Back to Journals » International Journal of Nanomedicine » Volume 17

Antibody-Functionalized Nanoformulations for Targeted Therapy of Colorectal Cancer: A Systematic Review

Authors Cepero A , Luque C , Cabeza L , Perazzoli G, Quiñonero F , Mesas C, Melguizo C , Prados J 

Received 31 March 2022

Accepted for publication 21 August 2022

Published 31 October 2022 Volume 2022:17 Pages 5065—5080

DOI https://doi.org/10.2147/IJN.S368814

Checked for plagiarism Yes

Review by Single anonymous peer review

Peer reviewer comments 3

Editor who approved publication: Prof. Dr. Anderson Oliveira Lobo



Ana Cepero,1– 3,* Cristina Luque,1– 3,* Laura Cabeza,1– 3 Gloria Perazzoli,1,2 Francisco Quiñonero,1– 3 Cristina Mesas,1– 3 Consolación Melguizo,1– 3 Jose Prados1– 3

1Institute of Biopathology and Regenerative Medicine (IBIMER), Center of Biomedical Research (CIBM), University of Granada, Granada, 18100, Spain; 2Department of Anatomy and Embryology, Faculty of Medicine, University of Granada, Granada, 18071, Spain; 3Biosanitary Institute of Granada (ibs.GRANADA), SAS-University of Granada, Granada, 18014, Spain

*These authors contributed equally to this work

Correspondence: Consolación Melguizo, Institute of Biopathology and Regenerative Medicine (IBIMER), Center of Biomedical Research (CIBM), University of Granada, Granada, 18100, Spain, Tel +34-958-249833, Email [email protected]

Abstract: The failure of chemotherapeutic treatment in colorectal cancer (CRC), the second most mortal cancer worldwide, is associated with several drug limitations, such as non-selective distribution, short half-life, and development of multiple resistances. One of the most promising strategies in CRC therapy is the development of delivery systems based on nanomaterials that can transport antitumor agents to the tumor site more efficiently, increasing accumulation within the tumor and thus the antitumor effect. In addition to taking advantage of the increased permeability and retention effect (EPR) of solid tumors, these nanoformulations can be conjugated with monoclonal antibodies that recognize molecular markers that are specifically over-expressed on CRC cells. Active targeting of nanoformulations reduces the adverse effects associated with the cytotoxic activity of drugs in healthy tissues, which will be of interest for improving the quality of life of cancer patients in the future. This review focuses on in vitro and in vivo studies of drug delivery nanoformulations functionalized with monoclonal antibodies for targeted therapy of CRC.

Keywords: nanoformulation, colon carcinoma, monoclonal antibody, 5-fluorouracil, targeted therapy

Introduction

Colorectal cancer (CRC) accounts for the third highest incidence of cancer and the second highest mortality in the world.1,2 Changes in lifestyle and dietary patterns, including increased consumption of processed food, sedentarism, smoking, alcohol, and low intake of fruits, vegetables, and calcium, among others, have been related to a significant increase in the incidence of CRC in recent years.3 Moreover, far from improving, CRC mortality is estimated to increase by more than 60.0% by 2035.4

The treatment of choice for non-metastatic CRC is surgery. However, the management of metastatic CRC, which occurs in 50% of patients,5 consists of surgical resection of the tumor (and metastases when possible), together with chemotherapy, radiotherapy and targeted therapy. 5-fluorouracil (5-FU), oxaliplatin (OXA) and irinotecan (IRI) are the chemotherapeutics of first choice, and can be used alone or in combination regimens such as FOLFOX (5-FU/leucovorin/OXA), FOLFIRI (5-FU/leucovorin/IRI) and FOLFOXIRI (5-FU/leucovorin/OXA/IRI).6 Unfortunately, these drugs have numerous side effects on proliferating cells, such as those found in the digestive tract, hair follicles or hematopoietic progenitors. In fact, FOLFOXIRI has been significantly associated with the development of grade 3 and 4 neurotoxicity and neutropenia, limiting its therapeutic success due to treatment discontinuation by patients.7 Likewise, the search for CRC cell-specific biomarkers has allowed the development of targeted therapy; including agents acting against EGFR (eg, cetuximab and panitumumab), as well as against VEGF (eg, bevacizumab and aflibercept).8,9 These biomarkers, in turn, can be used for the generation of new strategies for targeted drug delivery to tumor cells. However, all these therapeutic advances have failed to increase the survival of patients with advanced disease which remains below 15%.10

Treatment failure of metastatic CRC has various causes, including adverse effects of chemotherapy, drug non-specificity, and drug resistance mediated by ABC (ATP-binding cassette) transporters, among others.11 Thus, the development of new strategies to improve the treatment of these patients is a priority. In this context, nanomedicine represents a promising field for the development of new antitumor nanodrugs that could be released locally at the tumor site, overcoming the limitations of conventional chemotherapy and improving adherence to treatment and the quality of life of patients.12

The most widely used nanoformulations in cancer therapy include polymeric nanoparticles (NPs), lipid nanoformulations (liposomes and micelles) and inorganic NPs. These nanoformulations improve stability, solubility, and drug half-life, and are able to increase accumulation within the tumor due to the EPR effect of solid tumors, which is closely related to passive targeting and relies on paracellular transport of the nanoformulations through compromised blood vessels and subsequent non-specific interaction with tumor cells. However, their effectiveness is compromised due to high inter- and intra-tumor variability.12–14 Furthermore, some of these nanodrugs block resistance mechanisms that prevent the elimination and/or degradation of the drug by the tumor cell.15 Specifically, in CRC therapy, a wide variety of nanoformulations are being used, including liposomes and polymeric NPs,16 which have shown high efficacy. This is the case with liposomes associated with doxorubicin (DOX) and curcumin (co-encapsulation), which increased antitumor efficacy in CRC in vivo models,17 and with polymeric NPs loaded with Nutlin-3a and granulocyte colony stimulating factor- macrophages (GM-CSF), which have recently shown enhanced antiproliferative effects against CRC.18 Likewise, some nanoformulations against CRC attempt to avoid multidrug resistance (MDR) mechanisms. For instance, Jiang et al used nanocomposites based on a gold nanorod core-shell associated with DOX and functionalized with poly-histidine and d-α-tocopherol polyethylene glycol 1000 succinate (TPGS) that inhibited p-glycoprotein.19 Clinical trials are the final step in the use of these nanoformulations in CRC, as is the case with TKM-080301, a lipid NP loaded with an siRNA against the PLK1 protein, or CRLX101, a PEGylated cyclodextrin NP with camptothecin.16,20,21 These trials may prove the usefulness of these systems to improve the prognosis of CRC patients.

Specific interactions with tumor cells can be achieved by active targeting nanoformulations, designed with specific ligands that recognize with high affinity tumor cell receptors. This active targeting allows i) to improve the retention of passively accumulated nanoformulations due to ligand-receptor interaction, and ii) to provide specific interaction with tumor cells by reducing interactions with non-targeted cells.14 In this context, the functionalization of NPs by using monoclonal antibodies —the most widely used targeting-ligands22 that allow their targeting to the tumor cell represents a great opportunity for the improvement of oncological treatment. This tissue- or cell-specific delivery occurs through the development of antibody-NP conjugates (ANCs) that bind specifically to the cell type of interest significantly reducing their non-specificity and toxicity in non-tumor tissues.23 The use of ANCs in CRC has emerged as a field of interest. The EGFR tyrosine kinase receptor, involved in tumor growth and progression,24 is one of the surface molecules most commonly used to target ANCs in CRC25,26 and other types of cancer (eg, prostate, skin cancer and lung cancer).27–29 In fact, a patent has already been published for cetuximab (anti-EGFR) bound to carbon quantum dots, which showed a high targeting capacity in EGFR-overexpressing tumor cells.30,31 New therapies targeting cancer stem cells (CSC), which appear to be responsible for resistance to chemotherapy, radiotherapy and the development of metastases, are under investigation.32 In fact, novel nanovehicles that specifically target this cell population have been developed. For example, a recent study used functionalized DCLK1 folic acid conjugated hesperetin encapsulated in chitosan NPs to selectively target colorectal CSC.33 The stem cell biomarker CD133 has become a way of targeting colon,34 breast35 and ovarian stem cells,36 among others. Recently, this ligand has been used by Mohd-Zahid et al to synthesize PEGylated gold NPs that significantly increased intracellular drug (5-FU) accumulation in HCT116 CRC cells.37

The main objective of this systematic review is to analyze all the recent published studies on NPs, liposomes and micelles functionalized with monoclonal antibodies and associated to a molecule with antitumor activity against CRC. This review summarizes the main antibody-NP conjugates used in human CRC, including in vitro and in vivo assays, and supports the need for further studies to understand their main mechanism of action and their application in patients with CRC.

Materials and Methods

Study Eligibility

The purpose of this systematic review is to evaluate the most recent scientific publications containing information on the therapeutic efficacy of antitumor agents carried by NPs, liposomes or micelles functionalized with monoclonal antibodies or fragments of them in CRC. The systematic review was performed according to the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) criteria.38 Literature published more than 10 years ago was considered obsolete. In addition, according to the Burton-Kebler index for obsolescence, more than half of the scientific publications on this topic were included.39

Inclusion and Exclusion Criteria

This review included studies written in English and with accessible full text, published between May 2011 and May 2021, about NPs, liposomes or micelles that i) target cells of interest with monoclonal antibodies or fragments of them; and ii) function as antitumor molecule delivery agents against CRC. Studies with insufficient information on the data provided, as well as systematic reviews, reviews, meta-analyses and editorials were excluded.

Data Sources

The literature search was performed in the following electronic databases: Pubmed, SCOPUS and Web of Science. First, the medical subject heading (MeSH) terms were established; “Colorectal Neoplasms”, “Nanoparticles”, “Liposomes”, “Micelles” and “Antibodies” were used as descriptive terms. The final search strategy was ((“Colorectal Neoplasms” [MeSH Terms] OR ((“colon” [Title/Abstract] OR “colorectal” [Title/Abstract] OR “colonic” [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] OR “micelles” [MeSH Terms] OR “micelle*” [Title/Abstract]) AND (“antibodies” [MeSH Terms] OR “antibod*” [Title/Abstract]))). Slight modifications were made to adapt the search strategy to the different databases.

Study Selection

Two of the authors (C.L. and A.C.) conducted the literature search. In the first stage of the review, the titles and abstracts of the studies were screened and those meeting the inclusion criteria were selected for further review. In the second stage, the authors reviewed the full text of all the selected articles according to the previously established inclusion and exclusion criteria.

Data Extraction

After the selection process of publications was completed, the same two authors independently reviewed and extracted the data from the selected studies. According to Cohen’s Kappa statistical test,40 which exceeded 0.8, there was concordance between the two investigators.41 Discrepancies were resolved by consensus between C.L. and A.C., or between two other authors if necessary. The quality of the selected publications was determined by means of a specific questionnaire. The extracted data are summarized in Tables 1–4 and have been classified according to the type of nanoformulation used. In order to facilitate the understanding of the selected studies, details on the following variables have been included: antitumor agent transported, monoclonal antibody used for functionalization, type of nanoformulation, most relevant in vitro and in vivo results, and publication reference.

Table 1 Antiproliferative Activity of Lipid Nanoformulations in Colorectal Cancer Models

Table 2 Antiproliferative Activity of Polymeric Nanoformulations in Colorectal Cancer Models

Table 3 Antiproliferative Activity of Inorganic Nanoformulations in Colorectal Cancer Models

Table 4 Antiproliferative Activity of Hybrid and Peptide Nanoformulations in Colorectal Cancer Models

Results and Discussion

Study Description

After the initial literature search in the different electronic databases, 778 articles were retrieved following the inclusion and exclusion criteria. Titles and abstracts were reviewed to exclude articles that did not meet the selection criteria as well as duplicates. After full-text review of the 80 articles selected, 36 did not fit the subject of the review, 4 were not available in full text, and 1 was excluded due to low quality. Finally, 39 articles were included in this systematic review. The flow diagram of the search process is presented in Figure 1.

Figure 1 Flow diagram of articles included in this systematic review.

Most of the studies focused on the evaluation of the cytotoxic effects of antitumor agents delivered by lipid nanoformulations (14 articles) and inorganic nanoformulations (10 articles). The remaining studies used polymeric, peptide or hybrid nanoformulations (ie, derived from combinations of the previous ones). In addition, 28 articles analyzed carrier nanoformulations of conventional antitumor agents such as DOX (10 articles), 5-FU (8 articles) and paclitaxel (PTX) (4 articles). Moreover, most of the functionalizations used antibodies against the EGFR (15 articles), with cetuximab being the most widely used. The predominant targeting of EGFR, a transmembrane receptor of the tyrosine kinase family that mediates cell signalling cascades involved in cell proliferation, angiogenesis and apoptosis, is explained by its overactivation in numerous types of cancer, including CRC. Therefore, the use of monoclonal antibodies that act as selective competitors by blocking the binding of endogenous ligands and inhibiting the signalling cascade is a strategy under extensive study.24

Lipid Nanoformulations

Of the 39 articles included in this review, 14 analyzed the antitumor effect of lipid nanocarriers in CRC (Table 1), including liposomes coated with PEG chains (11 articles), the chemotherapeutic drug Doxil® (2 articles), and a cerasome (1 article). Of note, studies with Doxil® targeted against CD133 +34 and CD44 + tumor cells42 -both of which are clusters of differentiation associated with

CSC of CRC-32 showed a 3-fold and 2-fold improvement in cell cytotoxicity compared to non-targeted Doxil®, respectively.

Regarding cytotoxic drugs transported by lipid nanocarriers, DOX (5 articles) and 5-FU (3 articles) were the most commonly used. PEGylated liposomes loaded with 5-FU and functionalized with anti-FZD10, anti-EGFR and anti-ITGB6 were tested in CRC. Specifically, of the 14 articles on lipid nanoformulations, five were directed against EGFR. Interestingly, the use of anti-EGFR and anti-ITGB6 –which recognizes integrin β6, involved in invasion and metastasis of CRC- nanoformulations showed significant enhancement of in vitro cytotoxicity, including reduced tumor growth in murine CRC models compared to free 5-FU.47,52 Furthermore, anti-FZD10 nanoformulations targeting the FZD10 receptor of the WNT signaling cascade showed enhanced antitumor effect of the drug at low concentrations (1–2 µM).43 PTX,46 OXA,50 Celecoxib,55 Z-Leucinyl-Leucinyl-Norleucinal tripeptide,44 RA-V cyclopeptide45 and MG85 complex48 were also analyzed.

HT29 and HCT-116 were the most frequently used human colon adenocarcinoma cell lines in in vitro assays studying cytotoxicity (12 assays), apoptosis, migration, and tumor hypoxia.43,45 In fact, Corvo et al showed that encapsulation of the MG85 complex in PEGylated liposomes functionalized with Cetuximab led to 2- and 4-fold higher cytotoxicity in the HCT-116 cell line compared to free MG85.48 Likewise, the RA-V chemotherapeutic transported together with the HIF-1a inhibitor RX-0047 in a pH-sensitive liposome targeting cell death receptor 5 (DR-5) demonstrated greater activation of the caspase-8 cascade than both compounds alone. In addition, this liposome was able to decrease protein expression of the HIF-1 factor by reducing the tumor hypoxic environment in HT-29 and/or HCT-116 cells.45 On the other hand, HT29 tumor-bearing mice were preferred in in vivo studies. Liang et al showed that the use of immunoliposomes loaded with 5-FU in this murine model induced a significant (2-fold) regression of tumor volume and 1.5- to 1.7-fold higher cell apoptosis compared to 5-FU-loaded liposomes.52 Furthermore, Arabi et al demonstrated a strong antitumor effect with the use of anti-CD44-

conjugated Doxil®. Specifically, at doses of 10 and 15 mg/kg, the nanoformulation increased survival time (1.4- to 1.5-fold) and reduced tumor growth by more than 94% compared to control in C-26 tumor-bearing mice.42

In conclusion, most of the studies showed improved therapeutic effect after the use of targeted lipid nanoformulations, except for five of them — although one of these was not studied in in vivo or in vitro assays.53 Lipid nanoformulations appear to be the most widely used ANCs in CRC due to their favorable toxicological profile and high bioavailability.56,57 Moreover, the possibility of adding a PEG coating to the lipid decreases its recognition by the endothelial reticular system, increasing the amount of drug available in the tumor. In fact, the use of a PEG-coated lipid nanoformulation and antibody conjugation was found in eleven articles.

Polymeric Nanoformulations

Polymeric NPs are a promising option as drug delivery systems, as they have favorable biocompatibility characteristics, good solubility profiles, and long circulation times. However, there is a translational gap between animal models and patients,58 making them a strategy to be improved. Poly(lactic-co-glycolic acid) (PLGA), a copolymer composed of PLA and PGA, both approved by the Food and Drug Administration (FDA) and the European Medicines Agency (EMA),59 was the most frequently used polymeric NP to be functionalized with antibodies (3 out of 7 articles). Other polymers used for the synthesis of nanocarriers were poly(glycolic acid) (PGA),60 poly(lactic acid) (PLA),61 polycaprolactone (PCL)62 and citrus pectin-chitosan NPs (Table 2).63 On the other hand, 4 articles used co-polymerization with PEG to increase

circulation time, decrease immunogenicity and/or aid binding of the antibody.

All 7 articles analyzed different drugs with the exception of the studies by Fay et al and by Schmid et al, which focused on camptothecin encapsulated in PLGA-anti-DR5 and PLGA-anti TRAIL NPs, respectively.64,65 In fact, the former led to enhanced cytotoxicity (4-fold) and increased cell apoptosis of camptothecin compared to camptothecin alone.64 5-FU,61 PTX,59 IRI,62 carfilzomib,66 curcumin63 and triptolide60 were other drugs associated with polymeric NPs. Cetuximab was used to direct nanoformulations in 2 of the 7 articles.61,63 The most relevant result

was obtained by Sabra et al, who demonstrated that citrus pectin-chitosan NPs functionalized with curcumin-transporting Cetuximab achieved 29.8- and 30-fold higher cytotoxicity than free curcumin in CaCo-2 and HCT-116 cell lines, respectively. This targeted nanoformulation also induced 1.8-fold higher cell cycle arrest in G2/M phase compared to free curcumin in CaCo-2 cells.63

Finally, HCT-116 and HT29 CRC cell lines were the most commonly used in in vitro and in vivo assays. Recently, Yalikong et al assayed a triptolide delivery system based on PGA-L-Phe NPs targeted against HER2 that induced 3-fold higher cytotoxicity in HT29 cells and higher G1-S arrest compared to the drug alone.60 In addition, 4 of the 7 articles used murine CRC models to evaluate the therapeutic efficacy of nanoformulations such as mPEG-PCL/mal-PEG-PCL NPs targeted against the CD-133 biomarker and loaded with IRI, which induced greater tumor regression in HCT-116 tumor-bearing mice compared to non-targeted NPs and free IRI.62

Polymeric NPs present numerous advantages that promote their application as nanovehicles in targeted therapy of CRC, being biocompatibility one of the most characteristic properties, also present in lipid nanoformulations. However, control over the shape and size of polymeric NPs allows long circulation times by evasion of the endothelial reticular system, which is associated with increased passive accumulation in the tumor via the EPR effect. Moreover, control over design allows the development of precision chemistry for specific and orientated binding of monoclonal antibodies, which is essential for proper targeting.67

Inorganic Nanoformulations

As shown in Table 3, 10 articles studied the therapeutic effect of inorganic nanocarriers. Among metallic nanoformulations (8 articles), the most common were gold NPs (6 articles) followed by silica and ruthenium NPs (1 article each).68,69 Five of them selectively targeted EGFR+ CRC cells. Notably, mifepristone-loaded mesoporous silica NPs coated with anti-EPCAM antibody were designed to selectively detect circulating tumor cells (CTC) of CRC through the membrane glycoprotein EpCAM, which is one of the most widely used surface antigens to differentiate cancer cells of epithelial origin from healthy blood cells.70 This nanoformulation showed to decrease the adhesion of HT29 and SW620 colon cancer cells to endothelial cells by 1.3-fold. In fact, a decrease in lung metastases of SW620 tumor-bearing mice compared to the non-targeted NP and free mifepristone was demonstrated.68 Non-metallic conformations (2 articles) used EGFR antibody; one showed the efficacy of polyethylenimine (PEI)-coated quantum dots to increase cytotoxicity in HCT-116 cells,71 and the other analyzed the antitumor effect of nanodiamonds to inhibit tumor growth and to induce apoptosis.74

The drugs most frequently associated with inorganic nanoformulations were 5-FU and DOX (4 articles). Mohd-Zahid et al and Liszbinski et al functionalized gold NPs with anti-CD133 and anti-EGFR antibodies, respectively. The former enhanced the antitumor effect in the HCT-116 cell line compared to the non-targeted NP,37 The latter demonstrated an increase in apoptotic cells compared to free 5-FU and the non-targeted nanoformulation, but no decrease in cell viability.26 Concerning DOX, Lo et al synthesized graphene quantum dots coated with PEI polymer and functionalized with anti-EGFR that decreased DOX toxicity in a murine CRC model.71 In addition, Emami et al developed gold NPs loaded with DOX and targeted to the immune checkpoint protein PD-L1 that enhanced in vitro antitumor effect (1.6-fold) in comparison with the non-targeted nanoformulation, and increased the number of necrotizing cells and reactive oxygen species (ROS) generation.73 PTX,74 OXA,76 Zn (II) coordination compound,72 Mifepristone,68 RBT69 and TS26575 were other antitumor agents tested. The most relevant result was obtained with OXA-loaded gold NPs targeted against the DR5 receptor, a transmembrane protein belonging to the tumor necrosis factor receptor (TNFR) family, which is overexpressed in stages II and III of CRC.77 DR5-targeted gold NPs achieved 3.2-fold higher cytotoxicity and cell apoptosis than free OXA. In vivo studies also showed greater tumor growth inhibition than both the free drug and non-targeted NP in a murine model of CRC.76

Therefore, gold NPs stand out among inorganic NPs as ideal candidates for the specific transport of antitumor molecules.75 Moreover, gold NPs exhibit unique optical and Surface Plasmon Resonance (SPR) properties that are useful for tumor detection and for the development of image-based therapies, such as photothermal (PTT) and photodynamic therapy (PDT).78 For example, DOX-conjugated and anti-PD-L1 targeting gold nanoformulations plus NIR irradiation synergistically inhibited the in vitro proliferation of the CT-26 cell line via higher apoptosis and cell cycle arrest.73 However, unlike other types of nanovehicles, inorganic NPs present difficulties to be eliminated or excreted, resulting in adverse effects such as inflammation and tissue cysts. This limitation should be solved with the development of new biodegradable nanoformulations.79

Other Nanoformulations

Only 8 of the 39 selected articles analyzed hybrid (6 articles) or peptide (2 articles) nanoformulations (Table 4). Most hybrid nanoformulations were combinations of polymeric NPs and lipid nanoformulations (4 articles), usually in a polymeric micelle conformation (2 articles). In this context, Andrade et al developed CD44v6-targeted polymeric micelles loaded with niclosamide that demonstrated greater antitumor effect compared to the non-targeted nanoformulation and free drug (2.8-and 3-fold, respectively) in the HCT-116 cell line. These nanoformulations also showed activity against HT29 colonspheres and circulating tumor cells (CTC) in HT29 tumor-bearing mice.80 Hybrid nanocarriers of inorganic metal nanocomposites (gold and silica NPs) combined with PLGA NPs and liposomes, respectively, have been also developed.81,82 In fact, Chen et al used EGFR-targeted mesoporous silica NPs coated with PEGylated liposomes and loaded with 5-FU to increase the in vitro antitumor effect (5.8-fold), arrest in S phase (2.3-fold) and apoptosis (1.9-fold) of the drug.81 Moreover, Ye et al developed a peptide nanoformulation using DOX-loaded fetal bovine serum (BSA) NPs functionalized with Cetuximab83 and Agbana et al generated carfilzomib nanocarriers consisting of a polypeptide coated with PEG and targeted against the EpCAM molecule. In this last case, cell viability assays against the DLD-1 CRC adenocarcinoma line showed a 4.5-fold increase in cytotoxicity of free 5-FU.66

As shown in Table 4, 5 studies focused on the chemotherapeutics 5-FU and DOX. In addition, PTX,84 carfilzomib,66 gamboic acid85 and niclosamide80 were also analyzed (each one in 1 article). For instance, Gener et al synthesized PEGylated PLGA-polymeric micelles coated with Cetuximab to transport PTX that showed a 4.4- and 16.5-fold reduction in the IC50 value in the HCT-8 CRC cell line compared to the non-targeted nanoformulation and free PTX, respectively,84 Cetuximab was the most commonly used monoclonal antibody to target nanoformulations (3 articles). Anti-EpCAM,66 the Fab fragment of anti-CD44v6 antibodies,80 and a recombinant protein composed of an anti-EGFR antibody and the RGD peptide85 were also used for active targeting. The HCT-116 cell line was the most widely used to assess therapeutic efficacy, followed by colonspheres. In fact, 5-FU loaded polymeric micelles functionalized against SMC2, a central component of the condensin complex involved in DNA supercoiling, demonstrated a 4-fold higher cytotoxicity in HCT-116 colonspheres, but did not show differences in antitumor effect against HCT-116 adherent cells compared to free 5-FU.86

Finally, although involving a more complex synthesis process compared to other nanoformulations, some hybrid NPs were developed with the aim of benefiting from the greatest possible number of advantages in a single nanocarrier. For example, Chen et al synthesized mesoporous silica NPs with an easily modifiable surface area, but they tended to form aggregates under physiological conditions, leading to low hemocompatibility. The addition of a liposomal shell allowed the generation of hybrid NPs with high biocompatibility, stability and controlled release that showed promising results in targeted therapy against CRC.81

Conclusion

The use of targeted nanoformulations as tumor-selective delivery systems in CRC is a promising strategy to improve antitumor efficacy. The administration of current antitumor drugs is limited by several obstacles such as rapid excretion, degradation and whole-body distribution, with the subsequent development of side effects. Regarding the latter limitation, site-specific drug delivery is necessary. An active targeting delivery system using monoclonal antibodies is an excellent way to direct pharmacological agents against tumor cells, as it induces a higher level of cell internalization compared to conventional delivery systems. In recent years, significant progress in the synthesis of high quality NPs (eg, composition, shapes, and sizes) has been made. Specifically, in CRC therapy, the NPs most frequently associated with monoclonal antibodies include polymeric NPs, lipid nanoformulations (liposomes and micelles) and inorganic NPs. In relation to the drugs most commonly used for testing antibody-functionalized nanoformulations in CRC, 5-FU and DOX were the two most outstanding chemotherapeutic agents in the studies analyzed. Furthermore, the most frequently used antibody for the generation of these nanoformulations was Cetuximab, which recognizes EGFR. Finally, a wide variety of CRC cell lines were used in in vitro assays to determine the efficacy of the newly synthesized nanopharmaceuticals, but the HT-116 cell line was by far the most commonly used. Future research is likely to include extensive development of nanoformulations with multi-ligand binding systems and exquisite specificity. However, although most of the studies included in this review yielded positive results, further assays are needed to demonstrate the benefits of novel drugs regarding bioavailability, biodegradability and biocompatibility in in vivo CRC models.

Future Perspectives

Despite recent advances in nanoformulations, including targeting systems for CRC treatment, many challenges remain to obtain a feasible alternative to conventional chemotherapy. In the case of antibody-linked nanoformulations, their specificity and efficacy still need to be greatly improved, since in many cases the tumor cells can escape the action of the nanodrug. This may be especially relevant in the case of CSCs, which show intense drug resistance and the ability to induce recurrence and metastasis. In addition, it should be noted that the toxicity of NPs may be increased by the use of antibodies. Damage to normal cells that express specific antigens recognized by the nanodrugs may limit their use. The success of this new targeted therapy against CRC will depend on the development of strategies to solve this problem. In any case, as new targeting strategies continue to be developed, more effective NP platforms will be applied in the treatment of CRC.

Acknowledgments

A.C. and F.Q. to acknowledge FPU2017 and FPU2019 grants (ref. FPU17/02977 and FPU19/04112) from the Ministerio de Educación y Formación Profesional (Spain) and C.L. a P-FIS (ref. PI20-00284) from the Ministerio de Ciencia, Innovación y Universidades (Spain). This research was funded in part by Junta de Andalucía (A-CTS-666-UGR20, P20_00540, B-CTS-122-UGR20 and PYC20-RE-035 UGR) FEDER program and Instituto de Salud Carlos III(PI19/01478-FEDER).

Disclosure

The authors declare that they have no conflicts of interest in this work.

References

1. Mattiuzzi C, Lippi G. Current cancer epidemiology. J Epidemiol Glob Health. 2019;9(4):217–222. doi:10.2991/jegh.k.191008.001

2. Thomas L, Roberts L. Cancer Today: Origins, Prevention, and Treatment. ed. Washington DC, USA: National Academic Press; 1984. doi:10.17226/18700

3. Sawicki T, Ruszkowska M, Danielewicz A, Niedźwiedzka E, Arłukowicz T, Przybyłowicz KE. A review of colorectal cancer in terms of epidemiology, risk factors, development, symptoms and diagnosis. Cancers. 2021;13(9). doi:10.3390/cancers13092025

4. Araghi M, Soerjomataram I, Jenkins M, et al. Global trends in colorectal cancer mortality: projections to the year 2035. Int J Cancer. 2019;144(12):2992–3000. doi:10.1002/ijc.32055

5. Modest DP, Pant S, Sartore-Bianchi A. Treatment sequencing in metastatic colorectal cancer. Eur J Cancer. 2019;109:70–83. doi:10.1016/J.EJCA.2018.12.019

6. Berman RS, Valerie L, Ryan DP. Tratamiento del cáncer de colon (PDQ®)–Versión para profesionales de salud - Instituto Nacional Del Cáncer; 2014:6237. Available from: https://www.cancer.gov/espanol/tipos/colorrectal/pro/tratamiento-colorrectal-pdq#_105. Accessed November 17, 2021.

7. Leal F, Ferreira FP, Sasse AD. FOLFOXIRI regimen for metastatic colorectal cancer: a systematic review and meta-analysis. Clin Colorectal Cancer. 2017;16(4):405–409.e2. doi:10.1016/j.clcc.2017.03.012

8. Geredeli C, Yasar N. FOLFIRI plus panitumumab in the treatment of wild-type KRAS and wild-type NRAS metastatic colorectal cancer. World J Surg Oncol. 2018;16(1):67. doi:10.1186/s12957-018-1359-9

9. Qin S, Li J, Wang L, et al. Efficacy and tolerability of first-line cetuximab plus leucovorin, fluorouracil, and oxaliplatin (FOLFOX-4) versus FOLFOX-4 in patients with RASwild-type metastatic colorectal cancer: the open-label, randomized, Phase III TAILOR trial. J Clin Oncol. 2018;36(30):3031–3039. doi:10.1200/JCO.2018.78.3183

10. Morano F, Sclafani F. Duration of first-line treatment for metastatic colorectal cancer: translating the available evidence into general recommendations for routine practice. Crit Rev Oncol Hematol. 2018;131:53–65. doi:10.1016/j.critrevonc.2018.08.006

11. Lee GY, Lee JS, Son CG, Lee NH. Combating drug resistance in colorectal cancer using herbal medicines. Chin J Integr Med. 2021;27(7):551–560. doi:10.1007/s11655-020-3425-8

12. Garbayo E, Pascual-Gil S, Rodríguez-Nogales C, Saludas L, Estella-Hermoso de Mendoza A, Blanco-Prieto MJ. Nanomedicine and drug delivery systems in cancer and regenerative medicine. Wiley Interdiscip Rev. 2020;12(5). doi:10.1002/wnan.1637

13. Aghebati-Maleki A, Dolati S, Ahmadi M, et al. Nanoparticles and cancer therapy: perspectives for application of nanoparticles in the treatment of cancers. J Cell Physiol. 2020;235(3):1962–1972. doi:10.1002/jcp.29126

14. Narum SM, Le T, Le DP, et al. Passive targeting in nanomedicine: fundamental concepts, body interactions, and clinical potential. In: Chung EJ, Leon L, editors. Micro and Nano Technologies. Elsevier; 2020:37–53. doi:10.1016/B978-0-12-816662-8.00004-7

15. Li Y, Xu X. Nanomedicine solutions to intricate physiological-pathological barriers and molecular mechanisms of tumor multidrug resistance. J Control Release. 2020;323:483–501. doi:10.1016/j.jconrel.2020.05.007

16. Cabeza L, Perazzoli G, Mesas C, et al. Nanoparticles in colorectal cancer therapy: latest in vivo assays, clinical trials, and patents. AAPS PharmSciTech. 2020;21(5):1–15. doi:10.1208/s12249-020-01731-y

17. Sesarman A, Tefas L, Sylvester B, et al. Co-delivery of curcumin and doxorubicin in PEGylated liposomes favored the antineoplastic C26 murine colon carcinoma microenvironment. Drug Deliv Transl Res. 2018;9(1):260–272. doi:10.1007/S13346-018-00598-8

18. Bauleth-Ramos T, Feijão T, Gonçalves A, et al. Colorectal cancer triple co-culture spheroid model to assess the biocompatibility and anticancer properties of polymeric nanoparticles. J Control Release. 2020;323:398–411. doi:10.1016/j.jconrel.2020.04.025

19. Jiang Y, Guo Z, Fang J, et al. A multi-functionalized nanocomposite constructed by gold nanorod core with triple-layer coating to combat multidrug resistant colorectal cancer. Mater Sci Eng: C. 2020;107:110224. doi:10.1016/J.MSEC.2019.110224

20. Tian X, Nguyen M, Foote HP, et al. CRLX101, a nanoparticle–drug conjugate containing camptothecin, improves rectal cancer chemoradiotherapy by inhibiting DNA repair and HIF1α. Cancer Res. 2017;77(1):112–122. doi:10.1158/0008-5472.CAN-15-2951

21. Titze-de-Almeida R, David C, Titze-de-Almeida SS. The race of 10 synthetic RNAi-based drugs to the pharmaceutical market. Pharm. 2017;34(7):1339–1363. doi:10.1007/s11095-017-2134-2

22. Tarudji AW, Kievit FM. Active targeting and transport. In: Chung EJ, Leon L, editors. Micro and Nano Technologies. Elsevier; 2020:19–36. doi:10.1016/B978-0-12-816662-8.00003-5

23. Swierczewska M, Crist RM, McNeil SE. Evaluating nanomedicines: obstacles and advancements. Methods Mol Biol. 2018;3–16. doi:10.1007/978-1-4939-7352-1_1

24. Ayati A, Moghimi S, Salarinejad S, Safavi M, Pouramiri B, Foroumadi A. A review on progression of epidermal growth factor receptor (EGFR) inhibitors as an efficient approach in cancer targeted therapy. Bioorg Chem. 2020;99:103811. doi:10.1016/j.bioorg.2020.103811

25. Li Y, Du Y, Liang X, et al. EGFR-targeted liposomal nanohybrid cerasomes: theranostic function and immune checkpoint inhibition in a mouse model of colorectal cancer. Nanoscale. 2018;10(35):16738–16749. doi:10.1039/c8nr05803b

26. Liszbinski RB, Romagnoli GG, Gorgulho CM, Basso CR, Pedrosa VA, Kaneno R. Anti-EGFR-coated gold nanoparticles in vitro carry 5-fluorouracil to colorectal cancer cells. Materials. 2020;13(2). doi:10.3390/ma13020375

27. Eloy JO, Ruiz A, de Lima FT, et al. EGFR-targeted immunoliposomes efficiently deliver docetaxel to prostate cancer cells. Colloids Surf. 2020;194:111185. doi:10.1016/j.colsurfb.2020.111185

28. Reda M, Ngamcherdtrakul W, Gu S, et al. PLK1 and EGFR targeted nanoparticle as a radiation sensitizer for non-small cell lung cancer. Cancer Lett. 2019;467:9–18. doi:10.1016/j.canlet.2019.09.014

29. Petrilli R, Eloy JO, Saggioro FP, et al. Skin cancer treatment effectiveness is improved by iontophoresis of EGFR-targeted liposomes containing 5-FU compared with subcutaneous injection. J Control Release. 2018;283:151–162. doi:10.1016/j.jconrel.2018.05.038

30. Wu F, Yue L, Su H, Wang K, Yang L, Zhu X. Carbon dots @ platinum porphyrin composite as theranostic nanoagent for efficient photodynamic cancer therapy. Nanoscale Res Lett. 2018;13. doi:10.1186/s11671-018-2761-5

31. Zhu X, Wong WK, Wu F. Conjugated porphyrin carbon quantum dots for targeted photodynamic therapy; 2018:50.

32. Lim JR, Mouawad J, Gorton OK, Bubb WA, Kwan AH. Cancer stem cell characteristics and their potential as therapeutic targets. Med Oncol. 2021;38(7). doi:10.1007/S12032-021-01524-8

33. Lazer LM, Kesavan Y, Gor R, et al. Targeting colon cancer stem cells using novel doublecortin like kinase 1 antibody functionalized folic acid conjugated hesperetin encapsulated chitosan nanoparticles. Colloids Surf B Biointerfaces. 2022;217:112612. doi:10.1016/J.COLSURFB.2022.112612

34. Dadashi Noshahr K, Shamsi F, Valtchev P, et al. Optimization of post-insertion method to conjugate Doxil with anti-CD133 monoclonal antibodies: investigating the specific binding and cytotoxicity to colorectal cancer cells in vitro. Saudi Pharm J. 2020;28(11):1392–1401. doi:10.1016/j.jsps.2020.09.003

35. Yin H, Xiong G, Guo S, et al. Delivery of Anti-miRNA for triple-negative breast cancer therapy using RNA nanoparticles targeting stem cell marker CD133. Mol Ther. 2019;27(7):1252–1261. doi:10.1016/j.ymthe.2019.04.018

36. Mi Y, Huang Y, Deng J. The enhanced delivery of salinomycin to CD133+ ovarian cancer stem cells through CD133 antibody conjugation with poly(Lactic-co-glycolic acid)-poly(ethylene glycol) nanoparticles. Oncol Lett. 2018;15(5):6611–6621. doi:10.3892/ol.2018.8140

37. Mohd-Zahid MH, Zulkifli SN, Che Abdullah CA, et al. Gold nanoparticles conjugated with anti-CD133 monoclonal antibody and 5-fluorouracil chemotherapeutic agent as nanocarriers for cancer cell targeting. RSC Adv. 2021;11(26):16131–16141. doi:10.1039/d1ra01093j

38. Muka T, Glisic M, Milic J, et al. 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(1):49–60. doi:10.1007/s10654-019-00576-5

39. 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(13):1098–1105. doi:10.1002/asi.10105

40. Cohen J. Weighted kappa: nominal scale agreement provision for scaled disagreement or partial credit. Psychol Bull. 1968;70(4):213–220. doi:10.1037/h0026256

41. Wanden-Berghe C, Sanz-Valero J. Systematic reviews in nutrition: standardized methodology. Br J Nutr. 2012;107(SUPPL. 2). doi:10.1017/S0007114512001432

42. Arabi L, Badiee A, Mosaffa F, Jaafari MR. Targeting CD44 expressing cancer cells with anti-CD44 monoclonal antibody improves cellular uptake and antitumor efficacy of liposomal doxorubicin. J Control Release. 2015;220(PtA):275–286. doi:10.1016/j.jconrel.2015.10.044

43. Scavo MP, Cutrignelli A, Depalo N, et al. Effectiveness of a controlled 5-fu delivery based on fzd10 antibody-conjugated liposomes in colorectal cancer in vitro models. Pharmaceutics. 2020;12(7):1–19. doi:10.3390/pharmaceutics12070650

44. Cortese K, Marconi S, Aiello C, et al. Liposomes loaded with the proteasome inhibitor z-leucinyl-leucinyl-norleucinal are effective in inducing apoptosis in colorectal cancer cell lines. Membranes. 2020;10(5). doi:10.3390/membranes10050091

45. Yao Y, Feng L, Wang Z, Chen H, Tan N. Programmed delivery of cyclopeptide RA-V and antisense oligonucleotides for combination therapy on hypoxic tumors and for therapeutic self-monitoring. Biomater Sci. 2020;8(1):256–265. doi:10.1039/c9bm00905a

46. Du Y, Liang X, Li Y, et al. Liposomal nanohybrid cerasomes targeted to PD-L1 enable dual-modality imaging and improve antitumor treatments. Cancer Lett. 2018;414:230–238. doi:10.1016/j.canlet.2017.11.019

47. Udofot O, Affram K, Smith T, et al. Pharmacokinetic, biodistribution and therapeutic efficacy of 5-fluorouracil-loaded pH-sensitive PEGylated liposomal nanoparticles in HCT-116 tumor bearing mouse. J N Sci. 2016;2(1):43.

48. Corvo ML, Mendo AS, Figueiredo S, et al. Liposomes as delivery system of a Sn(IV) complex for cancer therapy. Pharm Res. 2016;33(6):1351–1358. doi:10.1007/s11095-016-1876-6

49. Tummala S, Gowthamarajan K, Satish Kumar MN, et al. Formulation and optimization of oxaliplatin immuno-nanoparticles using Box–Behnken design and cytotoxicity assessment for synergistic and receptor-mediated targeting in the treatment of colorectal cancer. Artif Cells Nanomed Biotechnol. 2016;44(8):1835–1850. doi:10.3109/21691401.2015.1111226

50. Zalba S, Contreras AM, Haeri A, et al. Cetuximab-oxaliplatin-liposomes for epidermal growth factor receptor targeted chemotherapy of colorectal cancer. J Control Release. 2015;210:26–38. doi:10.1016/j.jconrel.2015.05.271

51. Limasale YDP, Tezcaner A, Özen C, Keskin D, Banerjee S. Epidermal growth factor receptor-targeted immunoliposomes for delivery of celecoxib to cancer cells. Int J Pharm. 2015;479(2):364–373. doi:10.1016/j.ijpharm.2015.01.016

52. Liang B, Shahbaz M, Wang Y, et al. Integrinβ6-targeted immunoliposomes mediate tumor-specific drug delivery and enhance therapeutic efficacy in colon carcinoma. Clin Cancer Res. 2015;21(5):1183–1195. doi:10.1158/1078-0432.CCR-14-1194

53. Lozano N, Al-Ahmady ZS, Beziere NS, Ntziachristos V, Kostarelos K. Monoclonal antibody-targeted PEGylated liposome-ICG encapsulating doxorubicin as a potential theranostic agent. Int J Pharm. 2015;482(1–2):2–10. doi:10.1016/j.ijpharm.2014.10.045

54. Wicki A, Rochlitz C, Orleth A, et al. Targeting tumor-associated endothelial cells: anti-VEGFR2 immunoliposomes mediate tumor vessel disruption and inhibit tumor growth. Clin Cancer Res. 2012;18(2):454–464. doi:10.1158/1078-0432.CCR-11-1102

55. Dwi Y, Limasale P, Tezcaner A, Özen C, Keskin D, Banerjee S. Pharmaceutical nanotechnology Epidermal growth factor receptor-targeted immunoliposomes for delivery of celecoxib to cancer cells. Int J Pharmaceut. 2015;24:351–359. doi:10.1016/j.ijpharm.2015.01.016

56. Almeida B, Nag OK, Rogers KE, Delehanty JB. Recent progress in bioconjugation strategies for liposome-mediated drug delivery. Molecules. 2020;25(23). doi:10.3390/MOLECULES25235672

57. Fan Y, Marioli M, Zhang K. Analytical characterization of liposomes and other lipid nanoparticles for drug delivery. J Pharm Biomed Anal. 2021;192. doi:10.1016/j.jpba.2020.113642

58. Mitchell MJ, Billingsley MM, Haley RM, Wechsler ME, Peppas NA, Langer R. Engineering precision nanoparticles for drug delivery. Nat Rev Drug Discov. 2021;20(2):1. doi:10.1038/S41573-020-0090-8

59. Pereira I, Sousa F, Kennedy P, Sarmento B. Carcinoembryonic antigen-targeted nanoparticles potentiate the delivery of anticancer drugs to colorectal cancer cells. Int J Pharm. 2018;549(1–2):397–403. doi:10.1016/j.ijpharm.2018.08.016

60. Yalikong A, Li XQ, Zhou PH, et al. A triptolide loaded her2-targeted nano-drug delivery system significantly suppressed the proliferation of her2-positive and braf mutant colon cancer. Int J Nanomed. 2021;16:2323–2335. doi:10.2147/IJN.S287732

61. Wu P, Zhu H, Zhuang Y, Sun X, Gu N. Combined therapeutic effects of131I-Labeled and 5Fu-loaded multifunctional nanoparticles in colorectal cancer. Int J Nanomed. 2020;15:2777–2787. doi:10.2147/IJN.S215137

62. Ning ST, Lee SY, Wei MF, et al. Targeting colorectal cancer stem-like cells with anti-CD133 antibody-conjugated SN-38 nanoparticles. ACS Appl Matera Interfaces. 2016;8(28):17793–17804. doi:10.1021/acsami.6b04403

63. Sabra R, Billa N, Roberts CJ. Cetuximab-conjugated chitosan-pectinate (modified) composite nanoparticles for targeting colon cancer. Int J Pharm. 2019;572. doi:10.1016/j.ijpharm.2019.118775

64. Fay F, McLaughlin KM, Small DM, et al. Conatumumab (AMG 655) coated nanoparticles for targeted pro-apoptotic drug delivery. Biomaterials. 2011;32(33):8645–8653. doi:10.1016/j.biomaterials.2011.07.065

65. Schmid D, Fay F, Small DM, et al. Efficient drug delivery and induction of apoptosis in colorectal tumors using a death receptor 5-targeted nanomedicine. Mol Ther. 2014;22(12):2083–2092. doi:10.1038/mt.2014.137

66. Agbana P, Lee MJ, Rychahou P, Kim KB, Bae Y. Ternary polypeptide nanoparticles with improved encapsulation, sustained release, and enhanced in vitro efficacy of carfilzomib. Pharm Res. 2020;37(11). doi:10.1007/s11095-020-02922-9

67. Jarai BM, Kolewe EL, Stillman ZS, Raman N, Fromen CA. Polymeric nanoparticles. In: Chung EJ, Leon L, Rinaldi C, editors. Micro and Nano Technologies. Elsevier; 2020:303–324. doi:10.1016/B978-0-12-816662-8.00018-7

68. Gao Y, Gu S, Zhang Y, et al. The architecture and function of monoclonal antibody-functionalized mesoporous silica nanoparticles loaded with mifepristone: repurposing abortifacient for cancer metastatic chemoprevention. Small. 2016;12(19):2595–2608. doi:10.1002/smll.201600550

69. Xu M, Wen Y, Liu Y, et al. Hollow mesoporous ruthenium nanoparticles conjugated bispecific antibody for targeted anti-colorectal cancer response of combination therapy. Nanoscale. 2019;11(19):9661–9678. doi:10.1039/c9nr01904a

70. Ahn JC, Teng PC, Chen PJ, et al. Detection of circulating tumor cells and their implications as a biomarker for diagnosis, prognostication, and therapeutic monitoring in hepatocellular carcinoma. Hepatology. 2021;73(1):422–436. doi:10.1002/HEP.31165

71. Lo PY, Lee GY, Zheng JH, Huang JH, Cho EC, Lee KC. GFP Plasmid and chemoreagent conjugated with graphene quantum dots as a novel gene delivery platform for colon cancer inhibition in vitro and in vivo. ACS Appl Bio Mater. 2020;3(9):5948–5956. doi:10.1021/acsabm.0c00631

72. Pedrosa P, Corvo ML, Ferreira-Silva M, et al. Targeting cancer resistance via multifunctional gold nanoparticles. Int J Mol Sci. 2019;20(21). doi:10.3390/ijms20215510

73. Emami F, Banstola A, Vatanara A, et al. Doxorubicin and anti-PD-L1 antibody conjugated gold nanoparticles for colorectal cancer photochemotherapy. Mol Pharm. 2019;16(3):1184–1199. doi:10.1021/acs.molpharmaceut.8b01157

74. Lin YW, Raj EN, Liao WS, et al. Co-delivery of paclitaxel and cetuximab by nanodiamond enhances mitotic catastrophe and tumor inhibition. Sci Rep. 2017;7(1). doi:10.1038/s41598-017-09983-8

75. Fernandes AR, Jesus J, Martins P, et al. Multifunctional gold-nanoparticles: a nanovectorization tool for the targeted delivery of novel chemotherapeutic agents. J Control Release. 2017;245:52–61. doi:10.1016/j.jconrel.2016.11.021

76. Tummala S, Kumar MNS, Pindiprolu SK. Improved anti-tumor activity of oxaliplatin by encapsulating in anti-DR5 targeted gold nanoparticles. Drug Deliv. 2016;23(9):3505–3519. doi:10.1080/10717544.2016.1199606

77. Yue D, Sun X. Ixazomib promotes CHOP-dependent DR5 induction and apoptosis in colorectal cancer cells. Cancer Biol Ther. 2019;20(3):284–294. doi:10.1080/15384047.2018.1529095

78. Singh P, Pandit S, Mokkapati V, Garg A, Ravikumar V, Mijakovic I. Molecular sciences gold nanoparticles in diagnostics and therapeutics for human cancer. Int J Mol Sci. 2018. doi:10.3390/ijms19071979

79. Zhou H, Ge J, Miao Q, et al. Biodegradable inorganic nanoparticles for cancer theranostics: insights into the degradation behavior. Bioconjug Chem. 2020;31(2):315–331. doi:10.1021/ACS.BIOCONJCHEM.9B00699

80. Andrade F, Rafael D, Vilar-Hernández M, et al. Polymeric micelles targeted against CD44v6 receptor increase niclosamide efficacy against colorectal cancer stem cells and reduce circulating tumor cells in vivo. J Control Release. 2021;331:198–212. doi:10.1016/j.jconrel.2021.01.022

81. Chen R, Huang Y, Wang L, et al. Cetuximab functionalization strategy for combining active targeting and antimigration capacities of a hybrid composite nanoplatform applied to deliver 5-fluorouracil: toward colorectal cancer treatment. Biomater Sci. 2021;9(6):2279–2294. doi:10.1039/d0bm01904f

82. Lee SM, Kim HJ, Kim SY, et al. Drug-loaded gold plasmonic nanoparticles for treatment of multidrug resistance in cancer. Biomaterials. 2014;35(7):2272–2282. doi:10.1016/j.biomaterials.2013.11.068

83. Ye Z, Zhang Y, Liu Y, Liu Y, Tu J, Shen Y. Egfr targeted cetuximab-valine-citrulline (Vc)-doxorubicin immunoconjugates-loaded bovine serum albumin (bsa) nanoparticles for colorectal tumor therapy. Int J Nanomed. 2021;16:2443–2459. doi:10.2147/IJN.S289228

84. Gener P, Gouveia LP, Sabat GR, et al. Fluorescent CSC models evidence that targeted nanomedicines improve treatment sensitivity of breast and colon cancer stem cells. Nanomed. 2015;11(8):1883–1892. doi:10.1016/j.nano.2015.07.009

85. Zhang Z, Qian H, Huang J, et al. Anti-EGFR-iRGD recombinant protein modified biomimetic nanoparticles loaded with gambogic acid to enhance targeting and antitumor ability in colorectal cancer treatment. Int J Nanomed. 2018;13:4961–4975. doi:10.2147/IJN.S170148

86. Montero S, Seras-Franzoso J, Andrade F, et al. Intracellular delivery of anti-SMC2 antibodies against cancer stem cells. Pharmaceutics. 2020;12(2). doi:10.3390/pharmaceutics12020185

Creative Commons License © 2022 The Author(s). This work is published and licensed by Dove Medical Press Limited. The full terms of this license are available at https://www.dovepress.com/terms.php and incorporate the Creative Commons Attribution - Non Commercial (unported, v3.0) License. By accessing the work you hereby accept the Terms. Non-commercial uses of the work are permitted without any further permission from Dove Medical Press Limited, provided the work is properly attributed. For permission for commercial use of this work, please see paragraphs 4.2 and 5 of our Terms.