Next Article in Journal
Effects of Absorption Kinetics on the Catabolism of Melatonin Released from CAP-Coated Mesoporous Silica Drug Delivery Vehicles
Next Article in Special Issue
Microdosimetric Investigation and a Novel Model of Radiosensitization in the Presence of Metallic Nanoparticles
Previous Article in Journal
Novel Pyropheophorbide Phosphatydic Acids Photosensitizer Combined EGFR siRNA Gene Therapy for Head and Neck Cancer Treatment
Previous Article in Special Issue
Antitumor Activity of Nanoparticles Loaded with PHT-427, a Novel AKT/PDK1 Inhibitor, for the Treatment of Head and Neck Squamous Cell Carcinoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Peptide-Conjugated Nano Delivery Systems for Therapy and Diagnosis of Cancer

1
School of Chinese Medicine, Hong Kong Baptist University, Hong Kong, China
2
Centre for Regenerative Medicine and Health, Hong Kong Institute of Science and Innovation-CAS Limited, Hong Kong, China
3
Mr. & Mrs. Ko Chi-Ming Centre for Parkinson’s Disease Research, School of Chinese Medicine, Hong Kong Baptist University, Hong Kong, China
4
National Institute for Locomotor Disabilities (Divyangjan), Kolkata 700090, India
5
School of Basic and Applied Science, Galgotias University, Greater Noida 203201, India
6
Changshu Research Institute, Hong Kong Baptist University, Changshu Economic and Technological Development (CETD) Zone, Changshu 215500, China
*
Author to whom correspondence should be addressed.
Pharmaceutics 2021, 13(9), 1433; https://doi.org/10.3390/pharmaceutics13091433
Submission received: 5 July 2021 / Revised: 31 August 2021 / Accepted: 7 September 2021 / Published: 9 September 2021
(This article belongs to the Special Issue Nanocarriers for Cancer Therapy and Diagnosis)

Abstract

:
Peptides are strings of approximately 2–50 amino acids, which have gained huge attention for theranostic applications in cancer research due to their various advantages including better biosafety, customizability, convenient process of synthesis, targeting ability via recognizing biological receptors on cancer cells, and better ability to penetrate cell membranes. The conjugation of peptides to the various nano delivery systems (NDS) has been found to provide an added benefit toward targeted delivery for cancer therapy. Moreover, the simultaneous delivery of peptide-conjugated NDS and nano probes has shown potential for the diagnosis of the malignant progression of cancer. In this review, various barriers hindering the targeting capacity of NDS are addressed, and various approaches for conjugating peptides and NDS have been discussed. Moreover, major peptide-based functionalized NDS targeting cancer-specific receptors have been considered, including the conjugation of peptides with extracellular vesicles, which are biological nanovesicles with promising ability for therapy and the diagnosis of cancer.

1. Introduction

Despite numerous advancements and breakthroughs in cancer therapy and diagnosis, cancer still ranks among the topmost causes of mortality in every country. In the year 2020, approximately 19.3 million new cases of cancer occurred, and merely cancer accounted for about 10 million deaths around the world [1]. Conventional strategies including chemotherapy are challenged by the low specificity and high toxicity toward cancer cells [2,3]. There are various biological barriers that hinder the successful delivery of therapeutic agents to the tumor site, for example, the tumor microenvironment (TME), mononuclear phagocytic system, extravasation of nanoparticles (NPs), cellular barriers, and drug efflux transporters [4], as portrayed in Figure 1.
How TME and hypoxia impede drug delivery? The complex nature of the TME is one of the crucial hindrances to the delivery of therapeutic agents to the tumor site [5]. The TME consists of cellular components such as cancerous and noncancerous stromal cells, blood vessels, lymphatic vessels, and immune cells. In addition, the non-cellular components of TME are composed of cytokines, chemokines, mediators, and growth factors, which are generally affected by the growth of cancer cells [6]. The extracellular matrix (ECM) is another key element of the TME, which differs significantly in terms of composition and framework compared to that under normal tissue. The ECM in the TME is highly abundant, stiffer, and denser, forming another bottleneck to cancer therapy via shielding the cells from anti-cancer drugs. Moreover, the enhanced stiffness of ECM in hypoxic TME has been found to activate the antiapoptotic pathways and contribute toward the development of drug resistance in cancer cells [7].
Therapies such as photodynamic and radiotherapy depend on oxygen, which is restricted by hypoxic TME. When the eruptive growth of the cancer cells occurs, the supply of oxygen and the nutrient is restricted from their neighboring blood vessels [8]. Under hypoxia, the transcriptional factor, hypoxia-inducible factor-1α (HIF-1α), induces the metabolic change from oxidative phosphorylation to aerobic glycolysis, which is referred to as the Warburg effect [9]. The proliferation and glycolytic metabolism in the cancer cells enhance the development of excessive reactive oxygen species (ROS), which attack cellular components such as nucleic acid, causing genomic instability and thereby altering the morphology of the cell [9]. Notably, the ability of ROS to regulate cancer cell survival is found to be cell type specific, for example as observed in MCF-7 and MDA-MB-435 breast cancer cells [10]. In addition to the effect of ROS on cell proliferation, the ROS-mediated activation of extracellular-regulated kinase 1/2 (Erk1/2) are found to play an important role in the augmentation of cell survival, motility, and anchorage-dependent growth of multiple cancers, such as ovarian cancer, breast cancer, melanoma, and leukemia [10]. Such an occurrence, with upregulation of the efflux pump for secreting lactic acid and carbonic acid, leads to benefit the tumor cells as they live longer and succeed in their mission even in extreme condition [11]. Therefore, under hypoxic TME, the delivery of a therapeutic agent to the tumor site is obstructed [12,13]. Contrariwise, ROS have also been applied for therapeutics of cancer by designing strategies to enhance the cellular level of ROS exuberantly in order to include irrevocable damages, leading to the apoptosis of cancer cells. This can be accomplished via chemotherapy or radiotherapy depending on the cancer type. For example, a combinatorial therapy of pancreatic cancer with gemcitabine with trichostatin A, epigallocate-3-gallate (EGCG), capsaicin, and benzyl isothiocyanate (BITC) are found to be working via increasing the intracellular ROS level for triggering ROS [10]. In another research, it was found that gold(III) porphyrin 1a could be a potential anti-cancer lead by acting toward mitochondria, as ROS played a role in gold(III) porphyrin 1a-induced apoptosis [14]. In addition, photodynamic therapy using a synthetic photosensitizer, 5,10,15,20-tetra-sulfo-phenyl-porphyrin (TSPP), is found to enhance the generation of ROS, leading to the decrease in antioxidant capacity in tumor tissue [15]. In addition, palladium porphyrin complexes are also found to generate ROS with higher efficiency. Interestingly, the palladium porphyrin complex showed higher therapeutic activity as compared to free base porphyrin upon irradiation with light [16].
Mononuclear phagocytic system (MPS) as a barrier to drug delivery: To fetch the desired therapeutic response of the drug, its successful delivery at the tumor site is important, which again relies on the nature of the delivery system and its stability in the blood circulation. The blood carries various proteins including globulin, albumin, and fibrinogen. After entry of the nano delivery system (NDS) in the blood circulation, the blood serum proteins get adsorbed on their surface and form a complex, which is referred to as protein corona [17]. The process of forming protein corona is known as opsonization, which is generally followed by the phagocytosis via the macrophage, which is a type of immune cell in MPS [18]. Remarkably, the process of opsonization and phagocytosis by the MPS facilitates the elimination of NDS from the systemic blood circulation.
How extravasation reacts to NDS in the TME: The presence of the vascular endothelial layer is another hurdle, which is required to be overcome for the successful delivery of NDS at the tumor site. The vascular endothelial layer is composed of a semi-permeable lining of the inner walls of blood vessels. In addition, a proteoglycan layer of glycocalyx controls the permeability of molecules across the blood vessels [19]. The glycocalyx layer has been found to be involved in the enhanced interactions with cationic particles by providing a negative charge to the membrane of endothelial cells [20]. Therefore, the presence of glycocalyx is a limiting factor for the extravasation of NDS in the TME, as it potentially conceals the NDS [4,21]. In addition, there are other factors that affect the extravasation of NDS, such as the hydrodynamics of NDS, enhanced permeability, and retention, which favor the nano therapy of cancer [4].
Cellular barriers as a bane to the nano delivery system: The passage of NDS through the endothelium of the blood vessels into the target site is another obstacle. In general, the NDS cannot traverse through the endothelium; however, in disease conditions, such as cancer, the integrity of the endothelium is compromised owing to the activation of cytokines, and thereby, the endothelial cells’ gap is enhanced. Therefore, the NDS can reach the pathological site by traversing through the abnormal endothelial gaps. Unfortunately, after escaping the blood vessels associated with the endothelial barrier, the NDS confronts another hurdle while traversing through the dense interstitial space and extracellular matrix (ECM) to reach the target site. The composition of interstitial space including collagen and an elastic fiber network consisting of proteins and glycosaminoglycan, which form ECM, forms a hydrophilic gel by the interstitial fluid, which fills the interspersed spaces. Even though the ECM and interstitial space render structural integrity to the tissue, under pathological conditions, including cancer, the collagen content is bigger than that in healthy conditions. This suggests that the excessive firmness of ECM is a crucial barrier that could obstruct NDS delivery [22]. Notably, the charged particles have been found to possess enhanced interactions with the membrane, whereas uncharged particles—for example, PEGylated NDS—show lesser interaction due to the steric hindrance. This leads to the accumulation of NDS to form a cluster around the membrane and prevent the entry of successive NDS [4].
How drug-efflux transporters can pump out the therapeutic agents: Even though the NDS reach the target site after confronting various hurdles, there is a tiny fraction of those that could make it to therapeutic efficacy by exerting intracellular cytotoxicity. Interestingly, various solid tumors possess crucial machinery that facilitates the expulsion of drugs, which is often referred to as drug-efflux transporters. For example, the overexpression of P-glycoprotein (P-gp), a drug-efflux transporter, has been reported to be linked with the efflux of anti-cancer drugs, and also the clinical refractoriness of anti-cancer drugs is associated with P-gp [4]. In addition, there are several other hurdles associated with the obstruction of delivery of NDS to the target site, which have been extensively reviewed elsewhere [23].
Recently, the conjugation of peptides and NDS (CPNDS) has emerged as a versatile technique for multidisciplinary biomedical applications. Compared to antibodies-based targeted NDS, peptide-conjugated NDS offers various advances: for example, most of the therapeutic monoclonal antibodies (TMAs) do not target tumor-specific antigens (TSAs), it requires screening to select monoclonal antibodies for dominant epitopes, the target must be antigenic for conventional monoclonal antibodies, and it also depends on the strain of animals used. However, in case of peptides, the target is not necessarily antigenic, and there is no requirement of prior information about target molecules. In the context of intracellular transport, there is no selection criteria for TMAs, and it is difficult to select during the screening process; however, in case of peptide-based NDS, screening technologies offer a convenient selection of candidates, which could induce endocytosis rapidly. In the context of the conjugation process, only ≈50% of the monoclonal antibodies bind to the drug, making it difficult to predict the stoichiometry and drug position. Moreover, the conjugation chemistry is limited to aqueous solutions. On the other hand, in case of peptide-conjugated NDS, the augmented flexibility in conjugation chemistry for coupling to linker and drug allows a wider selection of drugs, including compounds that are insoluble in water. Notably, the significantly lower cost of production and enhanced product reproducibility make peptide-conjugated NDS a preferred choice compared to the antibody-based NDS [24].
The synergistic integration between peptides and NDS allows effective customization of their biological behaviors and facilitates overcoming the inherent limitations of the individual system. Past decades have witnessed the development of several types of CPNDS for various applications including therapeutic drug delivery and diagnostic imaging [25]. This work provides a comprehensive overview of the existing and latest technologies and their application for the development of CPNDS.

2. Techniques for Preparing CPNDS

The CPNDS can be prepared by the modification of as-prepared NDS by functionalization with various peptides. In general, the major strategies employed are the chemical conjugation method, ligand exchange method, and chemical reduction method.

2.1. Chemical Conjugation Method

In this method, the peptide of choice is attached to the NDS in two steps. First, the NDS is capped by stabilizers (by using either hydrophilic shells or PEG derivatives), which contain active groups that are suitable for binding peptides. Furthermore, the peptides can be conjugated on the surface of NDS via a reaction with stabilizers. This method has been found to be suitable for the immobilization of positively charged or neutral peptides on gold nanoparticles (AuNPs) capped with citrate [26,27,28]. Bartczak et al. demonstrated the conjugation of a positively charged KPQPRPLS peptide (which binds to epidermal growth factor receptor (EGFR)) to carboxy-terminated oligoethylene glycol stabilized AuNPs by employing an EDC/sulfo-NHS (1-ethyl-3-(3-dimethylaminopropyl) carbodiimide hydrochloride/N-hydroxy sulfosuccinimide) coupling technique [29]. In another research study, Fu et al. synthesized manganese-doped iron oxide NPs (MnIO NPs) by the functionalization of monocyclic peptide (MCP, the CXC chemokine receptor 4 (CXCR4) antagonist) [30].

2.2. Ligand Exchange Method

In this method, the existing ligand on the surface of NDS is displaced by the desired peptide ligand. This method is the simplest approach for functionalizing the surface of NDS with peptides [31,32]. This method has been extensively employed for preparing cysteine (Cys, C)-containing peptides functionalized AuNPs owing to the presence of the thiol group of cycteine, which is capable of forming a strong S-Au covalent bond with the surface of AuNPs [33,34,35,36,37]. Lévy et al. demonstrated that the cysteine–alanine–leucine–asparagine–asparagine (CALNN) pentapeptide is capable of converting citrate-capped AuNPs to stable and water-soluble AuNPs equipped with chemical features similar to proteins [37]. It has been shown that the CALNN peptide is mostly captured in the endoplasmic reticulum due to its higher affinity toward the ER signal and its capacity to penetrate the nucleus. Interestingly, the AuNPs modified with CALNN can be functionalized with various biomolecules including nucleic acid and biotin, which is applicable for biomedical application. Later, it was found that the Au-S covalent bond can be degraded by the thiol group, which is often found in the biological system. Notably, the Tang group resolved this limitation by developing a method for preparing peptide-functionalized AuNPs (peptide-Se-AuNPs) via the Au–Se bond in lieu of the Au–S bond by employing a peptide with Se-modified cysteine [38,39,40].

2.3. Chemical Reduction Method

This method involves three steps: first, the metal ion precursor is premixed with a peptide in a reaction solution. Second, a small amount of reducing agent is added to the reaction solution. Third, the as-prepared peptide-functionalized NPs are purified [41,42,43]. Notably, the peptide is responsible for reducing the metal ions and the stabilization of NPs. The presence of amino acid residues in the peptide, for example, tyrosine, aldehyde-functionalized proline, and tryptophan, are capable of reducing the metal ions to the metal via electron transfer [42,44]. Another research study showed that peptides could act as a stabilizing agent; however, other chemicals such as ascorbic acid and sodium borohydride can be used as reducing agents [45]. Corra et al. demonstrated that the HH-dL-dD-NH2 peptide can be employed as a capping agent to produce palladium NPs (PdNPs), platinum NPs (PtNPs), and AuNPs equipped with high monodispersed and colloidal stability in solution [43].

3. Peptide Conjugation of NDS for Therapy and Diagnosis of Cancer

A plethora of studies have shown the application of artificial bioactive peptides, and many of those have been commercialized [46,47]. Despite tremendous advancements, most peptides suffer from various limitations including lower binding affinity toward targets, lower selectivity compared to the proteins, susceptibility to digestion by proteases [48], and shorter half-life [49]. Interestingly, the integration of peptides with various non-biological materials such as small molecules, polymers, metals, and hydrogels have shown potential to resolve the inherent limitation of peptides [50,51]. Especially, NDS have shown promising capacity to form conjugates with peptide, which could not only alleviate the peptides’ function but also execute abiotic properties, leading to synergistic effects. Therefore, the CPNDS has been considered a promising tool for cancer therapy and diagnosis.
As noted previously, various factors such as hypoxic TME, MPS, cellular barrier, and drug-efflux transporters are major hurdles in nano delivery, and peptide-conjugated NDS have been found to be useful to overcome these scenarios. In the context of hypoxic TME, stimuli-responsive peptide-conjugated nano delivery systems have been developed. For example, pH-responsive insertion peptides possess feasible interactions with the cell membrane at neutral pH, but they can penetrate and form stable transmembrane complexes at acidic pH, which is suitable for targeting hypoxic TME [52]. To overcome the MPS, Tang et al. developed RES-specific blocking systems employing a “don’t-eat-us” approach, where a CD47-derived, enzyme-resistant peptide ligand was designed and placed on a d-self-peptide-labeled liposome (DSL). Interestingly, it facilitated the long-lasting masking of cell membranes, thereby reducing interactions between phagocytes and NDS [53]. Peptide-conjugated NDS have been found to be crucial to overcome the cellular barriers. There are many successful examples of peptide-conjugated particles helping in the targeted delivery of dug to the diseased cells and penetration across physiological barriers. For example, Georgieva et al. showed the conjugation of the G23 peptide to polymersomes for in vivo and in vitro delivery of therapeutic drug across the BBB [54]. Yao et al. reported that pDNA can be delivered across the BBB by conjugating dendrigraft poly-l-lysines (DGL) NP to poly (ethylene glycol) (PEG) and a LIM Kinase 2 derived cell-penetrating peptide (LNP) [55]. Peptide conjugation has been found to be effective in bypassing P-glycoprotein (P-gp), causing drug resistance [56,57,58].
For a long time, the selective targeted delivery of anti-cancer drugs to the target site has been a major bottleneck in cancer therapy. In the prevailing condition, peptides have shown a great potential for rendering targeted drug delivery selectively, warranting an alleviated performance for treating fatal diseases, including cancer [59,60]. NDS can be engineered via functionalization with specific peptides to achieve the targeted delivery of anti-cancer drugs to the target site (Figure 2). Table 1 enlists the promising CPNDS based on cancer type, their specific receptor, and the conjugated peptide. There are various receptors, which have been employed as a target for peptide-conjugated NDS for cancer therapy.

3.1. CPNDS Targeting Somatostatin Receptor

Somatostatin receptors (SSTR) are transmembrane GPCRs that have been found to be upregulated in several cancers, including adenocarcinoma and breast cancer [77,78,79]. Notably, the somatostatin peptide in its native form possesses a binding affinity toward SSTR, making them an alluring targeting agent for cancer treatment [80,81]. However, the somatostatin peptide possesses a shorter half-life owing to the enzymatic deterioration. Hence, octreotide was developed, which is an analog of the somatostatin peptide that can endure the enzymatic deterioration [80]. Various research groups employed the octreotide peptide-based functionalized NDS for cancer treatment [61,71]. Zhang et al. prepared octreotide-PEG-distearoylphosphatidylethanolamine (DSPE), followed by developing octreotide-modified PEGylated liposomes loaded with doxorubicin (DOX), which promoted the delivery of DOX via an intracellular route. Notably, octreotide-functionalized NDS showed enhanced toxicity toward SSTR2-positive cancer cells through endocytosis [71]. Another research by Chang et al. developed octreotide-functionalized PEGylated liposome loaded with cantharidin, which could efficiently induce the cell death of MCF7 breast cancer cells by specifically targeting somatostatin receptors and demonstrated the lowered toxicity as compared to cantharidin alone [61].
The SSTR-based diagnosis of cancer has also been demonstrated; for example, SSTR-based imaging of gastroenteropancreatic neuroendocrine tumors has been conducted by [111In-DTPA0]-octreotide(Octreoscan), octreotide chelator conjugates, 1,4,7,10-tetraazacyclodocecane-N,N′,N″,N‴-tetraacetic acid (DOTA)-d-Phe1-Tyr3-octreotide (DOTATOC), and DOTA-dPhe1-Tyr3-octreotate (DOTAT ATE), which showed enhanced affinity toward SSTR [82,83]. In another research, 89Zr- and gadolinium (Gd)-labeled PEGylated liposomes functionalized with octreotide, which demonstrated SSTr2-targeting specificity and dual PET/MR imaging features [84].

3.2. CPNDS Targeting Integrin Receptor

Integrin is a transmembrane heterodimeric protein essential for the regulation of the different biological functions of cancer cells, including cell–cell and cell–ECM interaction [85]. Among various forms of integrins, αvβ3, αvβ5, and α5β1 integrins are upregulated in cancer cells and associated with cancer cell phenotypes such as angiogenesis, tumor growth, and metastasis [86], suggesting that peptide-based ligands targeting integrins could be promising therapeutic agents for drug delivery as well as molecular imaging. One of the natural ligands of integrin is glycoproteins, which express themselves on the surface of the cell or protein of the extracellular matrix. Therefore, short peptide sequences that produce integrin-binding motives have gathered huge attention as a potential therapy; however, it was not found to pass the clinical trial successfully. Therefore, the integrin peptide ligand was alternatively used in conjugation with NDS for the specific delivery of drug to the cell, which is overexpressing the integrin receptor [87].
Various Arg–Gly–Asp (RGD)-based CPNDS were also developed as potential anti-cancer therapies and diagnostic probes [88,89,90]. For example, tripeptide RGD was reported as a ligand for αvβ3 integrin overexpressed in solid tumors [91]. The RGD-modified PEGylated liposome-encapsulated DOX enhanced drug accumulation in cancer cells by internalization through the integrin receptor-mediated endocytosis pathway and showed antitumor effects [74]. Furthermore, to enhance the targeting efficacy, cyclic RGD-modified PEGylated liposomes were developed; for example, c(RGDfK), c(RGDfC), and RGD10 (DGARYCRGDCFDG) were found to be more stable at neutral pH as compared to the noncyclic RGD peptide, which enabled them to resist proteolysis [63,88,92]. Additionally, they showed high affinity toward αvβ3 integrin in human BcaP-37 breast cancer, HT29 colon cancer, and A375 melanoma cells [63,93].
In the context of cancer diagnosis, RGD-modified probes have been developed, such as [18F] Galacto-RGD, [18F] Alfatide, [68Ga] NOTA-PRGD2, 99mTcHYNIC-3PEG4-E[c(RGDfK)2], and 64Cu-DOTA-QD-RGD, which allowed the visualization of tumors in vivo [94]. Moreover, [18F] Galacto-RGD did not accumulate in the normal brain, unlike 18F-fluorodeoxyglucose (FDG), when used clinically as a PET tracer, suggesting that the RGD PET tracer can be applied to the imaging of glioma. [18F] Alfatide showed a higher tumor/background ratio in brain metastases compared with before the affinity was optimized [95]. Integrin α5β1 shows potent anti-cancer activity, which is recognized by a non-RGD peptide, ATN-161 (Ac-Pro-His-Ser-Cys-Asn-NH2) [96]. By coupling the PEGylated DOX liposome and ATN-161 lysine analog, the ATN-161-modified PEGylated DOX liposome was produced. It was reported that the integrin-mediated endocytosis mediates the cellular uptake of the ANT-161-modified liposome. Thereby, the ATN-161-modified PEGylated DOX liposome showed the significant antitumor effect on breast cancer cells and human umbilical vein endothelial cells [62].

3.3. CPNDS Targeting Transferrin Receptor (TFR)

TFRs are transmembrane glycoproteins receptors that facilitate the iron uptake by interacting with transferrin, an iron-binding protein [97]. Since the TFR is found to be upregulated on the surface of various cancer cells including breast cancer, lung adenocarcinoma, glioma, and chronic lymphocytic leukemia, it became an attractive molecule for cancer therapeutics [98,99,100,101,102]. Interestingly, the transport of various substances including anti-cancer drugs across the blood–brain barrier (BBB) is found to be regulated via P-glycoprotein and tight junction [103]. As the expression level of TFR in the BBB is high, the NDS conjugated with TF can cross the BBB through receptor-based endocytosis. Research reported that dual targeting DOX liposomes conjugated with TF and folate yielded anti-cancer effects in C6 glioma cells [104]. Lee et al. developed peptide T7 (HAIYPRH) using a phagedisplay method and showed higher TFR binding activity compared with TF [105].
TFR also represents a unique target for the specific imaging of cancer cells, suggesting its applicability in the diagnosis of cancer progression. Zhang et al. developed a light-up probe TPETH-2T7 by conjugating a red-emissive photosensitizer with aggregation-induced emission (AIE) with peptide HAIYPRH(T7), enabling them to target TFR. The probe alone is non-emissive; however, it yields turn-on fluorescence in the presence of TfR. In vitro experiments showed that the probe specifically binds to TFR, which is overexpressed on the MDA-MB-231 breast cancer cells. Notably, the image-guided photodynamic cancer ablation is evidence of its cancer therapeutic ability as well [106]. Wang et al. developed self-assembled IR780-loaded transferrin NDS, which are applicable for imaging and targeting, and offered a combined value as photothermal and photodynamic therapy, which is suitable for cancer therapeutics [107]. Another class of transferrin, which is known as lactoferrin, has been found to be highly expressed in the BBB [108], and it possesses better permeability than transferrin [109,110]. Notably, Miao et al. functionalized lactoferrin to the surface of poly(ethylene glycol)-poly(lactic acid) nanoparticles to facilitate BBB/BBTB and glioma cell dual targeting. Interestingly, tLyP-1, a tumor-homing peptide, which contains a C-end Rule sequence that can facilitate tissue penetration via the neuropilin-1-dependent uptake pathway, was coadministrated with lactoferrin-functionalized NPs to augment its accumulation and deep penetration into the glioma parenchyma, suggesting its suitability for antiglioma drug delivery [111].

3.4. CPNDS Targeting the HER2 Receptor

HER2 is highly expressed in various cancers including breast cancer, gastric cancer, and ovarian cancer [112,113]. Trastuzumab, a recombinant monoclonal antibody, has been found to target specifically HER2 [114]. Additionally, combinatorial therapy with trastuzumab showed a higher anti-cancer therapeutic effect [115]. However, a tedious method for producing recombinant monoclonal antibodies makes it relatively costly. Conversely, the production of peptide-based ligands is cost-efficient and equipped with low antigenicity. Therefore, HER2-specific peptide ligands have gained attention; for example, Karasseva et al. developed KCCYSL peptide using the phagedisplaytechnique and demonstrated its activities against human breast and prostate cancer cells with HER2 overexpression [116]. In another research, the apH-responsive PEGylated DOX liposome was modified with KCCYSL, which could specifically bind to and internalize in HER2-positive cells, and then pH-tunable vesicles release DOX swiftly and significantly. Notably, this liposome inhibited the tumor growth in a breast cancer mouse model with HER2-positive BT474 breast cancer cells [64].
Another peptide AHNP (FCDGFYACYADVGGG) was created from a heavy-chain CDR3 loop of trastuzumab, which was found to have HER2-specific affinity [117]. In another research study, AHNP-PEG-DSPE was developed with three glycine amino acids, and it was applied to AHNP-modified PEGylated DOX liposomes. Notably, this liposome showed tumor inhibition properties in a breast cancer mouse model bearing HER2-positive TUBO cancer cells [65]. In the context of the diagnostic application, PEGylated chitosan-modified LTVSPWY (LTVSPWY-PEG-CS) was developed as an MRI imaging probe, which could detect cancer efficiently in vivo [76].

3.5. CPNDS Targeting Aminopeptidase N

Aminopeptidase N (or CD13) is associated with the growth of various cancers and suggested as a potential target for anti-cancer treatment. Interestingly, tripeptide Asn-Gly-Arg (NGR) is a ligand of aminopeptidase N (APN/CD13), which is found to be overexpressed in cancer cells and also target neoangiogenic blood vessels [118]. APN-targeted NDS have been developed by various groups; for example, after the intravenous injection of the c-Myc siRNA loaded in NGR-modified PEGylated liposomes, they are delivered efficiently to the HT1080 fibrosarcoma cytoplasm. Therefore, the result at the tumor site showed the suppression of c-Myc and evoked cellular apoptosis [54]. Antitumor activity was observed in HT1080 fibrosarcoma cells and HUVECs by the quantitative accumulation of docetaxel, which was loaded in NGR-modified PEG-b-PLA polymeric micelles [103]. When NGR, thermosensitive liposomes, and DOX were conjugated with CPP, it showed an inhibition of tumor growth in HT1080 fibrosarcoma cells [104]. If NGR peptides are conjugated with an imaging agent such as fluorescent dye, QDs, micelles, and liposomes show potential in visualizing the tumor. The glioma-associated vessels in a fluorescent imaging system were clearly shown, and CD31 were specifically recognized when PEGlyated CdSe/AnS QDs were modified with an NGR peptide [58].

3.6. CPNDS Targeting Luteinizing Hormone-Releasing Hormone (LHRH)

Another receptor, LHRH, is overexpressed in different cancers such as breast, colorectal, ovarian, and prostate cancers, and it is a crucial anti-cancer target [119,120]. Bajusz et al. developed LHRH-based peptides, SB-05, SB-86, SB-40, and SB-95 as cancer-specific ligands. Interestingly, these ligands showed high affinities toward the membrane receptors of human breast and prostate cancer cells as well as rat pituitary Dunning R-3327 prostate cancer cells [121,122]. AEZS-108 (previously known as AN-152), a hybrid molecule consisting of a synthetic peptide carrier covalently coupled to DOX, was found to facilitate the delivery of DOX specifically to cancer cells expressing LHRH, including in uveal melanoma [123] and prostate cancer [124]. Mingqiang et al. developed cisplatin-loaded LHRH-modified dextran NPs (Dex-SA-CDDP-LHRH), which could specifically target LHRH receptors overexpressed on the surface of 4T1 breast cancer cells [125].

3.7. CPNDS Targeting Epidermal Growth Factor Receptor (EGFR)

EGFR has been widely reported to be crucial for uncontrolled signal transduction associated with cellular growth [126]. Notably, the GE11 peptide binds specifically to EGFR, which is overexpressed in various cancers including breast cancer, lung cancer, and glioma [127]. Therefore, the GE11 peptide has been conjugated with different NDS; for example, Huang et al. developed GE11 peptide-conjugated liposomes loaded with the photosensitizer indocyanine green (ICG) and chemotherapy drug curcumin (CUR), which could demonstrate EGFR targeting as well as an anti-cancer effect [128]. Han et al. demonstrated that small peptide, AEYLR-conjugated, nano lipid carriers increased the specific cellular uptake in cancer cells with EGFR overexpression [129]. Mayr et al. synthesized platinum (IV) complexes conjugated with an EGFR-targeting peptide, LARLLT; however, it was found to be unsuitable for increasing the specific uptake of small-molecule drugs in cancer cells with overexpressed EGFR [130].

3.8. CPNDS Targeting Epithelial Cell Adhesion Molecule (EpCAM)

EpCAM (or CD326) is an epithelial cell marker that is frequently and most strongly expressed in tumor-associated antigens. It is expressed in various cancers including squamous cell carcinoma and adenocarcinoma [131]. Ma et al. demonstrated that the peptide SNFYMPL (SNF*) could target EpCAM. Next, they conjugated SNF* with poly(histidine)–PEG/DSPE copolymer micelles. Notably, SNF* labeling substantially enhanced the micelle binding with gastric adenocarcinoma and colon cancer cells and augmented the anti-cancer effects, and it also reduced the in vivo toxicities of the micelles. Therefore, SFN* peptide-based targeting paves the way for EpCAM-targeted cancer therapy as well as diagnosis [132].

3.9. CPNDS Targeting CD133

CD133 is commonly expressed in cancer stem cells from various cancers including glioma, colon cancer, prostate cancer, and lung cancer [133]. Yan et al. developed CD133 peptide-conjugated photosensitizer, CD133-pyropheophorbide-a (Pyro), which showed a targeted photodynamic effect in colorectal cancer stem cells (CRCSC). Conventional photosensitizers such as (Pyro) lack tumor selectivity, triggering unwanted toxicity to the nearby healthy tissue. Interestingly, CD133-Pyro augmented the targeting capacity of Pyro, and it was found that CD133-Pyro exhibits the targeted delivery ability both in CRCSCs and inhibited tumor growth in a mouse model, suggesting its applicability for the therapy of CRC via CRCSC targeting [134].

4. Cell-Penetrating Peptides (CPP)

Cell penetration of the peptide is classified into two categories. (A) On the basis of peptide origin, they are subdivided into three types: chimeric, protein derived, and synthetic. Chimeric CPPs are made of two different peptide motifs. Transportan is said to be chimeric CPP that has been derived from mastoparan and galanin. Examples of protein-derived CPPs are TAT and penetratin, which is a natural protein derivative. The synthetic peptides are of the polyarginine family [135]. (B) The second category of CPP classification is based on physiochemical property. Based on physiological property, there are three types of CPP: cationic, amphipathic, and hydrophobic. As a result of its positive charge, many CPPs are cationic. The example of cationic CPP is TAT, the transcriptional activator protein in HV-1 [136]. The amphipathic CPPs, because of the lysine residue in their structure, are the sequences with a high degree of amphipathicity: for example, Transportan, a 27 amino acid long peptide [137]. In case of hydrophobic CPP, only the hydrophobic motif or non-polar sequence are present [138].
Regarding the mechanism for the internalization of CPP, for the transportation of CPP across the biological membrane, the exact mechanism is still unclear. However, after going through certain literature, the outcome showed that there may be three possible pathways for CPP internalization into the membrane. The three most effective parameters for the internalization pathway of CPP into the cellular membrane are the peptide concentration, peptide sequence, and lipid component in each membrane [139,140].
On the basis of peptide concentration, the route for the uptake of different cationic CPPs varies. When the concentration is high, rapid cytosolic uptake is detected, and at the lower concentration of peptide, the mechanism of uptake is dominant [141,142]. The second influential parameter for the uptake mechanism of CPP is peptide sequence. The local concentration of TAT and penetratin, which are arginine-rich CPPs, in a biomembrane may be enhanced due to the highly positively charged CPPs [143,144]. For the internalization of CPPs, there are three possible mechanisms. (i) The first is direct penetration, which is an energy-independent pathway including various mechanisms including pore formation, a carpet-like model, and a membrane-thinning model [145,146]. (ii) The second mechanism is the endocytosis pathway, in which the transduction approach is energy dependent. In endocytosis, the inward folding of the plasma membrane takes place to carry material from outside of the cell and absorb them. The three different classes of endocytosis are pinocytosis, phagocytosis, and receptor-mediated endocytosis. (iii) The third mechanism is translocation through the formation of a transitory structure. In this, the interaction of CPP takes place with the cellular membrane, which causes the disruption of the lipid bilayer of the membrane following the formation of an inverted structure, the inverted micelles [147].

5. Conjugation of Peptides and Extracellular Vesicles (CPEVs) for Cancer Therapy

Extracellular vesicles (EVs) are nanovesicles with a size around 30–1000 nm, which are secreted from most of the cell types and are found in various biofluids including blood and urine [148,149,150]. Recently, EVs have emerged as a promising NDS with huge application in cancer therapy as well as diagnosis. A detailed review of the factors reacting with EV-based drug delivery systems has been reported by our group previously [151]. Interestingly, the surface modification of EVs has a great potential to achieve the targeting ability [152]. There are various methods that could be utilized to modify the surface of EVs to conjugate the ligand, such as physical approaches (sonication, extrusion, and freeze–thaw) that can change the surface properties of EVs via membrane rearrangements and biological approaches (genetically and metabolically engineering cells to express protein or cargo molecules of interest in secreted Evs) [152].
Various groups have demonstrated the applicability of the GE11 peptide for the specific targeting toward the EGFR receptor for different purposes [153,154], including drug delivery [155,156,157,158]. Importantly, Ohno et al., (2013) showed that the delivery of micro RNA (miRNA) to EGFR-expressing breast cancer cells can be achieved efficiently by EVs. For this, the donor cells were engineered to express the transmembrane domain of the platelet-derived growth factor receptor fused to the GE11 peptide. Notably, the exosome that was injected intravenously could deliver the let-7a miRNA to EGFR-expressing xenograft breast cancer tissue in RAG2(−/−) mice. The result showed that EVs can be employed to target the EGFR expressing cancer tissue with nucleic acid drug for therapeutic purposes [159]. In another research study, Nakase et al. developed a novel drug delivery system based on biofunctional peptide-modified exosomes, which includes arginine-rich cell-penetrating peptide-modified exosomes for the active induction of micropinocytosis and the effective intracellular delivery of therapeutic molecules, a pH-sensitive fusogenic peptide for enhanced cytosolic release of exosomal contents, and a receptor target system using an artificial coiled-coil peptide modified on exosomal membranes [160].

6. Conclusions and Future Remark

The complex TME of cancer exhibits various barriers including hypoxia, MPS, occurrence of extravasation, cellular barriers, and drug efflux transporters, which are required to be overcome by the NDS. Despite massive progress in the advancements of developing NDS, the issues of specific targeting and toxicity remain paramount. Among various approaches, the peptide-based functionalization of NDS has been extensively studied, which showed significant advantages including augmentation of the ability of NDS to target specific receptors or mutant proteins on the surface of cancer cells. To accomplish the conjugation of peptide and NDS, different techniques have been employed, such as chemical conjugation, ligand exchange, and chemical reduction. A plethora of receptors have been reported to be associated with the malignant progression of cancer such as SSTR, integrin, transferrin, HER2, APN, LHRH, EGFR, EpCAM, and CD133, which have been utilized for developing the peptide-based functionalized NDS for targeted cancer therapy as well as diagnosis. Moreover, the peptide-based functionalization with EVs paved the way for the conjugation of biological NDS with receptor-targeted peptides for cancer therapy and diagnosis. This reveals the pertinency of peptide–NDS conjugates in future cancer treatment and diagnosis. Notably, the full potential of this method is not utilized yet, and next-generation advanced CPNDS can be developed by integrating methods of artificial intelligence for potential peptide screening to be used for cancer theranostics. In the future, there is also a need for developing smart and environment-friendly peptide-conjugated NPs, which could be potentially achieved via the integration of artificial intelligence-based machine learning algorithm for peptide screening, and the green synthesis method-based production of NPs, which is applicable for cancer diagnosis and therapeutics.

Author Contributions

Conceptualization, I.G. and Z.Y.; resources, I.G., X.W., A.T., A.I., S.T., X.C., G.K., M.L. and Z.Y.; writing—original draft preparation, I.G.; writing—review and editing, I.G., A.T., A.I., S.T. and Z.Y.; visualization, I.G., A.T., A.I., Z.Y. and M.L.; supervision, Z.Y.; project administration, I.G. and Z.Y.; funding acquisition, Z.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Innovation and Technology Fund (ITF) of the Hong Kong Government, grant number ITS/348/18FX, and the special development funding of Hong Kong Baptist University, grant number SDF18–0319-P0.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

APNAminopeptidase N
ATN-161Ac-PHSCN-NH2peptide
CPNDSConjugation of peptide and NDS
EGFREpidermal growth factor receptor
EpCAMEpithelial cell adhesion molecule
HER2Human epidermal growth factor receptor 2
LHRHLuteinizing hormone-releasing hormone
MPSMononuclear phagocytic system
NDSNano delivery system
NGRAsparagine–glycine–arginine peptide
NPNanoparticles
QDQuantum Dot
RGDArginine–glycine–aspartic acid peptide
SSTRSomatostatin receptor
TFRTransferrin receptor
TMETumor microenvironment
TAT/T7HAIYPRH peptide

References

  1. 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]
  2. Pucci, C.; Martinelli, C.; Ciofani, G. Innovative approaches for cancer treatment: Current perspectives and new challenges. Ecancermedicalscience 2019, 13, 961. [Google Scholar] [CrossRef]
  3. Thakur, A.; Roy, A.; Ghosh, A.; Chhabra, M.; Banerjee, S. Abiraterone acetate in the treatment of prostate cancer. Biomed. Pharmacother. 2018, 101, 211–218. [Google Scholar] [CrossRef] [PubMed]
  4. Sriraman, S.K.; Aryasomayajula, B.; Torchilin, V.P. Barriers to drug delivery in solid tumors. Tissue Barriers 2014, 2, e29528. [Google Scholar] [CrossRef] [Green Version]
  5. Zhou, Y.; Chen, X.; Cao, J.; Gao, H. Overcoming the biological barriers in the tumor microenvironment for improving drug delivery and efficacy. J. Mater. Chem. B 2020, 8, 6765–6781. [Google Scholar] [CrossRef]
  6. Chitty, J.L.; Filipe, E.C.; Lucas, M.C.; Herrmann, D.; Cox, T.R.; Timpson, P. Recent advances in understanding the complexities of metastasis. F1000Research 2018, 7, 1169. [Google Scholar] [CrossRef] [PubMed]
  7. Henke, E.; Nandigama, R.; Ergün, S. Extracellular Matrix in the Tumor Microenvironment and Its Impact on Cancer Therapy. Front. Mol. Biosci. 2020, 6, 160. [Google Scholar] [CrossRef] [Green Version]
  8. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The Next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  9. Gwangwa, M.V.; Joubert, A.M.; Visagie, M.H. Crosstalk between the Warburg effect, redox regulation and autophagy induction in tumourigenesis. Cell. Mol. Biol. Lett. 2018, 23, 20. [Google Scholar] [CrossRef] [Green Version]
  10. Liou, G.-Y.; Storz, P. Reactive oxygen species in cancer. Free Radic. Res. 2010, 44, 479–496. [Google Scholar] [CrossRef] [Green Version]
  11. Roma-Rodrigues, C.; Pombo, I.; Raposo, L.; Pedrosa, P.; Fernandes, A.; Baptista, P. Nanotheranostics Targeting the Tumor Microenvironment. Front. Bioeng. Biotechnol. 2019, 7, 197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Dirkx, A.E.M.; Egbrink, M.G.A.O.; Castermans, K.; Van Der Schaft, D.W.J.; Thijssen, V.; Dings, R.; Kwee, L.; Mayo, K.H.; Wagstaff, J.; Ter Steege, J.C.A.B.; et al. Anti-angiogenesis therapy can overcome endothelial cell anergy and promote leukocyte-endothelium interactions and infiltration in tumors. FASEB J. 2006, 20, 621–630. [Google Scholar] [CrossRef] [Green Version]
  13. Klein, D. The Tumor Vascular Endothelium as Decision Maker in Cancer Therapy. Front. Oncol. 2018, 8, 367. [Google Scholar] [CrossRef]
  14. Wang, Y.; He, Q.-Y.; Sun, R.W.-Y.; Che, C.M.; Chiu, J.-F. Gold(III) Porphyrin 1a Induced Apoptosis by Mitochondrial Death Pathways Related to Reactive Oxygen Species. Cancer Res. 2005, 65, 11553–11564. [Google Scholar] [CrossRef] [PubMed]
  15. Clichici, S.; Filip, A.; Daicoviciu, D.; Ion, R.; Mocan, T.; Tatomir, C.; Rogojan, L.; Olteanu, E.D.; Muresan, A. The dynamics of reactive oxygen species in photodynamic therapy with tetra sulfophenyl-porphyrin. Acta Physiol. Hung. 2010, 97, 41–51. [Google Scholar] [CrossRef] [PubMed]
  16. Deng, J.; Li, H.; Yang, M.; Wu, F. Palladium porphyrin complexes for photodynamic cancer therapy: Effect of porphyrin units and metal. Photochem. Photobiol. Sci. 2020, 19, 905–912. [Google Scholar] [CrossRef]
  17. Nel, A.E.; Mädler, L.; Velegol, D.; Xia, T.; Hoek, E.M.V.; Somasundaran, P.; Klaessig, F.; Castranova, V.; Thompson, M. Understanding biophysicochemical interactions at the nano–bio interface. Nat. Mater. 2009, 8, 543–557. [Google Scholar] [CrossRef] [PubMed]
  18. Chow, A.; Brown, B.D.; Merad, M. Studying the mononuclear phagocyte system in the molecular age. Nat. Rev. Immunol. 2011, 11, 788–798. [Google Scholar] [CrossRef]
  19. Rehm, M.; Zahler, S.; Lötsch, M.; Welsch, U.; Conzen, P.; Jacob, M.; Becker, B.F. Endothelial Glycocalyx as an Additional Barrier Determining Extravasation of 6% Hydroxyethyl Starch or 5% Albumin Solutions in the Coronary Vascular Bed. Anesthesiology 2004, 100, 1211–1223. [Google Scholar] [CrossRef] [Green Version]
  20. Dull, R.O.; Dinavahi, R.; Schwartz, L.; Humphries, D.E.; Berry, D.; Sasisekharan, R.; Garcia, J.G.N. Lung endothelial heparan sulfates mediate cationic peptide-induced barrier dysfunction: A new role for the glycocalyx. Am. J. Physiol. Cell. Mol. Physiol. 2003, 285, L986–L995. [Google Scholar] [CrossRef] [Green Version]
  21. Kutuzov, N.; Flyvbjerg, H.; Lauritzen, M. Contributions of the glycocalyx, endothelium, and extravascular compartment to the blood–brain barrier. Proc. Natl. Acad. Sci. USA 2018, 115, E9429–E9438. [Google Scholar] [CrossRef] [Green Version]
  22. Barua, S.; Mitragotri, S. Challenges associated with penetration of nanoparticles across cell and tissue barriers: A review of current status and future prospects. Nano Today 2014, 9, 223–243. [Google Scholar] [CrossRef] [PubMed]
  23. Jia, J.; Wang, Z.; Yue, T.; Su, G.; Teng, C.; Yan, B. Crossing Biological Barriers by Engineered Nanoparticles. Chem. Res. Toxicol. 2020, 33, 1055–1060. [Google Scholar] [CrossRef] [PubMed]
  24. Firer, M.A.; Gellerman, G. Targeted drug delivery for cancer therapy: The other side of antibodies. J. Hematol. Oncol. 2012, 5, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Jeong, W.-J.; Bu, J.; Kubiatowicz, L.J.; Chen, S.S.; Kim, Y.; Hong, S. Peptide–nanoparticle conjugates: A next generation of diagnostic and therapeutic platforms? Nano Converg. 2018, 5, 1–18. [Google Scholar] [CrossRef] [PubMed]
  26. Baptista, P.; Kanaras, A.; Heuer-Jungemann, A.; Roma-Rodrigues, C.; Fernandes, A. Peptide-coated gold nanoparticles for modulation of angiogenesis in vivo. Int. J. Nanomed. 2016, 11, 2633–2639. [Google Scholar] [CrossRef] [Green Version]
  27. Liu, L.; Li, X.; Zhang, H.; Chen, H.; Abualrejal, M.M.; Song, D.; Wang, Z. Six-in-one peptide functionalized upconversion@ polydopamine nanoparticle-based ratiometric fluorescence sensing platform for real-time evaluating anticancer efficacy through monitoring caspase-3 activity. Sens. Actuators B Chem. 2021, 333, 129554. [Google Scholar] [CrossRef]
  28. Li, X.; Liu, L.; Fu, Y.; Chen, H.; Abualrejal, M.M.; Zhang, H.; Wang, Z.; Zhang, H. Peptide-enhanced tumor accumulation of upconversion nanoparticles for sensitive upconversion luminescence/magnetic resonance dual-mode bioimaging of colorectal tumors. Acta Biomater. 2020, 104, 167–175. [Google Scholar] [CrossRef]
  29. Bartczak, D.; Kanaras, A.G. Preparation of Peptide-Functionalized Gold Nanoparticles Using One Pot EDC/Sulfo-NHS Coupling. Langmuir 2011, 27, 10119–10123. [Google Scholar] [CrossRef]
  30. Fu, Y.; Li, X.; Chen, H.; Wang, Z.; Yang, W.; Zhang, H. CXC Chemokine Receptor 4 Antagonist Functionalized Renal Clearable Manganese-Doped Iron Oxide Nanoparticles for Active-Tumor-Targeting Magnetic Resonance Imaging-Guided Bio-Photothermal Therapy. ACS Appl. Bio Mater. 2019, 2, 3613–3621. [Google Scholar] [CrossRef]
  31. Wang, W.; Anderson, C.F.; Wang, Z.; Wu, W.; Cui, H.; Liu, C.-J. Peptide-templated noble metal catalysts: Syntheses and applications. Chem. Sci. 2017, 8, 3310–3324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Li, Y.; Tang, Z.; Prasad, P.N.; Knecht, M.R.; Swihart, M.T. Peptide-mediated synthesis of gold nanoparticles: Effects of peptide sequence and nature of binding on physicochemical properties. Nanoscale 2014, 6, 3165–3172. [Google Scholar] [CrossRef]
  33. Krpetic, Z.; Nativo, P.; Porta, F.; Brust, M. A Multidentate Peptide for Stabilization and Facile Bioconjugation of Gold Nanoparticles. Bioconjug. Chem. 2009, 20, 619–624. [Google Scholar] [CrossRef]
  34. Fernandes, R.; Smyth, N.R.; Muskens, O.; Nitti, S.; Heuer-Jungemann, A.; Ardern-Jones, M.R.; Kanaras, A.G. Interactions of Skin with Gold Nanoparticles of Different Surface Charge, Shape, and Functionality. Small 2014, 11, 713–721. [Google Scholar] [CrossRef] [PubMed]
  35. Samieegohar, M.; Sha, F.; Clayborne, A.Z.; Wei, T. ReaxFF MD Simulations of Peptide-Grafted Gold Nanoparticles. Langmuir 2019, 35, 5029–5036. [Google Scholar] [CrossRef] [PubMed]
  36. Luo, J.; Cheng, Y.; Gong, Z.-W.; Wu, K.; Zhou, Y.; Chen, H.-X.; Gauthier, M.; Cheng, Y.-Z.; Liang, J.; Zou, T. Self-Assembled Peptide Functionalized Gold Nanopolyhedrons with Excellent Chiral Optical Properties. Langmuir 2019, 36, 600–608. [Google Scholar] [CrossRef]
  37. Lévy, R.; Thanh, N.T.K.; Doty, R.C.; Hussain, I.; Nichols, R.; Schiffrin, D.J.; Brust, M.; Fernig, D. Rational and Combinatorial Design of Peptide Capping Ligands for Gold Nanoparticles. J. Am. Chem. Soc. 2004, 126, 10076–10084. [Google Scholar] [CrossRef]
  38. Hu, B.; Kong, F.; Gao, X.; Jiang, L.; Li, X.; Gao, W.; Xu, K.; Tang, B. Avoiding Thiol Compound Interference: A Nanoplatform Based on High-Fidelity Au-Se Bonds for Biological Applications. Angew. Chem. Int. Ed. 2018, 57, 5306–5309. [Google Scholar] [CrossRef] [PubMed]
  39. Pan, W.; Liu, X.; Wan, X.; Li, J.; Li, Y.; Lu, F.; Li, N.; Tang, B. Rapid Preparation of Au–Se–Peptide Nanoprobe Based on a Freezing Method for Bioimaging. Anal. Chem. 2019, 91, 15982–15987. [Google Scholar] [CrossRef]
  40. Luan, M.; Shi, M.; Pan, W.; Li, N.; Tang, B. A gold–selenium-bonded nanoprobe for real-timein situimaging of the upstream and downstream relationship between uPA and MMP-9 in cancer cells. Chem. Commun. 2019, 55, 5817–5820. [Google Scholar] [CrossRef]
  41. Si, S.; Mandal, T.K. Tryptophan-Based Peptides to Synthesize Gold and Silver Nanoparticles: A Mechanistic and Kinetic Study. Chem. A Eur. J. 2007, 13, 3160–3168. [Google Scholar] [CrossRef]
  42. Upert, G.; Bouillère, F.; Wennemers, H. Oligoprolines as Scaffolds for the Formation of Silver Nanoparticles in Defined Sizes: Correlating Molecular and Nanoscopic Dimensions. Angew. Chem. Int. Ed. 2011, 51, 4231–4234. [Google Scholar] [CrossRef]
  43. Corra, S.; Lewandowska, U.; Benetti, E.M.; Wennemers, H. Size-Controlled Formation of Noble-Metal Nanoparticles in Aqueous Solution with a Thiol-Free Tripeptide. Angew. Chem. Int. Ed. 2016, 55, 8542–8545. [Google Scholar] [CrossRef]
  44. Graf, P.; Mantion, A.; Foelske, A.; Shkilnyy, A.; Masic, A.; Thünemann, A.F.; Taubert, A. Peptide-Coated Silver Nanoparticles: Synthesis, Surface Chemistry, and pH-Triggered, Reversible Assembly into Particle Assemblies. Chem. A Eur. J. 2009, 15, 5831–5844. [Google Scholar] [CrossRef] [PubMed]
  45. Papst, S.; Brimble, M.A.; Tilley, R.D.; Williams, D.E. One-Pot Synthesis of Functionalized Noble Metal Nanoparticles Using a Rationally Designed Phosphopeptide. Part. Part. Syst. Charact. 2014, 31, 971–975. [Google Scholar] [CrossRef]
  46. Henninot, A.; Collins, J.C.; Nuss, J.M. The Current State of Peptide Drug Discovery: Back to the Future? J. Med. Chem. 2017, 61, 1382–1414. [Google Scholar] [CrossRef]
  47. Lau, J.L.; Dunn, M.K. Therapeutic peptides: Historical perspectives, current development trends, and future directions. Bioorg. Med. Chem. 2018, 26, 2700–2707. [Google Scholar] [CrossRef] [PubMed]
  48. Weinstock, M.T.; Francis, J.N.; Redman, J.S.; Kay, M.S. Protease-resistant peptide design-empowering nature’s fragile warriors against HIV. Biopolymers 2012, 98, 431–442. [Google Scholar] [CrossRef] [PubMed]
  49. Talmadge, J.E. Pharmacodynamic aspects of peptide administration biological response modifiers. Adv. Drug Deliv. Rev. 1998, 33, 241–252. [Google Scholar] [CrossRef]
  50. Shu, J.Y.; Panganiban, B.; Xu, T. Peptide-Polymer Conjugates: From Fundamental Science to Application. Annu. Rev. Phys. Chem. 2013, 64, 631–657. [Google Scholar] [CrossRef]
  51. Xiao, Y.; Reis, L.A.; Feric, N.; Knee, E.J.; Gu, J.; Cao, S.; Laschinger, C.; Londoño, C.; Antolovich, J.; McGuigan, A.P.; et al. Diabetic wound regeneration using peptide-modified hydrogels to target re-epithelialization. Proc. Natl. Acad. Sci. USA 2016, 113, E5792–E5801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Du, J.-Z.; Lane, L.A.; Nie, S. Stimuli-responsive nanoparticles for targeting the tumor microenvironment. J. Control. Release 2015, 219, 205–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Tang, Y.; Wang, X.; Li, J.; Nie, Y.; Liao, G.; Yu, Y.; Li, C. Overcoming the Reticuloendothelial System Barrier to Drug Delivery with a “Don’t-Eat-Us” Strategy. ACS Nano 2019, 13, 13015–13026. [Google Scholar] [CrossRef] [PubMed]
  54. Georgieva, J.V.; Brinkhuis, R.P.; Stojanov, K.; Weijers, C.A.G.M.; Zuilhof, H.; Rutjes, F.P.J.T.; Hoekstra, D.; Van Hest, J.C.M.; Zuhorn, I.S. Peptide-Mediated Blood-Brain Barrier Transport of Polymersomes. Angew. Chem. Int. Ed. 2012, 51, 8339–8342. [Google Scholar] [CrossRef]
  55. Yao, H.; Wang, K.; Wang, Y.; Wang, S.; Li, J.; Lou, J.; Ye, L.; Yan, X.; Lu, W.; Huang, R. Enhanced blood–brain barrier penetration and glioma therapy mediated by a new peptide modified gene delivery system. Biomaterials 2015, 37, 345–352. [Google Scholar] [CrossRef]
  56. Ma, P.; Mumper, R.J. Anthracycline nano-delivery systems to overcome multiple drug resistance: A comprehensive review. Nano Today 2013, 8, 313–331. [Google Scholar] [CrossRef] [Green Version]
  57. Mazel, M.; Clair, P.; Rousselle, C.; Vidal, P.; Scherrmann, J.-M.; Mathieu, D.; Temsamani, J. Doxorubicin-peptide conjugates overcome multidrug resistance. Anti-Cancer Drugs 2001, 12, 107–116. [Google Scholar] [CrossRef]
  58. Guillemard, V.; Saragovi, H.U. Prodrug chemotherapeutics bypass p-glycoprotein resistance and kill tumors in vivo with high efficacy and target-dependent selectivity. Oncogene 2004, 23, 3613–3621. [Google Scholar] [CrossRef] [Green Version]
  59. Modi, D.A.; Sunoqrot, S.; Bugno, J.; Lantvit, D.D.; Hong, S.; Burdette, J.E. Targeting of follicle stimulating hormone peptide-conjugated dendrimers to ovarian cancer cells. Nanoscale 2014, 6, 2812–2820. [Google Scholar] [CrossRef]
  60. Jiang, X.; Bugno, J.; Hu, C.; Yang, Y.; Herold, T.; Qi, J.; Chen, P.; Gurbuxani, S.; Arnovitz, S.; Ulrich, B.; et al. Targeted Treatment of FLT3-Overexpressing Acute Myeloid Leukemia with MiR-150 Nanoparticles Guided By Conjugated FLT3 Ligand Peptides. Blood 2015, 126, 3784. [Google Scholar] [CrossRef]
  61. Chang, C.-C.; Liu, D.-Z.; Lin, S.-Y.; Liang, H.-J.; Hou, W.-C.; Huang, W.-J.; Chang, C.-H.; Ho, F.-M.; Liang, Y.-C. Liposome encapsulation reduces cantharidin toxicity. Food Chem. Toxicol. 2008, 46, 3116–3121. [Google Scholar] [CrossRef] [PubMed]
  62. Dai, W.; Yang, T.; Wang, X.; Wang, J.; Zhang, X.; Zhang, Q. PHSCNK-Modified and doxorubicin-loaded liposomes as a dual targeting system to integrin-overexpressing tumor neovasculature and tumor cells. J. Drug Target. 2009, 18, 254–263. [Google Scholar] [CrossRef] [PubMed]
  63. Liu, X.-Y.; Ruan, L.-M.; Mao, W.-W.; Wang, J.-Q.; Shen, Y.-Q.; Sui, M.-H. Preparation of RGD-modified Long Circulating Liposome Loading Matrine, and its in vitro Anti-cancer Effects. Int. J. Med. Sci. 2010, 7, 197–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Bandekar, A.; Zhu, C.; Gomez, A.; Menzenski, M.Z.; Sempkowski, M.; Sofou, S. Masking and Triggered Unmasking of Targeting Ligands on Liposomal Chemotherapy Selectively Suppress Tumor Growth in Vivo. Mol. Pharm. 2012, 10, 152–160. [Google Scholar] [CrossRef]
  65. Zahmatkeshan, M.; Gheybi, F.; Rezayat, S.M.; Jaafari, M.R. Improved drug delivery and therapeutic efficacy of PEgylated liposomal doxorubicin by targeting anti-HER2 peptide in murine breast tumor model. Eur. J. Pharm. Sci. 2016, 86, 125–135. [Google Scholar] [CrossRef]
  66. Chen, Y.; Wu, J.J.; Huang, L. Nanoparticles Targeted With NGR Motif Deliver c-myc siRNA and Doxorubicin for Anticancer Therapy. Mol. Ther. 2010, 18, 828–834. [Google Scholar] [CrossRef]
  67. Hou, J.; Diao, Y.; Li, W.; Yang, Z.; Zhang, L.; Chen, Z.; Wu, Y. RGD peptide conjugation results in enhanced antitumor activity of PD0325901 against glioblastoma by both tumor-targeting delivery and combination therapy. Int. J. Pharm. 2016, 505, 329–340. [Google Scholar] [CrossRef]
  68. Cai, W.; Chen, K.; Li, Z.-B.; Gambhir, S.S.; Chen, X. Dual-Function Probe for PET and Near-Infrared Fluorescence Imaging of Tumor Vasculature. J. Nucl. Med. 2007, 48, 1862–1870. [Google Scholar] [CrossRef]
  69. Qin, Y.; Chen, H.; Zhang, Q.; Wang, X.; Yuan, W.; Kuai, R.; Tang, J.; Zhang, L.; Zhang, Z.; Zhang, Q.; et al. Liposome formulated with TAT-modified cholesterol for improving brain delivery and therapeutic efficacy on brain glioma in animals. Int. J. Pharm. 2011, 420, 304–312. [Google Scholar] [CrossRef]
  70. Huang, N.; Cheng, S.; Zhang, X.; Tian, Q.; Pi, J.; Tang, J.; Wang, F.; Chen, J.; Xie, Z.; Xu, Z.; et al. Efficacy of NGR peptide-modified PEGylated quantum dots for crossing the blood–brain barrier and targeted fluorescence imaging of glioma and tumor vasculature. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 83–93. [Google Scholar] [CrossRef]
  71. Zhang, J.; Jin, W.; Wang, X.; Wang, J.; Zhang, X.; Zhang, Q. A Novel Octreotide Modified Lipid Vesicle Improved the Anticancer Efficacy of Doxorubicin in Somatostatin Receptor 2 Positive Tumor Models. Mol. Pharm. 2010, 7, 1159–1168. [Google Scholar] [CrossRef]
  72. Wang, R.-H.; Cao, H.-M.; Tian, Z.-J.; Jin, B.; Wang, Q.; Ma, H.; Wu, J. Efficacy of dual-functional liposomes containing paclitaxel for treatment of lung cancer. Oncol. Rep. 2017, 38, 3285. [Google Scholar] [CrossRef] [Green Version]
  73. Saad, M.; Garbuzenko, O.B.; Ber, E.; Chandna, P.; Khandare, J.J.; Pozharov, V.P.; Minko, T. Receptor targeted polymers, dendrimers, liposomes: Which nanocarrier is the most efficient for tumor-specific treatment and imaging? J. Control. Release 2008, 130, 107–114. [Google Scholar] [CrossRef] [Green Version]
  74. Xiong, X.-B.; Huang, Y.; Lu, W.-L.; Zhang, X.; Zhang, H.; Nagai, T.; Zhang, Q. Intracellular delivery of doxorubicin with RGD-modified sterically stabilized liposomes for an improved antitumor efficacy: In vitro and in vivo. J. Pharm. Sci. 2005, 94, 1782–1793. [Google Scholar] [CrossRef]
  75. Wu, H.; Yao, L.; Mei, J.; Li, F. Development of synthetic of peptide-functionalized liposome for enhanced targeted ovarian carcinoma therapy. Int. J. Clin. Exp. Pathol. 2015, 8, 207–216. [Google Scholar]
  76. Du, Y.-Z.; Jie, L.-Y.; Cai, L.-L.; Wang, L.-J.; Ying, X.-Y.; Yu, R.-S.; Zhang, M. Actively-targeted LTVSPWY peptide-modified magnetic nanoparticles for tumor imaging. Int. J. Nanomed. 2012, 7, 3981–3989. [Google Scholar] [CrossRef] [Green Version]
  77. He, Y.; Yuan, X.-M.; Lei, P.; Wu, S.; Xing, W.; Lan, X.-L.; Zhu, H.-F.; Huang, T.; Wang, G.-B.; An, R.; et al. The antiproliferative effects of somatostatin receptor subtype 2 in breast cancer cells. Acta Pharmacol. Sin. 2009, 30, 1053–1059. [Google Scholar] [CrossRef] [Green Version]
  78. Kwekkeboom, D.J.; Krenning, E.P. Somatostatin receptor imaging. Semin. Nucl. Med. 2002, 32, 84–91. [Google Scholar] [CrossRef] [PubMed]
  79. Susini, C.; Buscail, L. Rationale for the use of somatostatin analogs as antitumor agents. Ann. Oncol. 2006, 17, 1733–1742. [Google Scholar] [CrossRef]
  80. Patel, Y.C. Somatostatin and Its Receptor Family. Front. Neuroendocr. 1999, 20, 157–198. [Google Scholar] [CrossRef] [PubMed]
  81. Florio, T.; Schettini, G. Multiple intracellular effectors modulate physiological functions of the cloned somatostatin receptors. J. Mol. Endocrinol. 1996, 17, 89–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Modlin, I.M.; Moss, S.F.; Oberg, K.; Padbury, R.; Hicks, R.; Gustafsson, B.I.; Wright, N.A.; Kidd, M. Gastrointestinal neuroendocrine (carcinoid) tumours: Current diagnosis and management. Med. J. Aust. 2010, 193, 46–52. [Google Scholar] [CrossRef]
  83. Kwekkeboom, D.J.; Kam, B.L.; van Essen, M.; Teunissen, J.J.M.; van Eijck, C.H.J.; Valkema, R.; de Jong, M.; de Herder, W.W.; Krenning, E.P. Somatostatin receptor-based imaging and therapy of gastroenteropancreatic neuroendocrine tumors. Endocr.-Relat. Cancer 2010, 17, R53–R73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Abou, D.S.; Thorek, D.L.J.; Ramos, N.N.; Pinkse, M.W.H.; Wolterbeek, H.T.; Carlin, S.; Beattie, B.J.; Lewis, J.S. 89Zr-Labeled Paramagnetic Octreotide-Liposomes for PET-MR Imaging of Cancer. Pharm. Res. 2012, 30, 878–888. [Google Scholar] [CrossRef]
  85. Su, C.-Y.; Li, J.-Q.; Zhang, L.-L.; Wang, H.; Wang, F.-H.; Tao, Y.-W.; Wang, Y.-Q.; Guo, Q.-R.; Li, J.-J.; Liu, Y.; et al. The Biological Functions and Clinical Applications of Integrins in Cancers. Front. Pharmacol. 2020, 11. [Google Scholar] [CrossRef]
  86. Serini, G.; Valdembri, D.; Bussolino, F. Integrins and angiogenesis: A sticky business. Exp. Cell Res. 2006, 312, 651–658. [Google Scholar] [CrossRef]
  87. Zhao, J.; Santino, F.; Giacomini, D.; Gentilucci, L. Integrin-Targeting Peptides for the Design of Functional Cell-Responsive Biomaterials. Biomedicines 2020, 8, 307. [Google Scholar] [CrossRef]
  88. Danhier, F.; Le Breton, A.; Préat, V. RGD-Based Strategies To Target Alpha(v) Beta(3) Integrin in Cancer Therapy and Diagnosis. Mol. Pharm. 2012, 9, 2961–2973. [Google Scholar] [CrossRef]
  89. Temming, K.; Schiffelers, R.; Molema, G.; Kok, R.J. RGD-based strategies for selective delivery of therapeutics and imaging agents to the tumour vasculature. Drug Resist. Updat. 2005, 8, 381–402. [Google Scholar] [CrossRef] [PubMed]
  90. Casagrande, C.; Manzoni, L. Integrin-Mediated Drug Delivery in Cancer and Cardiovascular Diseases with Peptide-Functionalized Nanoparticles. Curr. Med. Chem. 2012, 19, 3128–3151. [Google Scholar] [CrossRef]
  91. Chen, K.; Chen, X. Integrin Targeted Delivery of Chemotherapeutics. Theranostics 2011, 1, 189–200. [Google Scholar] [CrossRef] [Green Version]
  92. Hölig, P.; Bach, M.; Völkel, T.; Nahde, T.; Hoffmann, S.; Müller, R.; Kontermann, R.E. Novel RGD lipopeptides for the targeting of liposomes to integrin-expressing endothelial and melanoma cells. Protein Eng. Des. Sel. 2004, 17, 433–441. [Google Scholar] [CrossRef] [Green Version]
  93. Bogdanowich-Knipp, S.J.; Chakrabarti, S.; Siahaan, T.J.; Williams, T.D.; Dillman, R.K. Solution stability of linear vs. cyclic RGD peptides. J. Pept. Res. 1999, 53, 530–541. [Google Scholar] [CrossRef] [PubMed]
  94. Haubner, R.; Weber, W.A.; Beer, A.J.; Vabuliene, E.; Reim, D.; Sarbia, M.; Becker, K.-F.; Goebel, M.; Hein, R.; Wester, H.-J.; et al. Noninvasive Visualization of the Activated αvβ3 Integrin in Cancer Patients by Positron Emission Tomography and [18F]Galacto-RGD. PLoS Med. 2005, 2, e70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Wei, Y.-C.; Gao, Y.; Zhang, J.; Fu, Z.; Zheng, J.; Liu, N.; Hu, X.; Hou, W.; Yu, J.; Yuan, S. Stereotactic Comparison Study of 18F-Alfatide and 18F-FDG PET Imaging in an LLC Tumor-Bearing C57BL/6 Mouse Model. Sci. Rep. 2016, 6, 28757. [Google Scholar] [CrossRef] [Green Version]
  96. Ren, J.; Xu, S.; Guo, D.; Zhang, J.; Liu, S. Increased expression of α5β1-integrin is a prognostic marker for patients with gastric cancer. Clin. Transl. Oncol. 2013, 16, 668–674. [Google Scholar] [CrossRef] [PubMed]
  97. Neckers, L.M.; Trepel, J.B. Transferrin Receptor Expression and the Control of Cell Growth. Cancer Investig. 1986, 4, 461–470. [Google Scholar] [CrossRef]
  98. Daniels, T.R.; Bernabeu, E.; Rodríguez, J.A.; Patel, S.; Kozman, M.; Chiappetta, D.A.; Holler, E.; Ljubimova, J.Y.; Helguera, G.; Penichet, M.L. The transferrin receptor and the targeted delivery of therapeutic agents against cancer. Biochim. Biophys. Acta (BBA) Gen. Subj. 2012, 1820, 291–317. [Google Scholar] [CrossRef] [Green Version]
  99. Kondo, K.; Noguchi, M.; Mukai, K.; Matsuno, Y.; Sato, Y.; Shimosato, Y.; Monden, Y. Transferrin Receptor Expression in Adenocarcinoma of the Lung as a Histopathologic Indicator of Prognosis. Chest 1990, 97, 1367–1371. [Google Scholar] [CrossRef] [Green Version]
  100. Habeshaw, J.; Lister, T.; Stansfeld, A.; Greaves, M. Correlation of transferrin receptor expression with histological class and outcome in non-hodgkin lymphoma. Lancet 1983, 321, 498–501. [Google Scholar] [CrossRef]
  101. Prior, R.; Reifenberger, G.; Wechsler, W. Transferrin receptor expression in tumours of the human nervous system: Relation to tumour type, grading and tumour growth fraction. Virchows Arch. A 1990, 416, 491–496. [Google Scholar] [CrossRef]
  102. Seymour, G.J.; Walsh, M.D.; Lavin, M.; Strutton, G.; Gardiner, R.A. Transferrin receptor expression by human bladder transitional cell carcinomas. Urol. Res. 1987, 15, 341–344. [Google Scholar] [CrossRef]
  103. Rousselle, C.; Clair, P.; Lefauconnier, J.-M.; Kaczorek, M.; Scherrmann, J.-M.; Temsamani, J. New Advances in the Transport of Doxorubicin through the Blood-Brain Barrier by a Peptide Vector-Mediated Strategy. Mol. Pharmacol. 2000, 57, 679–686. [Google Scholar] [CrossRef] [Green Version]
  104. Gao, J.-Q.; Lv, Q.; Li, L.-M.; Tang, X.-J.; Li, F.-Z.; Hu, Y.-L.; Han, M. Glioma targeting and blood–brain barrier penetration by dual-targeting doxorubincin liposomes. Biomaterials 2013, 34, 5628–5639. [Google Scholar] [CrossRef]
  105. Lee, J.H.; Engler, J.A.; Collawn, J.F.; Moore, B.A. Receptor mediated uptake of peptides that bind the human transferrin receptor. JBIC J. Biol. Inorg. Chem. 2001, 268, 2004–2012. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, R.; Feng, G.; Zhang, C.-J.; Cai, X.; Cheng, X.; Liu, B. Real-Time Specific Light-Up Sensing of Transferrin Receptor: Image-Guided Photodynamic Ablation of Cancer Cells through Controlled Cytomembrane Disintegration. Anal. Chem. 2016, 88, 4841–4848. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, K.; Zhang, Y.; Wang, J.; Yuan, A.; Sun, M.; Wu, J.; Hu, Y. Self-assembled IR780-loaded transferrin nanoparticles as an imaging, targeting and PDT/PTT agent for cancer therapy. Sci. Rep. 2016, 6, 27421. [Google Scholar] [CrossRef] [PubMed]
  108. Huang, R.-Q.; Ke, W.-L.; Qu, Y.-H.; Zhu, J.-H.; Pei, Y.-Y.; Jiang, C. Characterization of lactoferrin receptor in brain endothelial capillary cells and mouse brain. J. Biomed. Sci. 2006, 14, 121–128. [Google Scholar] [CrossRef] [PubMed]
  109. Lalani, J.; Raichandani, Y.; Mathur, R.; Lalan, M.; Chutani, K.; Mishra, A.K.; Misra, A. Comparative Receptor Based Brain Delivery of Tramadol-Loaded Poly(lactic-co-glycolic acid) Nanoparticles. J. Biomed. Nanotechnol. 2012, 8, 918–927. [Google Scholar] [CrossRef]
  110. Lalani, J.; Rathi, M.; Lalan, M.; Misra, A. Protein functionalized tramadol-loaded PLGA nanoparticles: Preparation, optimization, stability and pharmacodynamic studies. Drug Dev. Ind. Pharm. 2012, 39, 854–864. [Google Scholar] [CrossRef]
  111. Miao, D.; Jiang, M.; Liu, Z.; Gu, G.; Hu, Q.; Kang, T.; Song, Q.; Yao, L.; Li, W.; Gao, X.; et al. Co-administration of Dual-Targeting Nanoparticles with Penetration Enhancement Peptide for Antiglioblastoma Therapy. Mol. Pharm. 2013, 11, 90–101. [Google Scholar] [CrossRef]
  112. Tai, W.; Mahato, R.; Cheng, K. The role of HER2 in cancer therapy and targeted drug delivery. J. Control. Release 2010, 146, 264–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Yarden, Y. Biology of HER2 and Its Importance in Breast Cancer. Oncology 2001, 61, 1–13. [Google Scholar] [CrossRef] [PubMed]
  114. Zhu, Z.-L.; Zhang, J.; Chen, M.-L.; Li, K. Efficacy and Safety of Trastuzumab Added to Standard Treatments for HER2-positive Metastatic Breast Cancer Patients. Asian Pac. J. Cancer Prev. 2013, 14, 7111–7116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Bang, Y.-J.; Van Cutsem, E.; Feyereislova, A.; Chung, H.; Shen, L.; Sawaki, A.; Lordick, F.; Ohtsu, A.; Omuro, Y.; Satoh, T.; et al. Trastuzumab in combination with chemotherapy versus chemotherapy alone for treatment of HER2-positive advanced gastric or gastro-oesophageal junction cancer (ToGA): A phase 3, open-label, randomised controlled trial. Lancet 2010, 376, 687–697. [Google Scholar] [CrossRef]
  116. Karasseva, N.G.; Glinsky, V.; Chen, N.X.; Komatireddy, R.; Quinn, T.P. Identification and characterization of peptides that bind human ErbB-2 selected from a bacteriophage display library. Protein J. 2002, 21, 287–296. [Google Scholar] [CrossRef] [PubMed]
  117. Park, B.-W.; Zhang, H.-T.; Wu, C.; Berezov, A.; Zhang, X.; Dua, R.; Wang, Q.; Kao, G.; O’Rourke, D.; Greene, M.I.; et al. Rationally designed anti-HER2/neu peptide mimetic disables P185HER2/neu tyrosine kinases in vitro and in vivo. Nat. Biotechnol. 2000, 18, 194–198. [Google Scholar] [CrossRef]
  118. Lei, H.; Cao, P.; Miao, G.; Lin, Z.; Diao, Z. Expression and Functional Characterization of Tumor-Targeted Fusion Protein Composed of NGR Peptide and 15-kDa Actin Fragment. Appl. Biochem. Biotechnol. 2010, 162, 988–995. [Google Scholar] [CrossRef]
  119. Dharap, S.S.; Wang, Y.; Chandna, P.; Khandare, J.J.; Qiu, B.; Gunaseelan, S.; Sinko, P.; Stein, S.; Farmanfarmaian, A.; Minko, T. Tumor-specific targeting of an anticancer drug delivery system by LHRH peptide. Proc. Natl. Acad. Sci. USA 2005, 102, 12962–12967. [Google Scholar] [CrossRef] [Green Version]
  120. Khandare, J.J.; Chandna, P.; Wang, Y.; Pozharov, V.P.; Minko, T. Novel Polymeric Prodrug with Multivalent Components for Cancer Therapy. J. Pharmacol. Exp. Ther. 2006, 317, 929–937. [Google Scholar] [CrossRef] [Green Version]
  121. Bajusz, S.; Janaky, T.; Csernus, V.J.; Bokser, L.; Fekete, M.; Srkalovic, G.; Redding, T.W.; Schally, A. Highly potent metallopeptide analogues of luteinizing hormone-releasing hormone. Proc. Natl. Acad. Sci. USA 1989, 86, 6313–6317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Bajusz, S.; Janaky, T.; Csernus, V.J.; Bokser, L.; Fekete, M.; Srkalovic, G.; Redding, T.W.; Schally, A. Highly potent analogues of luteinizing hormone-releasing hormone containing D-phenylalanine nitrogen mustard in position 6. Proc. Natl. Acad. Sci. USA 1989, 86, 6318–6322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Fodor, K.; Dobos, N.; Schally, A.; Steiber, Z.; Olah, G.; Sipos, E.; Szekvolgyi, L.; Halmos, G. The targeted LHRH analog AEZS-108 alters expression of genes related to angiogenesis and development of metastasis in uveal melanoma. Oncotarget 2020, 11, 175–187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Liu, S.V.; Tsao-Wei, D.D.; Xiong, S.; Groshen, S.; Dorff, T.B.; Quinn, D.; Tai, Y.-C.; Engel, J.; Hawes, D.; Schally, A.; et al. Phase I, Dose-Escalation Study of the Targeted Cytotoxic LHRH Analog AEZS-108 in Patients with Castration- and Taxane-Resistant Prostate Cancer. Clin. Cancer Res. 2014, 20, 6277–6283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Li, M.; Tang, Z.; Zhang, Y.; Lv, S.; Li, Q.; Chen, X. Targeted delivery of cisplatin by LHRH-peptide conjugated dextran nanoparticles suppresses breast cancer growth and metastasis. Acta Biomater. 2015, 18, 132–143. [Google Scholar] [CrossRef] [PubMed]
  126. Zhang, H.; Berezov, A.; Wang, Q.; Zhang, G.; Drebin, J.; Murali, R.; Greene, M.I. ErbB receptors: From oncogenes to targeted cancer therapies. J. Clin. Investig. 2007, 117, 2051–2058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Hossein-Nejad-Ariani, H.; AlThagafi, E.; Kaur, K. Small Peptide Ligands for Targeting EGFR in Triple Negative Breast Cancer Cells. Sci. Rep. 2019, 9, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Huang, X.; Chen, L.; Zhang, Y.; Zhou, S.; Cai, H.-H.; Li, T.; Jin, H.; Cai, J.; Zhou, H.; Pi, J. GE11 Peptide Conjugated Liposomes for EGFR-Targeted and Chemophotothermal Combined Anticancer Therapy. Bioinorg. Chem. Appl. 2021, 2021, 1–15. [Google Scholar] [CrossRef]
  129. Pan, W.-S.; Han, C.-Y.; Yue, L.-L.; Tai, L.-Y.; Zhou, L.; Li, X.-Y.; Xing, G.-H.; Yang, X.-G.; Sun, M.-S. A novel small peptide as an epidermal growth factor receptor targeting ligand for nanodelivery in vitro. Int. J. Nanomed. 2013, 8, 1541–1549. [Google Scholar] [CrossRef] [Green Version]
  130. Mayr, J.; Hager, S.; Koblmüller, B.; Klose, M.H.M.; Holste, K.; Fischer, B.; Pelivan, K.; Berger, W.; Heffeter, P.; Kowol, C.R.; et al. EGFR-targeting peptide-coupled platinum(IV) complexes. JBIC J. Biol. Inorg. Chem. 2017, 22, 591–603. [Google Scholar] [CrossRef] [Green Version]
  131. Went, P.T.; Lugli, A.; Meier, S.; Bundi, M.; Mirlacher, M.; Sauter, G.; Dirnhofer, S. Frequent EpCam protein expression in human carcinomas. Hum. Pathol. 2004, 35, 122–128. [Google Scholar] [CrossRef] [PubMed]
  132. Ma, X.; Kang, X.; He, L.; Zhou, J.; Zhou, J.; Sturm, M.B.; Beer, D.G.; Kuick, R.D.; Nancarrow, D.J.; Appelman, H.D.; et al. Identification of Tumor Specific Peptide as EpCAM Ligand and Its Potential Diagnostic and Therapeutic Clinical Application. Mol. Pharm. 2019, 16, 2199–2213. [Google Scholar] [CrossRef] [PubMed]
  133. Glumac, P.M.; Lebeau, A.M. The role of CD133 in cancer: A concise review. Clin. Transl. Med. 2018, 7, 18. [Google Scholar] [CrossRef]
  134. Yan, S.; Tang, D.; Hong, Z.; Wang, J.; Yao, H.; Lu, L.; Yi, H.; Fu, S.; Zheng, C.; He, G.; et al. CD133 peptide-conjugated pyropheophorbide-a as a novel photosensitizer for targeted photodynamic therapy in colorectal cancer stem cells. Biomater. Sci. 2020, 9, 2020–2031. [Google Scholar] [CrossRef] [PubMed]
  135. Bechara, C.; Sagan, S. Cell-penetrating peptides: 20 years later, where do we stand? FEBS Lett. 2013, 587, 1693–1702. [Google Scholar] [CrossRef]
  136. Kim, H.; Kitamatsu, M.; Ohtsuki, T. Enhanced intracellular peptide delivery by multivalent cell-penetrating peptide with bioreducible linkage. Bioorg. Med. Chem. Lett. 2018, 28, 378–381. [Google Scholar] [CrossRef]
  137. Huang, Y.-W.; Lee, H.-J. Cell-penetrating peptides for medical theranostics and targeted drug delivery. In Peptide Applications in Biomedicine, Biotechnology and Bioengineering; Elsevier: Amsterdam, The Netherlands, 2018; pp. 359–370. [Google Scholar] [CrossRef]
  138. Milletti, F. Cell-penetrating peptides: Classes, origin, and current landscape. Drug Discov. Today 2012, 17, 850–860. [Google Scholar] [CrossRef]
  139. Guidotti, G.; Brambilla, L.; Rossi, D. Cell-Penetrating Peptides: From Basic Research to Clinics. Trends Pharmacol. Sci. 2017, 38, 406–424. [Google Scholar] [CrossRef]
  140. Lundberg, P. A brief introduction to cell-penetrating peptides. J. Mol. Recognit. 2003, 16, 227–233. [Google Scholar] [CrossRef]
  141. Padari, K.; Koppel, K.; Lorents, A.; Hällbrink, M.; Mano, M.; de Lima, M.P.; Pooga, M. S413-PV Cell-Penetrating Peptide Forms Nanoparticle-Like Structures to Gain Entry Into Cells. Bioconjug. Chem. 2010, 21, 774–783. [Google Scholar] [CrossRef]
  142. Duchardt, F.; Ruttekolk, I.R.; Verdurmen, W.; Lortat-Jacob, H.; Bürck, J.; Hufnagel, H.; Fischer, R.; Heuvel, M.V.D.; Löwik, D.; Vuister, G.; et al. A Cell-penetrating Peptide Derived from Human Lactoferrin with Conformation-dependent Uptake Efficiency. J. Biol. Chem. 2009, 284, 36099–36108. [Google Scholar] [CrossRef] [Green Version]
  143. Bellet-Amalric, E.; Blaudez, D.; Desbat, B.; Graner, F.; Gauthier, F.; Renault, A. Interaction of the third helix of Antennapedia homeodomain and a phospholipid monolayer, studied by ellipsometry and PM-IRRAS at the air–water interface. Biochim. Biophys. Acta (BBA) Biomembr. 2000, 1467, 131–143. [Google Scholar] [CrossRef] [Green Version]
  144. Ziegler, A.; Blatter, X.L.; Seelig, A.; Seelig, J. Protein Transduction Domains of HIV-1 and SIV TAT Interact with Charged Lipid Vesicles. Binding Mechanism and Thermodynamic Analysis. Biochemistry 2003, 42, 9185–9194. [Google Scholar] [CrossRef]
  145. Lee, M.-T.; Hung, W.-C.; Chen, F.-Y.; Huang, H.W. Many-Body Effect of Antimicrobial Peptides: On the Correlation between Lipid’s Spontaneous Curvature and Pore Formation. Biophys. J. 2005, 89, 4006–4016. [Google Scholar] [CrossRef] [Green Version]
  146. Pouny, Y.; Rapaport, D.; Mor, A.; Nicolas, P.; Shai, Y. Interaction of antimicrobial dermaseptin and its fluorescently labeled analogs with phospholipid membranes. Biochemistry 1992, 31, 12416–12423. [Google Scholar] [CrossRef]
  147. Ter-Avetisyan, G.; Tünnemann, G.; Nowak, D.; Nitschke, M.; Herrmann, A.; Drab, M.; Cardoso, M.C. Cell Entry of Arginine-rich Peptides Is Independent of Endocytosis. J. Biol. Chem. 2009, 284, 3370–3378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Thakur, A.; Mishra, A.P.; Panda, B.; Sweta, K.; Majhi, B. Detection of Disease-Specific Parent Cells via Distinct Population of Nano-Vesicles by Machine Learning. Curr. Pharm. Des. 2020, 26, 3985–3996. [Google Scholar] [CrossRef] [PubMed]
  149. Thakur, A.; Mishra, A.P.; Panda, B.; Rodríguez, C.S.; Gaurav, I.; Majhi, B. Application of Artificial Intelligence in Pharmaceutical and Biomedical Studies. Curr. Pharm. Des. 2020, 26, 3569–3578. [Google Scholar] [CrossRef] [PubMed]
  150. Thakur, A.; Sidu, R.K.; Gaurav, I.; Sweta, K.; Chakraborty, P.; Thakur, S. Modified biopolymer-based systems for drug delivery to the brain. In Tailor-Made and Functionalized Biopolymer Systems; Woodhead Publishing: Shaxton, UK, 2021. [Google Scholar] [CrossRef]
  151. Gaurav, I.; Thakur, A.; Iyaswamy, A.; Wang, X.; Chen, X.; Yang, Z. Factors Affecting Extracellular Vesicles Based Drug Delivery Systems. Molecules 2021, 26, 1544. [Google Scholar] [CrossRef]
  152. Salunkhe, S.; Dheeraj; Basak, M.; Chitkara, D.; Mittal, A. Surface functionalization of exosomes for target-specific delivery and in vivo imaging & tracking: Strategies and significance. J. Control. Release 2020, 326, 599–614. [Google Scholar] [CrossRef]
  153. Jiao, H.; Zhao, X.; Han, J.; Zhang, J.; Wang, J. Synthesis of a novel 99mTc labeled GE11 peptide for EGFR SPECT imaging. Int. J. Radiat. Biol. 2020, 96, 1443–1451. [Google Scholar] [CrossRef] [PubMed]
  154. Agnes, R.S.; Broome, A.-M.; Wang, J.; Verma, A.; Lavik, K.; Basilion, J.P. An Optical Probe for Noninvasive Molecular Imaging of Orthotopic Brain Tumors Overexpressing Epidermal Growth Factor Receptor. Mol. Cancer Ther. 2012, 11, 2202–2211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Xu, W.-W.; Liu, D.-Y.; Cao, Y.-C.; Wang, X.-Y. GE11 peptide-conjugated nanoliposomes to enhance the combinational therapeutic efficacy of docetaxel and siRNA in laryngeal cancers. Int. J. Nanomed. 2017, 12, 6461–6470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Pi, J.; Jiang, J.; Cai, H.; Yang, F.; Jin, H.; Yang, P.; Cai, J.; Chen, Z.W. GE11 peptide conjugated selenium nanoparticles for EGFR targeted oridonin delivery to achieve enhanced anticancer efficacy by inhibiting EGFR-mediated PI3K/AKT and Ras/Raf/MEK/ERK pathways. Drug Deliv. 2017, 24, 1549–1564. [Google Scholar] [CrossRef] [Green Version]
  157. Ding, J.; Wang, K.; Tang, W.-J.; Li, D.; Wei, Y.-Z.; Lu, Y.; Li, Z.-H.; Liang, X.-F. Construction of Epidermal Growth Factor Receptor Peptide Magnetic Nanovesicles with Lipid Bilayers for Enhanced Capture of Liver Cancer Circulating Tumor Cells. Anal. Chem. 2016, 88, 8997–9003. [Google Scholar] [CrossRef] [Green Version]
  158. Rahmanian, N.; Hosseinimehr, S.J.; Khalaj, A.; Noaparast, Z.; Abedi, S.M.; Sabzevari, O. 99mTc-radiolabeled GE11-modified peptide for ovarian tumor targeting. DARU J. Pharm. Sci. 2017, 25, 13. [Google Scholar] [CrossRef] [Green Version]
  159. Ohno, S.-I.; Takanashi, M.; Sudo, K.; Ueda, S.; Ishikawa, A.; Matsuyama, N.; Fujita, K.; Mizutani, T.; Ohgi, T.; Ochiya, T.; et al. Systemically Injected Exosomes Targeted to EGFR Deliver Antitumor MicroRNA to Breast Cancer Cells. Mol. Ther. 2013, 21, 185–191. [Google Scholar] [CrossRef] [Green Version]
  160. Nakase, I. Biofunctional Peptide-Modified Extracellular Vesicles Enable Effective Intracellular Delivery via the Induction of Macropinocytosis. Processes 2021, 9, 224. [Google Scholar] [CrossRef]
Figure 1. Graphic depiction of major barriers obstructing the nano delivery system (NDS), such as the endosomal–lysosomal system, clearance of NDS via the mononuclear phagocytic system, and endothelial barrier acting in the event of extravasation of NDS in cancer.
Figure 1. Graphic depiction of major barriers obstructing the nano delivery system (NDS), such as the endosomal–lysosomal system, clearance of NDS via the mononuclear phagocytic system, and endothelial barrier acting in the event of extravasation of NDS in cancer.
Pharmaceutics 13 01433 g001
Figure 2. Schematic illustration showing the peptide-functionalized liposomal NDS acting on receptors overexpressed on the surface of cancer cells via targeted delivery. The peptide conjugated to the NDS binds specifically to the receptors upregulated on the surface of cancer cells, which is followed by its uptake by the cancer cells through receptor-mediated endocytosis. Subsequently, the payload of the NDS is released by the degradation of the lipid bilayer via the endosomal–lysosomal pathway.
Figure 2. Schematic illustration showing the peptide-functionalized liposomal NDS acting on receptors overexpressed on the surface of cancer cells via targeted delivery. The peptide conjugated to the NDS binds specifically to the receptors upregulated on the surface of cancer cells, which is followed by its uptake by the cancer cells through receptor-mediated endocytosis. Subsequently, the payload of the NDS is released by the degradation of the lipid bilayer via the endosomal–lysosomal pathway.
Pharmaceutics 13 01433 g002
Table 1. Peptides used in conjugation with NDS for targeting cancer-specific receptors.
Table 1. Peptides used in conjugation with NDS for targeting cancer-specific receptors.
Type of CancerTarget ReceptorPeptideRef.
Breast cancerSSTROctreotide[61]
α1β5 integrinATN-161[62]
αvβ3 integrinCyclic RGD[63]
HER2KCCYSL[64]
AHNP[65]
Colon cancer αvβ3 integrinCyclic RGD[63]
Fibrosarcoma AminopeptidaseNGR[66]
Glioma SSTROctreotide[67]
αvβ3 integrinCyclic RGD[68]
TFRT7/TAT[69]
AminopeptidaseNGR[70]
Lung CancerSSTROctreotide[71]
TFRT7/TAT[72]
LHRHLHRL[73]
Melanomaαvβ3 integrinRGD[74]
αvβ3 integrinCyclic RGD[63]
Ovarian cancer TFRT7[75]
HER2LTVSPWY[76]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gaurav, I.; Wang, X.; Thakur, A.; Iyaswamy, A.; Thakur, S.; Chen, X.; Kumar, G.; Li, M.; Yang, Z. Peptide-Conjugated Nano Delivery Systems for Therapy and Diagnosis of Cancer. Pharmaceutics 2021, 13, 1433. https://doi.org/10.3390/pharmaceutics13091433

AMA Style

Gaurav I, Wang X, Thakur A, Iyaswamy A, Thakur S, Chen X, Kumar G, Li M, Yang Z. Peptide-Conjugated Nano Delivery Systems for Therapy and Diagnosis of Cancer. Pharmaceutics. 2021; 13(9):1433. https://doi.org/10.3390/pharmaceutics13091433

Chicago/Turabian Style

Gaurav, Isha, Xuehan Wang, Abhimanyu Thakur, Ashok Iyaswamy, Sudha Thakur, Xiaoyu Chen, Gaurav Kumar, Min Li, and Zhijun Yang. 2021. "Peptide-Conjugated Nano Delivery Systems for Therapy and Diagnosis of Cancer" Pharmaceutics 13, no. 9: 1433. https://doi.org/10.3390/pharmaceutics13091433

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