearing rabbitsECDecrease in tumor size[41]
(maghemite/PLGA)/Cs NPs-250 kHz, 4 kA/mCytotoxicity assay (T84)Healthy miceCRCHigh cytotoxicity effect and good MRI results[42]
PLGA SPIONsDOXO205 kHz, 2 kA/mCytotoxicity assay (CT26)CT26 tumor-bearing miceCRCHigh cytotoxicity assay, drug release, decrease in tumor size and good MRI results[43]
Bacteria derived MNPs-187 kHz, 23 kA/m-HT29 tumor-bearing miceCRCIn vivo apoptotic and necrotic areas and good MRI results[44]
Solid-lipid MNPs-250 kHz, 4 kA/mCytotoxicity assay (HT29)-CRCHigh cytotoxicity effect [45]
Bacteria-derived MNPs5-FU250 kHz, 4 kA/mLiberation assay -CRCHigh drug release[46]
Bacteria-derived MNPsOXA197 kHz, 18 kA/mLiberation assay -CRCHigh drug release [47]
Cobalt ferrite NPs-261 kHz, 8–19.8 kA/mCytotoxicity assay (CT26)CT26 tumor-bearing miceCRCHigh cytotoxicity effect and decrease in tumor size[48]
MNPsCDDP237 kHz, 20 kA/mCytotoxicity assay (Caco-2)-CRCHigh cytotoxicity effect [49]
PEG-PCL-PEG/FA MNPs5-FU13,560 kHz, 0.4 kA/mCytotoxicity assay (HT29)-CRCHigh cytotoxicity effect [50]
MNPs-100 kHz, 4 kA/mMRI assay-CRCGood MRI results[51]
Iron oxide nanocubesDOXO182 kHzPatient-derived CSCs Patient-derived CSCs tumor-bearing miceCRCHigh cytotoxicity assay, decrease in tumor size[52]
Iron oxide NPs/Au NPs core/shell nanohybrid-13,560 kHzCytotoxicity assay (CT26)CT26 tumor-bearing miceCRCHigh cytotoxicity effect, decrease in tumor size, increased magnetic targeting and good MRI results[53]
ZnCoFe2O4 and ZnMnFe2O4 NPs-1.35 kA/mCytotoxicity assay (CT26)CT26 tumor-bearing miceCRCHigh cytotoxicity effect, decrease in tumor size and better targeting[54]
Polymers functionalized MNPsNiclosamide405 kHzCytotoxicity assay (HCT116)-CRCHigh Cytotoxicity effect [55]
Magnetic solid lipid NPs coated with FA and DextranDOXONot specifiedCytotoxicity assay (CT26)CT26 tumor-bearing miceCRCHigh cytotoxicity effect, decrease in tumor size and metastases[56]
Acid citric and EDC/NHC functionalized MNPs-87 kHz-340 kHz, 79.57 kA/mCytotoxicity assay (not specified)-CRCHigh cytotoxicity effect [57]
PMAO-PEG MNPs-650 kHz, 16.71 kA/m Cytotoxicity assay (HCT116)-CRCHigh cytotoxicity effect [58]
APTS/PRO functionalized SPIONs loaded with TNF-alfa-110 kHz, 8.75 kA/mCytotoxicity assay (SW480, HepG2)-CRCHigh cytotoxicity effect [59]
Carboxydextran coated MNPs-390 kHz, 28 kA/mCytotoxicity assay (HCT116)Peritoneal-dissemination miceCRCHigh cytotoxicity effect and metastases decrease[60]
Carboxydextran coated MNPsBortezomib233 kHz, 29.39 kA/mCytotoxicity assay (Caco-2)-CRCHigh cytotoxicity effect [61]
Liposome encapsulated citric acid-coated MNPsDOXO300 kHz, 59.3 kA/m Cytotoxicity assay (CT26)-CRCHigh cytotoxicity effect and drug release[62]
Monosaccharides coated MNPs-292 kHz, 51.0 kA/mCytotoxicity assay (CT26)-CRCHigh cytotoxicity effect [63]
PLGA SPIONs 930 kHz, 13 kA/m-CT26 tumor-bearing miceCRCIncreased magnetic targeting[64]
PMAO MNPs-606 kHz, 14 kA/m-CC-531 tumor-bearing ratsCRCHeterogeneous cytotoxicity results[65]
Cs MNPs5-FU435 kHz, 15.4 kA/m-HT29 tumor-bearing miceCRCSensitizes cells for further therapies and DNA damage[66]
AP1 (atherosclerotic plaque-specific peptide-1); APTES (3-aminopropyltriethoxysilane); CRC (colorectal cancer); Cs (chitosan); CSCs (Cancer stem cells); CSF1R (Colony stimulating factor 1 receptor); DOXO (doxorubicin); EC (esophageal cancer); EDC (1-ethyl-3-(3-dimethylaminopropyl) carbodiimide); FA (folic acid); 5-FU (5-fluorouracil); GC (gastric cancer); iPS (Induced pluripotent stem cells); IRI (irinotecan); MNPs (magnetic nanoparticles); MPVA (magnetic poly(vinyl alcohol)-based nanovehicles); MRI (magnetic resonance imaging); NHC (N-hydroxysuccinimide); NPs (nanoparticles); OXA (Oxaliplatin); PEG (polyethylene glycol); PLGA (poly(lactic-co-glycolic acid)); PMAO (poly(maleic anhydride-alt-1-octadecene)); SPIONs (Superparamagnetic iron oxide nanoparticles); TAT (Transactivator of transcription peptide).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gago, L.; Quiñonero, F.; Perazzoli, G.; Melguizo, C.; Prados, J.; Ortiz, R.; Cabeza, L. Nanomedicine and Hyperthermia for the Treatment of Gastrointestinal Cancer: A Systematic Review. Pharmaceutics 2023, 15, 1958. https://doi.org/10.3390/pharmaceutics15071958

AMA Style

Gago L, Quiñonero F, Perazzoli G, Melguizo C, Prados J, Ortiz R, Cabeza L. Nanomedicine and Hyperthermia for the Treatment of Gastrointestinal Cancer: A Systematic Review. Pharmaceutics. 2023; 15(7):1958. https://doi.org/10.3390/pharmaceutics15071958

Chicago/Turabian Style

Gago, Lidia, Francisco Quiñonero, Gloria Perazzoli, Consolación Melguizo, Jose Prados, Raul Ortiz, and Laura Cabeza. 2023. "Nanomedicine and Hyperthermia for the Treatment of Gastrointestinal Cancer: A Systematic Review" Pharmaceutics 15, no. 7: 1958. https://doi.org/10.3390/pharmaceutics15071958

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop