Next Article in Journal
Preferences of Healthcare Professionals on 3D-Printed Tablets: A Pilot Study
Previous Article in Journal
Combination of Lanosterol and Nilvadipine Nanosuspensions Rescues Lens Opacification in Selenite-Induced Cataractic Rats
Previous Article in Special Issue
Peptide Targeted Gold Nanoplatform Carrying miR-145 Induces Antitumoral Effects in Ovarian Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress of Conjugated Nanomedicine for Cancer Treatment

1
The Key Laboratory of Biomedical Information Engineering of Ministry of Education, School of Life Science and Technology, Xi’an Jiaotong University, Xi’an 710049, China
2
Department of Epidemiology, Shaanxi Provincial Cancer Hospital, Xi’an 710061, China
3
Shaanxi Provincial Centre for Disease Control and Prevention, Xi’an 710054, China
4
Center of Digestive Endoscopy, Shaanxi Provincial Cancer Hospital, Xi’an 710061, China
5
Research Institute of Xi’an Jiaotong University, Hangzhou 311200, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Pharmaceutics 2022, 14(7), 1522; https://doi.org/10.3390/pharmaceutics14071522
Submission received: 26 June 2022 / Revised: 14 July 2022 / Accepted: 19 July 2022 / Published: 21 July 2022
(This article belongs to the Special Issue Bioconjugation and Nanomaterials for Clinical Translation)

Abstract

:
The conventional cancer therapeutic modalities include surgery, chemotherapy and radiotherapy. Although immunotherapy and targeted therapy are also widely used in cancer treatment, chemotherapy remains the cornerstone of tumor treatment. With the rapid development of nanotechnology, nanomedicine is believed to be an emerging field to further improve the efficacy of chemotherapy. Until now, there are more than 17 kinds of nanomedicine for cancer therapy approved globally. Thereinto, conjugated nanomedicine, as an important type of nanomedicine, can not only possess the targeted delivery of chemotherapeutics with great precision but also achieve controlled drug release to avoid adverse effects. Meanwhile, conjugated nanomedicine provides the platform for combining several different therapeutic approaches (chemotherapy, photothermal therapy, photodynamic therapy, thermodynamic therapy, immunotherapy, etc.) with the purpose of achieving synergistic effects during cancer treatment. Therefore, this review focuses on conjugated nanomedicine and its various applications in synergistic chemotherapy. Additionally, the further perspectives and challenges of the conjugated nanomedicine are also addressed, which clarifies the design direction of a new generation of conjugated nanomedicine and facilitates the translation of them from the bench to the bedside.

1. Introduction

Cancer is the leading cause of death worldwide and a critical barrier to increasing life expectancy. According to the estimation of the World Health Organization (WHO), nearly 19.3 million new cases and 10 million deaths are closely related to cancer globally in 2020 [1]. Although new types of systemic cancer therapy (e.g., immunotherapy and targeted therapy) have been developed over the past few years, it is undeniable that chemotherapy still occupies a crucial position [1,2]. However, conventional chemotherapy suffers from intrinsic limitations, such as serious side effects, intrinsic and acquired multidrug resistance (MDR) and poor targeting capacity, causing more than 90% of new drugs to be abandoned before or during clinical trials [3]. Hence, the technologies that effectively improve the pharmacokinetics and the tumor accumulation of chemotherapeutic drugs are urgently needed [3,4]. Thereinto, employing carriers to deliver anti-cancer drugs to tumor tissues is a promising and innovative approach to increasing the efficacy of drugs while avoiding adverse effects [5,6,7].
Nanotechnology has been regarded as the next revolution to influence various industrial fields, including biomedicine. Especially, nanomedicine for cancer treatment has achieved tremendous progress in the last few decades. Due to unique physical and chemical properties, nanomedicine can enhance the pharmacokinetics of drugs via improving their stability, solubility, circulating half-life, and targeting capacity, and overcome drug resistance, finally resulting in enhanced chemotherapeutic efficacy. With the gradual maturity of nanomedicine research, nanomedicine has gradually entered clinical tumor treatment. Since the first cancer nanomedicine, Doxil, was approved by the Food and Drug Administration (FDA) in 1995, there have been at least 17 kinds of cancer nanomedicine applied in the clinic, including Abraxane (albumin-bound paclitaxel), Oncaspar (PEGylated L-asparaginase), Genexol-PM (mPEG-PDLLA micellar paclitaxel), NanoTherm® AS1, etc. (Table 1). In addition to these commercial nanomedicines, a vast number of chemotherapeutic nanomedicines are undergoing clinical trials and experimental studies currently, encompassing many types of nanomedicine, including liposomes, polymer-based nanoparticles, albumin nanoparticles as well as inorganic nanoparticles [8,9].
Conjugated nanomedicine, formulated from drug conjugates that connect drugs via chemical linkers, has attracted more attention due to its unique potential in cancer treatment, which are listed as follows: 1. It increases the drug-carrying capacity of chemotherapeutic drugs. Both hydrophilic and hydrophobic substances nanomedicine can selectively deliver chemotherapeutics to tumor cells, even intracellular organelles via the passive and active targeting capacity, especially ones containing targeting ligands. 2. Several stimuli-responsive chemical bonds can be exploited as linkers to form drug conjugates, with the purpose of selective and controlled drug release. 3. Some medical imaging agents can also be conjugated with therapeutic agents for a diagnostic function. 4. Since fewer excipients without therapeutic effects are required, the biohazard caused by the degradation of the excipients can be effectively minimized. 5. The conjugated nanomedicine is also easily mass-produced due to the simple structure and easy preparation of drug conjugates. Based on these advantages, conjugated nanomedicine is considered a hopeful strategy to ameliorate cancer treatment outcomes in the future. Indeed, in recent years, conjugated nanomedicine has obtained more and more attention and the number of publications per year with the keyword “conjugate nanomedicine” maintains a high level (Figure 1, from the web of science database). Therefore, this review provides an overview of conjugated nanomedicines, summarizing the types of drug conjugates and representative applications of these conjugated nanomedicines in synergistic chemotherapy. Furthermore, the further perspectives and challenges of the conjugated nanomedicine are also addressed, which clarifies the design direction of a new generation of conjugated nanomedicine and facilitates the translation of them from bench to bedside (Figure 2).

2. Categories of Drug Conjugates

Conjugated nanomedicine formulated by drug conjugates has proved to enhance the therapeutic effect of drugs through improving their transmission to the disease sites by the virtue of the ligand-mediated active targeting process and enhanced permeability and retention (EPR) effect-based passive targeting process. Besides, the physicochemical properties of nanomedicine, such as size, shape, and surface chemical properties also determine the accumulation and deep penetration of nanoparticles into tumor tissues [15]. Besides, the controllable release of drugs under certain internal/external stimuli is also crucial for the desired therapeutic outcomes. To develop conjugated nanomedicine with properties as required, the design and preparation of promising drug conjugates is the first step toward success. The following sections will discuss the categories of drug conjugates in detail along with several representative examples.

2.1. Polymer-Drug Conjugates

Polymer-drug conjugates, also known as polymer prodrugs, are usually composed of polymer skeletons, chemical linkers (cleavable or uncleavable) and therapeutic drugs [16]. The first study of polymer-drug conjugates can be traced back to 1955 when Von Horst Jatzkewitz found that mescaline, a kind of psychedelic alkaloid, presented biological activity in mice even being coupled with poly (vinylpyrrolidone), accompanied by an extended residence time of mescaline in vivo [17]. In the 1970s, Ringsdorff, Kopecek and Duncan pioneered the development of therapeutic strategies based on polymer-drug conjugates [18,19]. Thereinto, Ringsdorff put forward the concept of a pharmacologically active polymer carrier, which can improve drug solubility and control drug release in a targeted way. About twenty years later, a new kind of polymer–protein conjugate named Adagen® (pegademase bovine) was first approved by FDA for enzyme replacement therapy against severe combined immunodeficiency disease (SCID) associated with a deficiency of adenosine deaminase [18,19]. Since then, this drug delivery strategy received much more attention and a series of polymer–drug conjugates have been prepared for treating various illnesses. Furthermore, these conjugates could also be formulated into nanomedicines with types of polymer capsules [20,21], polymeric nanoparticles [22,23,24], dendrimers [25,26] and so on.
As for the historical development of polymer–drug conjugates in the field of cancer treatment, Matsumura and Maeda reported the antitumor carcinogen SMANCS, a conjugate of partially half-butyl-esterified styrene-co-maleic acid polymer[butyl-SMA] and neocarzinostatin (NCS), which preferentially accumulated in tumor tissues after intravenous administration in the 1980s [27]. Besides, PK1 (FCE28068) is the first water-soluble polymer small molecular drug conjugate applied in clinical trials. It is composed of N-(2-hydroxypropyl) methacrylamide (HPMA) and doxorubicin (DOX) linked via a peptide-based lysosome-cleavable bond. Based on the results from clinical trials, PK1 displays a prolonged half-life of DOX and improved biosafety as compared with parent DOX treatment. It also displays the antineoplastic activity against breast cancer and non-small cell lung cancer (NSCLC) during Phase II clinical studies [28]. However, the tumor accumulation of PK1 is still insufficient, compromising the final therapeutic efficacy and inducing the termination of PK1 development [27]. Subsequently, Jameson and co-workers formed the conjugate Onzeald of tetrapod polyethylene glycol(PEG) and irinotecan (antineoplastic enzyme inhibitor) through lipid bonds. During a preclinical study, the half-life and tumoral concentration of irinotecan increased as compared with parent irinotecan, leading to an inhibited tumor growth and improved therapeutic index [29]. Learning from the above examples the majority of cytotoxic chemotherapeutic drugs present problems including poor aqueous solubility, limited tumor exposure and off-target toxicity. Once the drugs are linked to polymer carriers, their aqueous solubility and stability can be improved along with modified pharmacokinetics. Moreover, it also shows the potential for polymer–drug conjugates to precisely deliver drugs to tumor tissues via passive targeting and/or active targeting effects, and overcoming the undesired side effects of cytotoxic chemotherapeutics against healthy tissues [30]. In addition, these conjugated polymeric nanoparticles also displayed advantages, including overcoming MDR and reducing immunogenicity [31].
In the past decade, many polymer-drug conjugates have been commercially available and some others are undergoing clinical trials [30], on account of the rapid advancement of polymeric materials. Among them, PEG occupies the important position, which are represented by a series of clinically applied PEG-drug conjugates (e.g., Adynovate®, Oncaspar®, Plegridy®, etc.) [32,33]. Additionally, some synthetic biodegradable polymers bearing minimal long-term side effects and systemic toxicity, especially the ones approved by the FDA, have also been used to prepare the polymer-drug conjugates, such as poly(lactic-co-glycolic acid) (PLGA) [34], poly(lactic acid) (PLA) [35], poly(ε-caprolactone) (PCL) [36], poly(amino acid) (PAA) [37] and so on. For example, the allyl functionalized PLA is used as the precursor of the polymer backbone. A UV-induced mercaptan-alkene reaction was carried out to combine sulfobetaine (SB) with a PLA-based backbone to yield the carrier materials, followed by the encapsulation of paclitaxel (PTX). This drug delivery system based on polymer–drug coupling showed complete degradability, continuous drug release ability and significant anticancer effect [38]. PAA has also been widely used in the field of drug delivery systems due to its good biocompatibility, biodegradability and functionalization [39]. Because there are many active functional groups on the side chain of PAA, the drugs and bioactive molecules could be easily conjugated to PAAs, followed by preparation into nanoparticles for cancer treatment [40]. For example, Ma and co-workers conjugated dexamethasone (DEX, an anti-inflammatory agent) onto the side chain of mPEG-b-poly(L-lysine) (mPEG-b-PLL) via a redox and pH dual sensitive linker. The resultant polymer–drug conjugates were further self-assembled into micelles for the treatment of colorectal cancer. According to their results, cyclooxygenase-2 (COX-2) and α-smooth muscle actin (α-SMA) were dramatically decreased during therapy and the immunosuppressive microenvironment of the CT26 tumor was also relieved, resulting in the compromised tumor-promoting inflammation. It also provided a promising option for applying anti-inflammatory drugs for cancer treatment [41]. Natural polysaccharides, including chondroitin sulfate (CS), hyaluronic acid (HA), pullulan (PUL), and heparin (HEP) are also good candidates for conjugating drugs due to their outstanding virtues, such as biocompatibility, biodegradability, non-immunogenicity and toxicity, easy chemical modification, and low cost [42,43,44,45,46]. The antimalarial drug dihydroartemisinin (DHA) can inhibit cancer cell proliferation and induce apoptosis, while it has been associated with some limitations, such as poor aqueous solubility and rapid metabolism in the systemic circulation. Robin and co-workers successfully synthesized a HA–DHA conjugate via the conjugation of the carboxylic group of HA with the hydroxyl group of DHA. The enhanced cytotoxicity of nanoparticles against lung cancer (A549 cell line) was supported by the generation of reactive oxygen species (ROS), loss of mitochondrial membrane potential and exhibition of better cytotoxicity than native DHA [47].

2.2. Antibody-Drug Conjugates

Antibody–drug conjugates (ADC) are conjugates that couple specific monoclonal antibodies with cytotoxic drugs through specific linkages. They are targeted biological agents, which can selectively transport cytotoxic drugs to tumor cells with monoclonal antibodies as navigation. ADC generally includes the following three components: antibodies with high specificity and affinity to targets, connectors with high stability and small-molecule cytotoxic drugs with promising therapeutic efficacy [48]. ADC selectively delivers cytotoxic drugs to tumor cells by using the specific binding ability of monoclonal antibodies to cell surface target antigens. After ADC enters the body, it recognizes the specific antigen on the cell surface through autoantibody components. Then, the ADC antigen complex can be internalized by tumor cells via receptor-mediated endocytosis, followed by being degraded by intracellular enzymes and lysosomes. Cytotoxic agents are then released in the cytoplasm, finally causing cell death through an induction mechanism [49] (Figure 3). Moreover, chemical connectors in ADC with promising stimuli-responsiveness have been recognized as the prerequisites for minimizing the premature drug release in plasma and promoting the controllable release of payloads to cancer cells, which kindles the enthusiasm of researchers for constituting sensitive ADC, especially in the field of tumor precision therapy for decades.
At present, FDA has approved a series of ADCs for clinical cancer treatment, and many other ADCs are undergoing clinical trials [50]. As for traditional chemotherapy, anticancer drugs mainly execute rapidly dividing cells, but undesirable toxicity to normal cells is inevitable, resulting in high side effects. To solve this problem, the first generation of ADCs was developed by connecting cytotoxic drugs with monoclonal antibodies (mAb). These ADCs can target cancer cells and then selectively destroy them with lower side effects on healthy tissues. Although the role of monoclonal antibodies in cancer treatment was not fully understood at that time, inspired by the fact that many antibodies can preferentially bind to tumor cells, people connected anticancer drugs, such as melphalan, vinblastine, methotrexate, and DOX to monoclonal antibodies to form the first generation of ADCs. In 2000, the first ADC, gemtuzumab ozogamicin (GO), was approved by FDA, mainly for the treatment of patients with acute myeloid leukemia (AML). GO is a conjugate of an anti-CD33 mAb and calicheamicin, which are connected via the acid-cleavable hydrazone linkers. However, the results from the required post-approval study demonstrated the inadequate improvement of the survival rate and more severe toxicity of GO-based synergistic therapy over chemotherapy alone, forcing Pfizer to withdraw GO from the market in 2010 [51]. Some other studies also pointed out the drawbacks of first-generation ADCs, including poor potency of the cytotoxic drug, low localization of monoclonal antibodies and poor stability of linkages [50,51,52,53,54,55] Moreover, an undesirable immune response could be induced during ADCs-mediated therapy, which is caused by antibody components, rather than cytotoxic drugs [56].
The second generation of ADCs were optimized by using more effective tubulin targeting agents (e.g., monomethyl auristatin E (MMAE), maidenlig nindm1) for therapeutic purposes. For example, brentuximabvedotin (sgn-35) containing MMAE was developed for the treatment of Hodgkin’s and anaplastic large cell lymphoma. As for Ado-Trastuzumab Emtansine, also known as T-dm1, the antibody trastuzumab (humanized IgG1 anti HER-2 antibody) is chemically linked to the drug maidenlignindm1. Noteworthily, once trastuzumab binds to the HER-2/neu receptor on target cells, it can also prevent homologous or heterodimerization of the receptor (HER2/HER3) and inhibit the activation of mitogen activated protein-kinase(MAPK) and PI3K/AKT signal pathway, finally preventing the growth of tumor cells. At the same time, T-dm1 can be internalized into cancer cells and then binds to tubulin to induce cell death [57,58]. Although efforts have been made to develop ADCs, studies have shown that the blood stability of the second generation ADCs is not ideal, showing unfavorable in vivo toxicity, such as hepatotoxicity, cardiotoxicity, peripheral neuropathy, thrombocytopenia, and ocular toxicity [59].
After the emergence of clinical problems related to the second generation of ADCs, the third generation of ADCs were designed and prepared. The main understanding of the third generation ADCs is to design immunoglobulin G (IgG) molecules with proper drug binding positions, so as to obtain more uniform drug conjugates. In the third generation of ADCs, the instability of ADC in circulation can be overcome by biocoupling chemistry. The main goal is to reduce the uncoupling of drugs in blood circulation for minimized off-target toxicity, improve the therapeutic index and expand the treatment window. Seattle Genetics has developed Vadastuximab talirine (SGN-CD33A), which contains a novel synthetic pyrrolobenzodiazepine (PBD) dimer. It was coupled to a humanized anti-CD33 IgG1 antibody through a maleimidocaproyl valine–alanine dipeptide linker and is structurally related to anthramycin, leading to targets cell death by cross-linking DNA and effectively preventing cell division. Vadastuximab talirine not only demonstrates robust activity in a series of acute myeloid leukaemia(AML) animal models but also overcomes transporter-mediated MDR [59,60,61,62]. Additionally, the better introduction of “cleavable” linkers between antibodies and payloads also promotes the clinical development of the third generation of ADCs, due to their strengths in controllable drug release at the target site. These cleavable linkers include acid-sensitive linkers (e.g., hydrazone linkage), reducible disulfides, and enzyme cleavable linkers, which have been comprehensively reviewed by Bargh and co-workers [63]. Taking reducible disulfide as an example, Pillow and co-workers directly attached maytansinoid bearing thiols (DM1) to engineered cysteine residues in an antibody. The obtained ADCs were stable during blood circulation due to the shielding of disulfides by antibodies. Once internalized by target cells followed by antibody catabolism, the disulfide linkers were exposed to the reducing cellular environment, leading to a rapid disulfide catabolite and desirable drug release [64]. Rémy Gébleux and co-workers also developed two ADCs by using a similar coupling strategy, denoted as SIP(F8)-SS-DM1 and IgG(F8)-SS-DM1. In this work, the F8 antibody, directed against the alternatively spliced Extra Domain A (EDA) domain of fibronectin was selected, but in two different formats which are IgG and small immune protein (SIP), respectively. Based on their results, IgG(F8)-SS-DM1 was more stable in mouse plasma than SIP(F8)-SS-DM1, demonstrating a novel mechanism in the drug release from the disulfide-based ADCs. However, the ADCs in SIP format displayed a better therapeutic outcome compared with ADCs in the IgG format in immunocompetent mice bearing F9 tumors, revealing that the format of antibodies plays a significant role in determining the final therapeutic efficacy [65]. Immunotoxins, that ADCs created by chemically conjugating antibodies to whole protein toxins (lack of natural binding domain), have been proved to enhance the performance of toxins by taking the advantage of the desired specificity of antibodies to the target cells and the potency of the toxins to kill cells effectively [66,67]. In this kind of ADC, the disulfide linkers could also be introduced to control the release of toxins. For instance, two immunotoxins (ITxs) were constructed via chemical conjugation (disulfide linker) of the ribosome-inactivating proteins (Saporin-S6) with anti-CD20 mAb Rituximab, with the difference in the structures (monomeric and dimeric) and molecular weight, known as HMW-ITx (dimeric) and LMW-ITx (monomeric), respectively. Accordingly, the HMW-ITx was more cytotoxic than the LMW-ITx in two CD20+ lymphoma cell lines, Raji and D430B, thanks to its higher toxin loading and more efficient antigen capping, although they displayed similar activity in inhibiting protein synthesis in a cell-free system. Moreover, as compared with parent Saporin-S6, both ITxs are more active [68].
The third generation ADCs have many other advantages, including improved stability, optimized pharmacokinetics, slow deconjugation, and high activity against cells that express lower levels of antigen. These ADCs are site-specific and homoconjugated, which provides a basis for applying ADCs in cancer treatment. However, many of these ADCs are still in the research stage, and only a relatively small number have been tested in the clinic. Although they can improve cancer treatment, there are still concerns about their limited treatment range [69,70,71,72]. More details about the potential of ADCs have been well reviewed elsewhere, along with their status in the oncology market [73]. At present, a total of 14 ADC drugs in the world have been approved for listing (Table 2). Among them, myLotarg of Pfizer/Wyeth is the first ADC drug to be listed, for the treatment of acute myeloid leukemia, but due to fatal liver injury from Mylotarg, Pfizer withdrew Mylotarg in 2010. In 2021, FDA successively approved the listing application of Zvnionta and Tivdak. The former was a CD19 ADC drug and the latter is a tissue factor (TF) ADC drug [41,74,75].

2.3. Peptide-Drug Conjugates

Peptides are short chains of amino acids, which are distinguished from proteins by their shorter length. Based on the definition from the FDA, a peptide is a polymer composed of less than 40 amino acids (500–5000 Da). Peptide–drug conjugates (PDCs) have a structure similar to that of ADCs, which also consist of three important components, including functional peptides, linkers and cytotoxic drugs. In 2021, Melflufen® (melphalan flufenamide), as a first-in-class anticancer PDC, was approved by the FDA for the treatment of relapsed and refractory multiple myeloma (RRMM) in combination with dexamethasone. Once in tumor cells, the aminopeptidase fusion domain of Melflufen can be affected by aminopeptidase and lipase, and the hydrophilic alkylating agent melphalan is further released to inhibit the DNA reparation and angiogenesis during therapy. Moreover, a series of hydroxypeptidases expressed in multiple myeloma and other tumors can be cracked by a hydrolysis reaction and quickly released in the cytotoxic payload from Melflufen [76].
For some PDCs, the homing peptides are responsible for directing the entire PDCs to the targeted tumor cells by recognizing the specific receptors overexpressed on the cellular surface, with the purpose of decreasing the side effects from off-target delivery. These homing peptides include RGD (targeting integrins), GnRH (targeting gonadotropin releasing hormone receptor(GnRH-R)), SST (targeting somatostatin receptors (SSTR1–5)), EGF (targeting epidermal growth factor receptor (EGFR)), Angiopep-2 (targeting low-density lipoprotein receptor-related protein-1 (LRP-1)) and so on [77]. It has been reported that the secondary structure of these peptides pronouncedly influences their binding affinity. Moreover, there are still some limitations compromising their homing effects, such as fast degradation by enzymes at terminal sites, chemical instability and rapid renal clearance. To circumvent these problems, the techniques for cyclization and stapling of linear peptides have been developed with the details reviewed by Cooper et al. recently [78]. In addition to the linear peptides, peptide dendrimers have also been studied as building blocks in PDCs. The employment of peptide dendrimers is considered to be beneficial because of their adjustable amino acid characteristics and good biocompatibility. Recently, Oliveria and co-workers developed the peptide dendrimer–gemcitabine (GEM) conjugates for the treatment of colorectal cancer. YIGSR, a kind of peptide has been shown to selectively bind to laminin receptors (LR) overexpressed in many cancers. By conjugating YIGSR peptide with GEM, YIGSR-conjugated peptide dendrimers could be selectively internalized into HCT-116 cancer cells with high expression of LR [79]. Although peptides are currently underrepresented in clinical trials compared with small molecules and biological agents, they still provide excellent versatility and can help to design targeted therapies.
In addition to the peptides with the function of cell targeting, a series of cell penetrating peptides (CPP), that promote the drugs to enter cells through a non-specific mechanism, have also been prepared to develop the PDCs. CPPs usually have characteristic features, such as hydrophobicity, amphipathicity and net positive charge. CPPs-mediated cellular internalization is an energy-dependent cell process, such as endocytosis or receptor-mediated uptake [80,81]. Although CPPs can enhance the in vitro and in vivo efficacy of impenetrable molecules in biomedical applications, it still has the limitations of low osmotic concentration and poor target selectivity [82,83,84,85]. Furthermore, cationic CPPs present problems, such as their inability to selectively home to the target. To improve the performance of CPPs-based PDCs, recently, a study found that multimers of lysine (K) and leucine (L) of the amphiphilic α-helical LK sequence can penetrate cells at the nanomolar level, which was 100–1000 folds lower than the transverse concentration of traditional CPPs [86]. The stronger the interaction between CPPs and cell surface receptors, the faster the PDCs enter the tumor cells [87,88]. Similarly, other research groups have reported that amphiphilic CPP in the form of dimers showed stronger cell penetration activity than monomer CPP [89,90]. Amphiphilic cyclic cell-penetrating peptides (cCPPs) are a relatively new class of peptides. The cCPPs possess several advantages over the conventional linear CPPs, such as low immunogenicity, proteolytic resistance, improved cellular uptake, facilitated serum stability, and better interaction with the membrane receptors. Park and co-workers conjugated cabazitaxel (CBT) to a cCPP via an ester bond to assist CBT to penetrate into the tumoral cells. The conjugates showed less toxicity to normal human embryonic kidney (HEK-293) cells compared to free CBT while displaying approximately three- to four-fold higher antiproliferative activity on cancer cell lines, compared to the free CBT analog. Although the increased drug delivery can be attributed to the presence of CPPs, the widespread use of these types of PDCs is still limited due to their low cellular specificity [91].
Moreover, some peptides in PDCs can also be cleaved under various stimuli, acting as linkers for controllable drug release. Tripodi and co-workers conjugated daunorubicin with a cyclic peptide containing the NGR motif, which is a tripeptide sequence capable of recognizing CD13 receptor isoforms on tumor cells. An enzyme-cleavable tetrapeptide (GFLG) was selected to conjugate the cyclic peptide with daunorubicin. Cathepsin B overexpressed in the tumor microenvironment (TME) could cleave GFLG connectors to daunorubicin to avoid the undesired toxicity on normal tissues. Based on the results, the mice bearing subcutaneous Kaposi’s sarcoma can tolerate this conjugate and express plasma stability and antitumor activity both in vitro and in vivo. Compared with free daunorubicin, the PDC-based therapy indeed decreased toxic side effects and improved the efficacy of tumor growth inhibition in mice [92].
In another case, the peptides in PDCs could also participate in the assembly of nanoparticles, facilitating the entry of drugs into the nucleus, finally influencing the nuclear DNA or related enzymes for a therapeutic purpose. In detail, Cai and co-workers combined negatively charged 10-hydroxycamptothecin (HCPT) with peptide amphiphiles FFERGD to obtain HCPT–FFERGD (HP) firstly, followed by forming two kinds of complexes with positively charged cisplatin (complex 1 and 2) by adjusting the molar ratio of HP and cisplatin. Interestingly, complex 1 (HP/cisplatin = 1:1) and complex 2 (HP/cisplatin = 1:1.5) can self-assemble into nanostructures of different shapes, which were rod-shaped nanofibers for complex 1 and spherical nanoparticles for complex 2, respectively. These obtained nanostructures exhibited the following advantages: Firstly, HCPT and positively charged platinum cannot easily enter the cell membrane but can enter the cell membrane smoothly after coupling with a polypeptide. Secondly, both rod-shaped nanofibers and spherical nanoparticles can effectively enter the nucleus, which was not observed for the group of HCPT and HP-based nanoparticles. Thirdly, the results also showed that both rod-shaped nanofibers and spherical nanoparticles not only effectively inhibited cancer cells in vitro and in vivo but also enhanced the inhibitory effect on drug-resistant tumor cells. The team likened their nanomedicine to a “Trojan horse”, that transports soldiers (anticancer drugs) through the walls of the castle (the cell and nuclear membrane) for precise targeted and synergistic therapy (Figure 4) [93].

2.4. Drug-Drug Conjugates

2.4.1. Conjugates of Multiple Chemotherapeutic Drugs

Drug–drug conjugates present accurately defined chemical structures with a low molecular weight and high drug loading content (even up to 100%) as compared with other drug conjugates. They are usually connected by two anticancer drugs, which can self-assemble into nanoparticles without additional stabilizers. The formed nanoparticles can enhance drug accumulation in tumors through the EPR effect. Under certain stimuli, the structural integrity of nanoparticles could be damaged along with the rapid release of drugs for the synergistic chemo–chemo therapy. Compared with a single administration, the conjugates of these chemotherapeutic drugs showed excellent cytotoxicity and fewer side effects [75,94,95]. In some cases, two identical chemotherapeutic drugs are bound to the same active site through appropriate junctions to form the drug–drug conjugates, known as drug dimers. The self-assembly of drug dimers can improve the water solubility of hydrophobic drugs as well as their stability in physiological conditions. For example, Li and co-workers prepared a doxorubicin dimer (D-DoxCAR), which is synthesized via a carbamate linkage. It was further prepared into nanoparticles (D-DoxCAR-NP) with a drug content up to 86%. The acidic tumor environment induced the cleavage of carbamate linker to induce the rapid release of DOX, resulting in an enhanced anti-tumor effect on HepG2 cells compared with parent DOX [96].
When two therapeutic drugs with different aqueous solubility and therapeutic mechanisms are combined through a properly designed joint, the therapeutic drug–drug conjugations could be obtained for self-assembly into nanoparticles. The resulting nanoparticles showed enhanced therapeutic efficacy through synergistic therapy compared to drug administration alone [97]. For instance, Podophyllotoxin (PPT) displays a significant anticancer effect by destabilizing microtubules and preventing cell division. However, it has been associated with some limitations, such as poor aqueous solubility and severe off-target side effects. To overcome these disadvantages, Hou and co-workers conjugated hydrophobic PPT and hydrophilic methotrexate (MTX) via a reduction-responsive disulfide bond. The nanoprodrug formulated by this amphiphilic drug conjugates the successfully improved tumor delivery of PPT [98]. Huang and co-workers synthesized an amphiphilic drug–drug conjugate from the hydrophilic anticancer drug irinotecan (Ir) and hydrophobic anticancer drug chloramphenicol (Cb) through hydrolyzable ester ligation. Amphiphilic Ir-Cb then self-assembled into nanoparticles and achieved passive tumor targeting via the EPR effect. The ester bonds were further cleaved in tumor cells to realize the recovery of cytotoxicity of both drugs [99]. The biological applications of curcumin (CCM) are severely limited by some undesired properties, such as poor water solubility, short serum half-life, and low bioavailability. Cheng and co-workers utilized bifunctional PEG, bearing both azide and carboxylic acid groups, to realize the conjugation of Erlotinib (ELT) and CCM. Compared with free drugs, the conjugates prolonged the half-life of drugs in blood retention, and enhanced drug accumulation in tumor tissues (Figure 5A–C) [99].

2.4.2. Photothermal Agent-Drug Conjugates

Photothermal therapy (PTT), as one type of phototherapy, uses the heat generated from photothermal agents (PTAs) upon near-infrared (NIR) laser irradiation to induce the death of cancer cells [100,101,102,103,104]. The PTAs should meet the requirements, including low cytotoxicity, easy preparation and modification, and good solubility in biocompatible liquids. Besides, the absorption of PTA is usually adjusted to the range between 750 and 1350 nm (Biological Windows (BWs), BW I: 700 nm–980 nm/BW II: 1000 nm–1350 nm) to enhance the penetration of light in the tissues. The ideal PTA should also have a higher photothermal conversion efficiency and good accumulation in tumor tissues. Nowadays, the PTAs can be divided into two categories, which are inorganic PTAs (e.g., metallic NPs, carbon-based NPs) and organic molecular-based ones, respectively, both of which could be prepared into drug conjugates for cancer treatment. In this section, we focus on the organic PTAs-related conjugation and the inorganic PTAs-related ones will be detailed and introduced in Section 2.5 “Inorganic Nanoparticle-Drug Conjugates”.
For organic molecular-based PTAs, their electrons absorb photon energy upon light illumination and transform from the ground state to the excited singlet state. When the electrons return to the ground state, due to the non-radiative relaxation processes in which excited singlet states collide with their neighboring molecules, photothermal effects could be induced. The increase in kinetic energy leads to heat surrounding the microenvironment and induces irreversible damage to tissues when the temperature exceeds the threshold [105]. Organic PTA includes organic dye molecules (e.g., indocyanine green and heptylcyanine) and organic nanoparticles (e.g., semiconductor polymer nanoparticles (SPNP)) [106,107]. PTT can target tumors precisely with adjustable doses of external laser irradiation, thus minimizing the damage to surrounding healthy tissue. It has been recognized as an effective non-invasive therapy available for most types of cancer [108]. However, there are still some obstacles during PTT, such as low transmission efficiency of PTA in tumors, and heat-induced overexpression of heat shock proteins (HSPs) in tumors post-PTT responsible for thermal resistance. Therefore, the combination of PTAs with anticancer drugs can potentially overcome these drawbacks of PTT in a synergistic manner [109,110,111,112]. Furthermore, PTAs–drug conjugates related nanomedicines continue to present the advantages of drug conjugates for an improved therapeutic outcome. For example, in order to improve the transfer efficiency of photothermal agent IR820 and the chemotherapeutic drug PTX, Zhang and co-workers coupled IR820 with PTX to form pH and enzyme-sensitive carrier-free nanopharmaceuticals, which were used for fluorescence imaging-guided synergistic chemotherapy-PTT. The nanosystems show high drug loading content and promising stability. IR820-PTX also solved the problems of the short life span of IR820 in vivo and poor solubility of PTX and effectively inhibited tumor growth via combined PTT and chemotherapy [112]. Similarly, Ao and co-workers linked camptothecin (CPT) to IR820 via a redox disulfide ligand to form the PTA-chemotherapeutic drug conjugate IR820-SS-CPT. The drug load content of IR820-SS-CPT in nanoparticles formed by self-assembly was close to 100%, and the water solubility of CPT and the membrane permeability of IR820 were significantly higher than that of a single drug [113]. Du and co-workers designed a self-assembled vector-free nanomachine (IR820/ATO NPs) based on ATovaquone (ATO) and IR820, which successfully addressed the problem of hydrophilic IR820 being unstable and easily removed in vivo. More interestingly, ATO can act as an oxidative phosphorylation system (OXPHOS) inhibitor to inhibit mitochondrial respiration and reduce the synthesis of ATP after entering tumor cells, thus leading to the downregulation of HSPs and synergistic enhancement of the sensitivity of PTT treatment in tumor cells [114].

2.4.3. Photosensitizer-Drug Conjugates

PDT is another important type of phototherapy, during which the photosensitizers (PSs) in target lesions are stimulated by light sources at a specific wavelength to produce reactive oxygen species (ROS), finally resulting in cell apoptosis and necrosis [97,115,116,117]. Upon light irradiation, PSs can be excited from the singlet basic energy state S0 to the excited singlet state S1 first, followed by partially transforming to the long-lived excited triplet state T1 through the intersystem crossing, which is the therapeutic form of the PSs. Subsequently, there are two mechanisms for ROS generation at the present stage: in the type I pathway, the excited PSs participate in the electron transfer reaction to produce free radicals and free radical ions. In the type II pathway, the PSs transfer energy to molecular oxygen (3O2) to produce highly reactive singlet oxygen (1O2). It needs to be mentioned that type II-based PDT is more popular than type I-based PDT [118]. Thereafter, the generated ROS can trigger oxidative stress on tumor cells, leading to the activation of the protein kinase pathway, expression of transcription factors and cytokines, and release of factors mediating apoptosis, resulting in the apoptosis or necrosis of tumor cells. Furthermore, PDT can effectively target tumor blood vessels, causing damage to the tumor vasculature, resulting in the injury of vascular endothelial cells, disorders of endothelial structure, and a significant reduction in the tumor cell nutrition supply. PDT has also been reported to induce acute local and systemic inflammatory responses, ultimately stimulating T cell activation and generating anti-tumor immune responses [119,120,121,122].
As for PSs, they can be divided into non-porphyrin PSs (e.g., rose red (RB) and methylene blue (MB), ruthenium (II) complexes, fullerenes, etc.) and porphyrin PSs (such as porphyrin, phthalocyanine (Pc), naphthalene phthalocyanine (Nc), etc.) [105]. In recent years, some photosensitive compounds extracted from herbal plants, such as chlorophyll and curcumin, have proven to possess photodynamic activity, which can also be used as PSs in PDT. Compared with other routinely used PSs, natural PSs usually causes lower side effects. Despite the success of PDT, new PSs and innovative methods are still needed to improve the practical application of PDT in clinical oncology. More and more studies have reported that the combination of PSs and nanomaterials, such as photosensitizer-drug conjugates-derived nanoparticles and PS-antibody conjugates [123], can improve the efficiency of PDT and diminish the undesirable side effects. Hao and co-workers formed CPT-TK-HPPH nanoparticles by a photosensitizer-drug conjugate, that combined CPT with PS 2-(1-hexoxyethyl)-2-dvinyl pyropheophorbide-a (HPPH) via a ROS-responsive thioketal bond. The platinum was then loaded into CPT-TK-HPPH to produce CPT-TK-HPPH/PT nanoparticles. The resultant nanoparticles can efficiently catalyze hydrogen peroxide (H2O2) to produce oxygen (O2), realizing tumor oxygenation for improved HPPH-mediated PDT. Moreover, the generated ROS could further induce the cleavage of the thioketal linker for the release of CPT on demand. The CPT–TK–HPPH/PT NP effectively inhibited colon tumor proliferation and growth in vitro and in vivo [124]. In another example, Ha and co-workers combined the chemo-drug combretastatin A-4 (CA4) with a tumor-targeting biotin portion and a PS Zn (II) phthalocyanine (ZnPc), in which a ROS-sensitive aminoacrylate linker was introduced for the controlled release of CA4 during PDT [125]. Besides ROS-responsive linkers, Um and co-workers used caspase 3 cleavable peptide (Asp-Glu-Val-Asp, DEVD) as a linker for the conjugation of Ce6 (PS) and MMAE (anti-cancer drug) and developed the Ce6-DEVD-MMAE nanoparticles. Compared with traditional PDT using high-energy irradiation, the new therapeutic strategy used lower-energy irradiation to induce the apoptosis of cancer cells. Along with MMAE-mediated anticancer activity, strong cytotoxic effects could be induced upon exposure to lower-energy irradiation. More importantly, Ce6–DEVD–MMAE nanoparticles did not display any toxicity in the absence of light illumination due to the drug conjugation strategy, which was different from free MMAE (1–10 nM) which had obvious cytotoxicity (Figure 5D,E) [126].
Additionally, recombinant antibody fragments have also been conjugated with PSs to realize antibody-directed PDT, of which the structure is similar to ADCs. These antibody–PS conjugates present promising strengths in superior drug loading, more favorable pharmacokinetics, enhanced potency and target cell selectivity [123]. For example, Ebaston and co-workers successfully conjugated trastuzumab (Ab, targeting Her2 receptors) with mI2XCy(PS). Interestingly, the hydroxyl group in mI2XCy was further protected by acetyl (Ac) to quench the ROS generation and fluorescence emission of PS, with the purpose of reducing the side effects caused by existing PSs to organs due to insufficient specificity. Upon the Ac group being cleaved by the intracellular esterases, the photodynamic activity could be restored and effective ROS generation could be observed upon NIR light irradiation. As desired, these Ab-mI2XCy-Ac conjugates displayed negligible side effects and promising tumor growth inhibition in the Her2 positive BT-474 tumor mouse model, which is almost the same as for the permanently active antibody–PS conjugates (Ab-mI2XCy) [127].

2.5. Inorganic Nanoparticle-Drug Conjugates

Among the various categories of nanocarriers, inorganic material-based ones are a matter of huge interest in developing nanoplatforms for cancer diagnosis and treatment, due to their excellent physical and chemical properties in the aspect of magnetic, thermal, optical, and catalytic performance. Moreover, these inorganic nanoplatforms are able to encapsulate and release drugs in a controllable manner, extend the systemic circulation, decrease the undesirable side effects, and improve biocompatibility and pharmacological profiles. Broadly, the inorganic nanoparticles could be divided into two different categories, which are metallic nanoparticles and nonmetallic nanoparticles, respectively [126].
Bulk metal-related electronic sensors have been widely applied in the field of disease diagnosis. Interestingly, the post-processing of metals into nanoparticles endows them with special features, rendering them good candidates for biomedical application. The common metallic nanostructures comprise gold nanoparticles, silver nanoparticles, iron oxide nanoparticles, calcium nanoparticles and so on. They are capable of both active and passive targeting, finally obtaining promising flexibility. Furthermore, the metallic nanoparticles themselves also display therapeutic functions during cancer treatment. For example, gold nanoparticles (AuNPs), which could be subdivided into gold nanospheres, gold nanorods, gold nanoshells, gold nanoclusters, etc., have shown potential applications in PTT and radiofrequency ablation [128]. Silver nanoparticles (AgNPs) present several intrinsic anticancer functions, including halting the generation of ROS by influencing the cellular mitochondrial system, reducing the velocity of ATP synthesis, and altering the corresponding biological pathways necessary for the survival of tumor cells [129]. Iron nanoparticles, such as iron oxide nanoparticles (IONPs) and superparamagnetic iron oxide nanoparticles (SPIONs), are indispensable conditions to realize magnetic hypothermia treatment (MHT). More importantly, the drugs can be covalently connected to these metallic nanoparticles through the chemical reaction between different functional groups, such as thiol/sulfide, carboxylic/amine, azide/alkyne and so on. Therefore, the therapeutic index and the pharmacokinetics of drugs can be significantly improved compared with native drugs [130,131,132]. For example, AgNPs were chemically conjugated with the somatostatin analog octreotide (OCT) through amide bonds to form AgNPs–OCT by Abdellatif and co-workers, which was further combined with alginate to produce AgNPs–OCT–Alg. It was found that AgNPs–OCT–Alg can not only effectively accumulate in lung tissue with promising delivery applicability and cytotoxicity, but also reduce drug side effects [133]. Gold nanorods (AuNRs) are good candidates as PTAs for PTT due to two different kinds of surface plasmon resonance (SPR), in which the strong longitudinal mode can be adjusted to visible light and the NIR region [134,135,136]. Jongseon and co-workers prepared AuNRs with different aspect ratios as PTAs to conjugate folic acid (FA)–PEG monoblock copolymer (FAP) and pheophorbide a (Pheo), yielding nanoplatform denoted as FAPAuNR–Pheos. The nanoplatform exhibited excellent performance in singlet oxygen generation, photothermal conversion, and glutathione(GSH)-responsive release of Pheo. Furthermore, FAPAuNR–Pheo with tumor-targeting FA ligands exhibit promising tumor-targeting activity and a synergistic PTT/PDT effect (Figure 6) [137].
Nonmetallic nanoparticles commonly refer to nanostructures based on silicon and carbon materials. These nanoparticles could be mesoporous, represented by the mesoporous silica nanoparticles (MSNs), which offer them promising properties, such as high drug loading capacity and pleasant surface chemical modification. These characteristics promote a great potential for smart drug delivery. MSNs are traversed by pores with a nanometer scale and are usually synthesized through a top-down approach with the assistance of chemicals and metals [138]. Due to the special porous structures, they present high drug loading ability and diverse surface functionalization. Moreover, the particle sizes, pore sizes and particle shape are adjustable by changing the precursor (e.g., tetraethyl orthosilicate (TEOS)) concentrations and stirring conditions. In order to increase the drug loading content and loading stability, the drugs can also be retained in the silica network through chemical conjugation (on the surface of MSNs or inside the pores) and released in a controllable manner. In addition, MSNs are also biodegradable. Their decomposition products, orthosilicic acid, are harmless to the health, which is one of its advantages over other inorganic nanoparticles [138,139]. However, exposed monodisperse silica microspheres tend to exhibit high aggregation when directly exposed to the biological environment, limiting their application in the biomedical field which could be overcome by the proper surface modification on mesoporous silica. Peng and co-workers designed a nanoplatform based on MSN for synergistic PTT/chemotherapy. The polymer poly(PEGMA-co-HEMA) was firstly modified onto the surface of MSNs via surface-initiated atom transfer radical polymerization. Then, DOX was anchored onto the polymer via reversible covalent bond cis-aconitic anhydride with pH sensitivity. Indocyanine green (ICG), as PTA, was further loaded into the pores of MSN for PTT, obtaining MSN@poly(PEGMA-co-HEMA-g-DOX)/ICG. This nanoplatform presents an improved synergistic effect on both HepG2 and Hela cells by virtue of photothermal action and promoted linker cleavage [140].
Carbon nanoparticles applied in drug delivery are usually sp2 carbon materials, including single-walled carbon nanotubes (SWCNTs) and multiwalled carbon nanotubes (MWCNTs), graphene, fullerene and carbon dots. They possess unparallel advantages, such as easy surface modification, strong adsorption, high photothermal conversion efficiency, supramolecular π–π stacking and excellent biocompatibility. Actually, therapeutics could also be covalently functionalized onto the carbon nanoparticles, since the carbon materials could bear the carboxylic acid groups once undergoing treatment with strong acid solutions, which provides a reaction site for the drug conjugation. Besides, carbon materials could also be functionalized via 1,3-dipolar cycloaddition of azomethine ylides, which provide carbon nanoparticles with customized substituents relying on the structure of α-amino acids and aldehydes employed during modification [141]. Dhar and co-workers prepared an axial folate derivative (FA)-containing platinum(IV) complex and further conjugated it with SWCNTs. This platinum-single-walled carbon nanotube structure containing folic acid has the targeting ability, which increases the activity of the platinum-based anticancer drug and significantly enhances the cell killing performance [142]. Recently, quantum dots (QDs) and ceramic-based nanoparticles have been paid more attention to due to their anticancer potential [143,144].

3. Representative Applications of Conjugated Nanomedicine

Chemotherapy is one of the standard methods for the clinical treatment of malignant tumors. Due to the heterogeneity of tumors and the complexity of their pathological mechanisms, a single chemotherapeutic drug is usually unable to eradicate cancer cells. It may also encounter some problems, such as toxic side effects induced by high doses of drugs and obtaining MDR after repeated treatment. These problems then increase the likelihood of cancer metastasis or recurrence [145,146,147]. The emergence of the combination of multiple antineoplastic drugs makes up for the deficiency of single drug application. Accordingly, the overall treatment benefit of the multidrug combination is usually higher than that of single drug administration by virtue of different therapeutic mechanisms. More importantly, the drug dose used during synergistic therapy usually decreases and the unfavorable side effects could be weakened under the premise of the same or better therapeutic efficacy [148].
Nevertheless, the traditional “cocktail” therapeutic strategies also display limitations. The actual concentration of individual drugs in tumors is uncontrollable, mainly due to differences between drugs in physical and chemical properties, pharmacokinetics, tissue distribution and so on. Since the drug ratios play a significant role in their synergistic manner, the final therapeutic outcome cannot be guaranteed during “cocktail” treatment. In contrast, conjugated nanomedicine usually presents a relatively fixed drug ratio in blood and tumor tissues depending on the initial composition of the materials. The stable drug connection also prevents the undesired pre-mature drug release, further promoting the accurate delivery of multiple drugs to targets as well as promising therapeutic efficacy. In this section, we will introduce some representative applications of conjugated nanomedicine in synergistic chemotherapy.

3.1. Synergistic Chemo-Chemo Therapy

Different types of chemotherapeutic drugs can be classified according to their function and mechanism of action on cancer cells. The most common classes of chemotherapeutic drugs are alkylating agents, antimetabolic agents, anthracyclines, topoisomerase inhibitors, mitotic inhibitors, and corticosteroids. The choice of two or more chemotherapeutic drugs depends on the stage and type of cancer, the synergistic behaviors of various drugs and other factors. The choice of drugs determines whether the effect is synergistic, additive, or antagonistic [149]. Conventional “cocktail” treatment generally presents an inadequate enhancement of therapeutic efficiency since different anticancer drugs often display diverse pharmacokinetics and transmembrane ability, resulting in the uncontrollable distribution of drugs in tumors [150,151,152]. To overcome the above challenges, nano-carrier based multiple drugs co-delivery systems have been developed for cancer synergistic chemo–chemo therapy, which are capable of delivering two or more chemotherapeutics to tumors in a desirable ratio by virtue of the advantages of nanomedicine. Once connected, the hydrophobic and hydrophilic anti-tumor drugs, serving as building blocks, could also provide the impetus to fabricate the nanoparticles. This kind of conjugated nanomedicine not only displays the advantages of high drug loading content but also relatively stable drug delivery capacity, reduced side effects and enhanced anticancer activity benefitting from the synergistic effect of different drugs [153,154,155]. Previously, we have introduced the conjugated nanomedicine (denoted as MTX-SS-PPT NAs) developed by Hou and co-workers, which is self-assembled from the drug conjugates containing hydrophilic drug MTX and the hydrophobic drug PPT. In this nanomedicine, the disulfide bonds connecting the two drugs contribute to the degradation of MTX-SS-PPT NAs in tumor cells under reduction conditions. As a slow-release of the active drug, the nanoagent can significantly improve the biocompatibility of PPT and reduce its toxicity [98].
The development and metastasis of solid tumors highly depend on the formation of neovascularization. However, the use of angiogenesis inhibitors alone cannot meet the needs of cancer treatment. Sun and co-workers conjugated hydrophilic chemotherapeutic drugs (fluorosarboside, FUDR) with hydrophobic antiangiogenic drugs (pseudoboric acid B, PAB), followed by formulating nanoparticles in an aqueous solution. These nanoparticles not only displayed promising anti-tumor activity but also had efficient antiangiogenesis properties, leading to a good cancer therapeutic outcome in mice bearing subcutaneous HeLa tumors [156]. Dasatinib (DAS) is a competitive oral dual Src/Abl kinase inhibitor, which can inhibit a variety of Src signal pathways and further inhibit tumor cell migration, invasion and angiogenesis [157]. Yang and co-workers linked DAS with cisplatin octahedral coordination derivative diamino dichlorodihydroxyplatinum (DH-CP) through an ester bond to form an amphiphilic drug–drug conjugate (CP–DDA) at the ratio of 2:1. Then, the stable nanoparticles (CP–DDA NPs) were formed by the self-assembly of CP–DDA in an aqueous solution. The nanoparticles displayed promising stability during blood circulation and increased accumulation of drugs in the tumor site through the EPR effect. After being internalized by cancer cells, under the action of high GSH and esterase, the DAS and CP could be released in situ for inhibiting Src activity and inducing cell apoptosis, respectively, resulting in a synergistic anti-tumor effect [158].

3.2. Synergistic PDT/PTT-Chemo Therapy

Phototherapy, including PDT and PTT, is a tumor resection and function-preserving interventional therapy, which shows great potential in clinical application. In the process of phototherapy, non-toxic phototherapeutic agents (PSs or PTAs) can be activated upon light irradiation, thus inducing cell death without causing undesirable collateral damage to normal tissue. However, it is difficult to completely eradicate solid tumors with single phototherapy. It has been reported that the combination of PTT and/or PDT with chemotherapy can provide therapeutic advantages including (1) giving play to synergistic effects during treatment. (2) Decreasing the undesirable side effects from anticancer drugs via lowering the drug dosage. (3) Facilitating the deep tumor penetration of chemotherapeutic drugs under hyperthermia treatment (PTT). (4) Promoting the cellular internalization of drugs in the presence of a large amount of ROS (PDT). (5) Inducing an immune response by phototherapy, including innate/adaptive immunity and antitumor immunity, to maximize the therapeutic outcome [105,159]. The unfavorable pharmacokinetic properties and desynchrony in the tumor accumulation of chemotherapeutic drugs and phototherapeutic agents still hinder the success of synergistic PDT/PTT-chemotherapy. Therefore, phototherapeutic nanomedicine has also aroused great interest in order to continuously improve its performance [97,160], among which conjugated nanomedicine occupies an important position [161,162]. As for synergistic PDT-chemotherapy, for example, Chen and co-workers developed a supramolecular system with optimized PS (4,4-difluoro-boradiazaindacene, BODIPY) and anti-cancer drug (PTX) loading efficiency. The adamantyl BODIPY (Ada-BODIPY) and PTX (Ada-PTX) were connected to the block copolymer (PEG-PGA-β-CD) through the host–guest interaction between adamantane and β-cyclodextrins (β-CD), followed by being prepared into nanoparticles of Ada-PTX (60%)-BODIPY(40%)-PNS. These nanoparticles remained in the precise drug loading ratio during circulation with the minimized pre-mature release of drugs. Upon NIR laser irradiation, ROS could be generated efficiently for PDT as well as the cleavage of ROS-sensitive aminoacrylate linker in Ada-PTX. Finally, PDT and cascaded Ada-PTX activation showed a significant inhibitory effect on tumor growth [163]. A similar nanosystem (PheoA-SN38-HC) has been developed by Lee and co-workers, which contains the ROS-cleavable thioketal-SN38 for the drug release during PDT, showing good tumor targeting for CD44 positive cancer cells and effective tumor inhibition mediated by synergistic PDT-chemotherapy (Figure 7A–C) [164].
Hypoxia is a key feature of the solid tumor microenvironment (TME) resulting from rapid malignant cell proliferation and vascular deformation during tumor angiogenesis, which is not beneficial to PDT. To overcome this problem, Xu and co-workers designed a nanoplatform self-assembled from amphiphilic oligomer Ce6-PEG Platinum(IV) (Ce6-PEG-Pt(IV), CPP) with upconversion nanoparticles (UCNPs) in the hydrophobic core. In this system, platinum(IV) diazido complexes bearing cis-diamine ligands can be activated to produce O2 as well as cytotoxic Pt(II) simultaneously upon laser illumination, successfully compensating for the consumption of O2 during the PDT process. The released active platinum(II) could also trigger efficient chemotherapeutic effects, resulting in dramatically enhanced synergistic PDT-chemotherapy [165].
As for synergistic PTT-chemotherapy, Li and co-workers designed a prodrug–hemicyanine conjugate (Cy-azo) based nanoplatform to achieve the combination of H-aggregation-improved PTT and sequential hypoxia-activated chemotherapy. In Cy-azo, nitrogen mustard was introduced into the NIR fluorescent group heptamethyl cyanamide through an azo bond, and the superposition of the conjugates promoted H aggregates, showing higher photothermal conversion efficiency than traditional cyanine dyes. In addition, under hypoxic conditions, the nitrogen mustard can be activated due to the cleavage of azo bonds and released in the hypoxic TME to induce cell death, thereby greatly reducing the side effects of chemotherapy [166]. In another work, Zhou and co-workers prepared the dual drugs-conjugated polydopamine nanoparticles (PDOXCBs) through the one-pot aqueous copolymerization of two dopamine prodrugs, which combined the NIR-mediated PTT with cocktail chemotherapy into one nanoplatform. Upon NIR irradiation, PDOXCBs presented a dramatic photothermal effect with the assistance of polydopamine nanoparticles as PTAs. Meanwhile, chemotherapeutic drugs, including DOX and chlorambucil (CB), could be released from the nanoplatforms under the stimuli of pH 5.0 and the reduced environment, respectively. The synergistic PTT-chemotherapy based on PDOXCB27 upon NIR irradiation displayed a highly lowered IC50 value on MCF-7 cells and a combination index of 0.36, revealing a promising combination between PTT and cocktail chemotherapy (Figure 7D,E) [167].

3.3. Synergistic Immune-Chemo Therapy

In recent years, with the development of molecular biology and tumor biology, tumor immunotherapy has become a new treatment method with a good application prospect. During cancer immunotherapy, the collective immune system can be activated by strengthening the natural immune defense of patients to fight against cancer cells and relieving the immunosuppressive microenvironment, to eradicate tumors and inhibit tumor metastasis and recovery. On the one hand, immunotherapy is aimed at training immune cells to recognize and remove target cells carrying tumor antigens, and enhancing immune-mediated tumor cell lysis, displaying the advantages of good curative effect, fewer adverse reactions and the prevention of recurrence. On the other hand, the down-regulation of the immunosuppressive signal pathways in tumor tissues could also facilitate the final immunotherapeutic effects [168]. So far, there have been several types of immunotherapy achieving great success in tumor therapy, such as immune checkpoint blockade (ICB), adoptive T cell transfer, cytokine therapy, agonist immunotherapy, vaccines and so on. However, due to the complexity and heterogeneity of tumors, systemic defects, such as immune escape and immunotoxicity of tumors make the overall efficacy of immunotherapy only about 20%. It needs to further improve the efficiency of tumor immunotherapy via inhibiting immune escape and enhancing the immunotherapeutic response rate. Among the various therapeutic strategies to improve the efficacy of immunotherapy, drug conjugates have been recognized as one of the good choices. Nano-drug delivery systems can enhance the retention, accumulation, penetration and target cell uptake of tumor immunotherapeutic drugs in tumor sites [169,170].
Besides, the combination of immunotherapy with other therapeutic modalities, such as chemotherapy and phototherapy, increases the immunotherapeutic effects. It has been recognized that chemotherapy can induce immunogenic cell death (ICD) to express or release damage-associated molecular patterns (DAMPs), including calreticulin (CRT), high-mobility group box 1 (HMGB-1), and adenosine triphosphate (ATP). These DAMPs are capable of enhancing the immunogenicity of cancer cells and stimulating the immune system to fight against tumors. Given this, Geng and co-workers developed the aptamer-drug conjugate nanomicelles to facilitate the antitumor immune response via DOX-mediated chemotherapy. In detail, an amphiphilic telodendrimer (ApMDC) consisting of an aptamer AS1411 and a monodendron connected with four DOX through acid-labile hydrazone spacers was firstly synthesized, followed by co-self-assembly with an ApMDC analog, in which the aptamer is substituted by PEG. Based on their results, the optimized micelles could induce ICD. Besides, the chemotherapy also promoted the tumor-specific immune responses of anti-PD-1 therapy (Figure 8A,B) [171]. In another work, Hu and co-workers developed a ROS-sensitive nanosystem (denoted as pep-PAPM@PTX) for synergistic chemotherapy and ICB therapy. The PD-L1-targeting D-peptide (NYSKPTDRQYHF, pep) was conjugated to the carrier materials and exposed to the surface of micelles with the function of anti-PD-L1 therapy. Accordingly, this micelle could bind to PD-L1 on the cell surface and promote its uptake via the lysosome-involved internalization, thus inhibiting PTX-activated PD-L1 upregulation and downregulating PD-L1 expression. It dramatically facilitated T cell infiltration and enhanced tumor immune activation, finally synergizing with PTX-mediated chemotherapy to achieve promising anticancer effects against triple-negative breast cancer (TNBC) [172]. Bai and co-workers designed a GSH/pH dual-responsive prodrug nano-platform (known as DDA) for synergistic chemotherapy/PDT/immunotherapy. The nano-platform can effectively enhance the immune response by promoting the maturation of dendritic cells and reducing the number of immunosuppressive immune cells, showing the enhanced adjuvant effect of anti-PD-1 therapy [173].
In addition to be applied for synergistic immune-chemo therapy, conjugated nanomedicine serves as a promising tool, and can also enhance the therapeutic efficacy of immunotherapy alone or other forms of synergistic immunotherapy, which is necessary to be discussed in this section. The company named Cyrtlmmune Sciences has developed a conjugated nanomedicine-related antitumor drug CYT-6091 (trade name: AurimuneTM) for cytokine immunotherapy. As mentioned in the previous section, gold nanoparticles can serve as PTAs and drug carriers simultaneously. The CYT-6091 was synthesized by covalent binding with recombinant human tumor necrosis factor (rhTNF) onto colloidal gold nanoparticles coated with mercapto functionalized PEG. It can be specifically stored in tumor tissues and has no obvious toxic and side effects [174]. It also provided the potential to combine with gold nanoparticles-mediated thermal therapy.
More recently, Xue and co-workers reported that the CD73 enzyme was highly expressed in tumor cells and immunosuppressive cells, including regulatory T cells (Treg cells), myeloid suppressor cells (MDSCs) and M2-like tumor-associated macrophages (TAM.M2). However, CD73 was negative for non-immunosuppressive cells, known as lytic T lymphocytes, natural killer cells (NK cells) and dendritic cells (DC cells). Based on these findings, they developed an IR-700 dye-coupled anti-CD73 antibody (α-CD73-Dye), which could bind to CD73+ cells selectively. Upon NIR laser irradiation, these conjugates could perform photoimmunotherapy against targeted cells and prevent tumors from acquiring resistance to ICB, finally leading to advanced tumor eradication [175].
Another effective cancer immunotherapeutic modality is chimeric antigen receptor T-cell immunotherapy (CAR-T therapy), which is an adoptive T cell metastasis therapy (adaptive T cell metastasis, ACT), which infuses the patient’s T cells back into the patient to fight cancer. Compared with ordinary T cells, CAR-T is not restricted by the major histocompatibility complex (MHC), thus avoiding the immune escape of tumor cells with the low expression of MHC molecules on their surface. However, the immunosuppressive tumor microenvironment inhibits the infiltration of T cells, limiting the effect of CAR-T therapy. Luo and co-workers combined human serum albumin (€) and IL-12 into nanoparticles, which were further modified onto the surface of CAR-T cells via bioorthogonal chemistry to yield IL-12 nonstimulant engineering CAR T cells (INS-CAR T) hybrids. The IL-12 released from nanoplatforms can promote the secretion of CCL5, CCL2 and CXCL10, thus increasing the infiltration of CD8+ CAR T cells, relieving the immunosuppressive TME. Based on their results, the anti-tumor ability of CAR-T cells has been improved and the growth of solid tumors was inhibited with negligible side effects (Figure 8C–E) [176].

3.4. Synergistic PTT-TDT

As the prominent character of solid tumors, hypoxia impedes the therapeutic effect of oxygen-dependent radical-based cancer therapy, such as PDT and radiotherapy. To address these hypoxia issues, researchers have developed another type of oxygen-independent radical-based cancer treatment strategy, known as TDT. During TDT, the alkyl radicals can be produced upon heating with high efficiency due to the presence of thermally decomposable radical initiators, such serving as radical donors. More importantly, the aforementioned PTT could also serve as a heat source to induce the generation of alkyl radicals. Upon NIR light irradiation, light-triggered heat and heat-caused alkyl radicals can jointly damage vital cellular components and further induce cell death [177,178]. To improve the accumulation of radical initiators and PTAs in tumors and avoid the undesired pre-mature release of them during blood circulation, the PTA-initiator conjugated nanomedicine has also been developed for enhanced synergistic PTT-TDT. Xia and co-workers conjugated the photothermal PSs (porphyrin) with radical initiator 2,2′-azobis [2-(2-imidazolinI-2-yl) propane dihydrochloride (AIBI) and prepared the nanoparticles (tripolyphosphate (TPP)-NN NPs) with the assistance of pluronic F-127 as surfactant. The aggregated porphyrin could generate heat upon 638 nm laser illumination and then trigger initiator AIBI to produce alkyl radicals to induce cell death even in a hypoxia environment. TPP-NN NPs have shown the potential to inhibit the growth of cervical tumors without notable systemic toxicity [179].
In our previous study, an all-organic nanoparticle (denoted as ZPA@HA-ACVA-AZ NBs) realized the “precise strike” of hypoxic tumors via synergistic PTT/TDT. The loading strategy of radical initiators (ACVA) was optimized by the conjugation of alkyl chain-functionalized initiators ACVA-HDA to HA, thus averting the unfavorable adverse effect in normal tissues while improving the efficiency for the targeted delivery of radical initiators to solid tumors. Then, this amphiphilic hyaluronic acid (HA)-based lipoid (HA-ACVA-AZ) was used as a carrier to encapsulate the special zinc(II) phthalocyanine aggregates (ZPA), acting as PTAs for highly efficient PTT upon 808 nm laser irradiation. Therefore, the sequentially generated heat and alkyl radicals could simultaneously trigger cell death and restrain cancer metastasis under the action of PTT/TDT and CA IX inhibition [180]. Our group also developed carrier-free nano-theranostic agents (denoted as AIBME@IR780-APM NPS) for magnetic resonance imaging (MRI)-guided synergistic PTT/TDT. As an extension of previous work, we were devoted to the incorporation of diagnostic functions into the nanoplatform to improve the accuracy of synergistic therapy. As for the subject of this nanomedicine, the first IR780 derivative, IR780-ATU, was designed to conjugate the chelating agents (acylthiourea, ATU) with PTA with the purpose of chelating transition metal Mn2+ ions to perform the T1-weighted contrast-enhanced MRI. The other derivative, IR780-PEG, renders a nanosystem with high sterical stability, and increased solubility of hydrophobic IR780/dimethyl 2,2′-azobis(2-methylpropionate) (AIBME, radical initiator) and reduced risk from reticuloendothelial system (RES) uptake. Upon IR780-mediated PTT launched under NIR laser irradiation, AIBME could generate highly cytotoxic alkyl radicals, combing the heat from PTT to synergistically induce cell death, ignoring tumor hypoxia [181].

4. Challenges and Future Perspectives of Bioconjugation and Nanomedicine

Although conjugated nanomedicines have made significant progress in the treatment of tumors, they have shown great potential and application prospects, especially in protecting the loaded components, increasing drug selective accumulation and intratumoral penetration in tumor tissues, and reducing serious side effects in normal tissues. However, there are still many factors influencing the therapeutic efficiency and further clinical application of conjugated nanomedicine, such as the complexity of tumor biology, the biological interaction between the nanomedicine and the in vivo biological substances, and the large-scale production.
At present, more than 200 nanomedicines are in clinical research. Based on the reported data, although 95% of nanomedicines could pass the phase I clinical test, only 48% and 14% of these nanomedicines achieved good performance in phase II and III clinical tests. The major reasons causing the clinical failure of these nanomedicines are concluded as follows: 1. Biological barrier. Nanomedicines enter the body through a complex biological process to exert their efficacy. For example, before exerting their anticancer effects, nanomedicine should circulate in the body, accumulate in the tumor, penetrate into the deeper region of the tumor, internalize into tumor tissue and release the drugs. Every step in this process plays an essential role in achieving the desired therapeutic efficiency of nanomedicine. Thus, how to overcome these biological barriers during these steps critically influences their efficiency. 2. The linker in the conjugated nanomedicine significantly influences the targeted- and controlled-release of therapeutics. Thus, more intelligent and sensitive linkers should be explored and applied to achieve the “precise attack” of tumors. 3. Large-scale production. Large-scale production is also one of the major obstacles for nanomedicine transferring from bench to bedside. The process of large-scale can affect the physical and chemical characteristics of nanomedicine, such as the particle size, surface properties and drug loading capacity, which are vital for the biological effects of nanoparticles. Moreover, the toxic solvent usage during large-scale preparation may also influence the stability of materials or antibodies in conjugated nanomedicine. 4. Non-selective distribution. Although the targeted nanomedicine can accumulate in tumor tissues through the EPR effect and active targeting capacity in animal models, only 0.7% of the injected nanomedicine can actually be detected in tumors in the human body. Thus, it is still necessary to further explore the new mechanism for efficient targeted drug delivery. Thus, more specific antibodies and targeting strategies should be established to improve the selective distribution of conjugated nanomedicine. 5. The correlation between preclinical research and clinical trial results is not strong. Because the therapeutic efficacy of nanomedicine depends on its pharmacokinetics, tissue distribution, tumor accumulation, penetration, drug release, etc., the above characteristics are very different between animal models and patients. Furthermore, there is huge heterogeneity among different patients and tumor types. Therefore, improving the clinical therapeutic efficacy and comprehensive benefits is the key to promoting the clinical transformation of anti-tumor nanomedicines.
Lessons learned from completed, ongoing, or terminated clinical trials can help guide the forward development of nanomedicine. Clinical cancer treatment plans should refer to the cancerous tumors’ size, stage, location, and grade, thus the new generation of conjugated nanomedicines should be optimized and standardized in many aspects of clinical research, including patient screening, proper drug selection and proper combination with other therapies to further accelerate the development of conjugated nanomedicines.
Overall, through changing the strategies of many therapeutics’ applications, conjugated nanomedicine as one major branch of nanomedicine has further improved cancer treatment and developed many potential applications against cancer. With the deeper exploration of tumor pathogenesis, pharmacology, nanomedicine, etc., more intelligent and effective conjugated nanomedicines would be developed and transferred from the laboratory to the clinical application, finally benefiting more cancer patients and achieving victory over cancer in the near future.

Funding

This work was supported by the National Natural Science Foundation of China (51903203, 51703178, 81901808), the China Postdoctoral Science Foundation (2019M653661, 2019M663742), the Natural Science Foundation of Zhejiang Province (LWY20H180002), the Natural Science Foundation of Shaanxi Province (2022JM-183), Shaanxi Provincial Key R&D Program (No.: 2022SF-342), the Fundamental Research Funds for the Central Universities (xpt012022030, xzy012022037).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

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. Chen, Q.; Ke, H.; Dai, Z.; Liu, Z. Nanoscale theranostics for physical stimulus-responsive cancer therapies. Biomaterials 2015, 73, 214–230. [Google Scholar] [CrossRef] [PubMed]
  3. Wong, C.; Siah, K.W.; Lo, A.W. Estimation of clinical trial success rates and related parameters. Biostatistics 2019, 20, 273–286. [Google Scholar] [CrossRef] [PubMed]
  4. Dong, P.; Rakesh, K.P.; Manukumar, H.M.; Mohammed, Y.H.E.; Karthik, C.S.; Sumathi, S.; Mallu, P.; Qin, H.L. Innovative nano-carriers in anticancer drug delivery-a comprehensive review. Bioorg. Chem. 2019, 85, 325–336. [Google Scholar] [CrossRef]
  5. Shen, J.; Wolfram, J.; Ferrari, M.; Shen, H. Taking the vehicle out of drug delivery. Mater. Today 2017, 20, 95–97. [Google Scholar] [CrossRef] [Green Version]
  6. Rajitha, B.; Malla, R.R.; Vadde, R.; Kasa, P.; Prasad, G.L.V.; Farran, B.; Kumari, S.; Pavitra, E.; Kamal, M.A.; Raju, G.S.R.; et al. Horizons of nanotechnology applications in female specific cancers. Semin. Cancer Biol. 2021, 69, 376–390. [Google Scholar] [CrossRef]
  7. Raju, G.S.R.; Dariya, B.; Mungamuri, S.K.; Chalikonda, G.; Kang, S.M.; Khan, I.N.; Sushma, P.S.; Nagaraju, G.P.; Pavitra, E.; Han, Y.K. Nanomaterials multifunctional behavior for enlightened cancer therapeutics. Semin. Cancer Biol. 2021, 69, 178–189. [Google Scholar] [CrossRef]
  8. de Lázaro, I.; Mooney, D.J. Obstacles and opportunities in a forward vision for cancer nanomedicine. Nat. Mater. 2021, 20, 1469–1479. [Google Scholar] [CrossRef]
  9. Lammers, T.; Kiessling, F.; Ashford, M.; Hennink, W.; Crommelin, D.; Storm, G. Cancer nanomedicine: Is targeting our target? Nat. Rev. Mater. 2016, 1, 16069. [Google Scholar] [CrossRef] [Green Version]
  10. Bobo, D.; Robinson, K.J.; Islam, J.; Thurecht, K.J.; Corrie, S.R. Nanoparticle-based medicines: A review of fda-approved materials and clinical trials to date. Pharm. Res. 2016, 33, 2373–2387. [Google Scholar] [CrossRef]
  11. Pearce, A.K.; O’Reilly, R.K. Insights into active targeting of nanoparticles in drug delivery: Advances in clinical studies and design considerations for cancer nanomedicine. Bioconj. Chem. 2019, 30, 2300–2311. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, M.T.; Chen, Y.; Marhoul, J.; Jacobson, F. Statistical modeling of the drug load distribution on trastuzumab emtansine (kadcyla), a lysine-linked antibody drug conjugate. Bioconj. Chem. 2014, 25, 1223–1232. [Google Scholar] [CrossRef]
  13. Barenholz, Y. Doxil®—The first FDA-approved nano-drug: Lessons learned. J. Control. Release 2012, 160, 117–134. [Google Scholar] [CrossRef] [PubMed]
  14. Cabral, H.; Matsumoto, Y.; Mizuno, K.; Chen, Q.; Murakami, M.; Kimura, M.; Terada, Y.; Kano, M.; Miyazono, K.; Uesaka, M. Accumulation of sub-100 nm polymeric micelles in poorly permeable tumours depends on size. Nat. Nanotechnol. 2011, 6, 815–823. [Google Scholar] [CrossRef] [PubMed]
  15. Ventola, C.L. Progress in nanomedicine: Approved and investigational nanodrugs. Pharm. Ther. 2017, 42, 742. [Google Scholar]
  16. Alven, S.; Nqoro, X.; Buyana, B.; Aderibigbe, B.A. Polymer-drug conjugate, a potential therapeutic to combat breast and lung cancer. Pharmaceutics 2020, 12, 406. [Google Scholar] [CrossRef]
  17. Jatzkewitz, H.P. Bound to blood plasma expander (polyvinylpyrrolidone) as a new depot form of a biologically active primary amine (mescaline) Z. Naturforsch 1955, 10, 27–31. [Google Scholar] [CrossRef]
  18. Murguia-Favela, L.; Min, W.; Loves, R.; Leon-Ponte, M.; Grunebaum, E. Comparison of elapegademase and pegademase in ada-deficient patients and mice. Clin. Exp. Immunol. 2020, 200, 176–184. [Google Scholar] [CrossRef]
  19. Alconcel, S.N.; Baas, A.S.; Maynard, H.D. Fda-approved poly (ethylene glycol)–protein conjugate drugs. Polym. Chem. 2011, 2, 1442–1448. [Google Scholar] [CrossRef]
  20. Amgoth, C.; Dharmapuri, G. Synthesis and characterization of polymeric nanoparticles and capsules as payload for anticancer drugs and nanomedicines. Mater. Today Proc. 2016, 3, 3833–3837. [Google Scholar] [CrossRef]
  21. De Koker, S.; Hoogenboom, R.; De Geest, B.G. Polymeric multilayer capsules for drug delivery. Chem. Soc. Rev. 2012, 41, 2867–2884. [Google Scholar] [CrossRef] [PubMed]
  22. Calzoni, E.; Cesaretti, A.; Polchi, A.; Di Michele, A.; Tancini, B.; Emiliani, C. Biocompatible polymer nanoparticles for drug delivery applications in cancer and neurodegenerative disorder therapies. J. Funct. Biomater. 2019, 10, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Mamidi, N.; Delgadillo, R.M.V. Design, fabrication and drug release potential of dual stimuli-responsive composite hydrogel nanoparticle interfaces. Colloids Surf. B Biointerfaces 2021, 204, 111819. [Google Scholar] [CrossRef] [PubMed]
  24. Mohan, A.; Girdhar, M.; Kumar, R.; Chaturvedi, H.S.; Vadhel, A.; Solanki, P.R.; Mamidi, N. Polyhydroxybutyrate-based nanocomposites for bone tissue engineering. Pharmaceuticals 2021, 14, 1163. [Google Scholar] [CrossRef]
  25. Palmerston Mendes, L.; Pan, J.; Torchilin, V. Dendrimers as nanocarriers for nucleic acid and drug delivery in cancer therapy. Molecules 2017, 22, 1401. [Google Scholar] [CrossRef] [Green Version]
  26. Kesharwani, P.; Choudhury, H.; Meher, J.G.; Pandey, M.; Gorain, B. Dendrimer-entrapped gold nanoparticles as promising nanocarriers for anticancer therapeutics and imaging. Prog. Mater. Sci. 2019, 103, 484–508. [Google Scholar] [CrossRef]
  27. Duncan, R. The dawning era of polymer therapeutics. Nat. Rev. Drug Discov. 2003, 2, 347–360. [Google Scholar] [CrossRef]
  28. Parveen, S.; Arjmand, F.; Tabassum, S. Clinical developments of antitumor polymer therapeutics. RSC Adv. 2019, 9, 24699–24721. [Google Scholar] [CrossRef] [Green Version]
  29. Jameson, G.S.; Hamm, J.T.; Weiss, G.J.; Alemany, C.; Anthony, S.; Basche, M.; Ramanathan, R.K.; Borad, M.J.; Tibes, R.; Cohn, A. A multicenter, phase i.; dose-escalation study to assess the safety, tolerability, and pharmacokinetics of etirinotecan pegol in patients with refractory solid tumorsetirinotecan pegol phase i in patients with solid tumors. Clin. Cancer Res. 2013, 19, 268–278. [Google Scholar] [CrossRef] [Green Version]
  30. Thakor, P.; Bhavana, V.; Sharma, R.; Srivastava, S.; Singh, S.B.; Mehra, N.K. Polymer–drug conjugates: Recent advances and future perspectives. Drug Discov. Today 2020, 25, 1718–1726. [Google Scholar] [CrossRef]
  31. Ekladious, I.; Colson, Y.L.; Grinstaff, M.W. Polymer–drug conjugate therapeutics: Advances, insights and prospects. Nat. Rev. Drug Discov. 2019, 18, 273–294. [Google Scholar] [CrossRef] [PubMed]
  32. Greenwald, R.B.; Choe, Y.H.; McGuire, J.; Conover, C.D. Effective drug delivery by pegylated drug conjugates. Adv. Drug Deliv. Rev. 2003, 55, 217–250. [Google Scholar] [CrossRef]
  33. Ing, M.; Gupta, N.; Teyssandier, M.; Maillère, B.; Pallardy, M.; Delignat, S.; Lacroix-Desmazes, S. Immunogenicity of long-lasting recombinant factor viii products. Cell. Immunol. 2016, 301, 40–48. [Google Scholar] [CrossRef] [Green Version]
  34. Makadia, H.K.; Siegel, S.J. Poly lactic-co-glycolic acid (plga) as biodegradable controlled drug delivery carrier. Polymers 2011, 3, 1377–1397. [Google Scholar] [CrossRef] [PubMed]
  35. Vlachopoulos, A.; Karlioti, G.; Balla, E.; Daniilidis, V.; Kalamas, T.; Stefanidou, M.; Bikiaris, N.D.; Christodoulou, E.; Koumentakou, I.; Karavas, E. Poly (lactic acid)-based microparticles for drug delivery applications: An overview of recent advances. Pharmaceutics 2022, 14, 359. [Google Scholar] [CrossRef] [PubMed]
  36. Espinoza, S.M.; Patil, H.I.; San Martin Martinez, E.; Casañas Pimentel, R.; Ige, P.P. Poly-ε-caprolactone (pcl), a promising polymer for pharmaceutical and biomedical applications: Focus on nanomedicine in cancer. Int. J. Polym. Mater. Polym. Biomater. 2020, 69, 85–126. [Google Scholar] [CrossRef]
  37. Tinajero-Díaz, E.; de Ilarduya, A.M.; Cavanagh, B.; Heise, A.; Muñoz-Guerra, S. Poly (amino acid)-grafted polymacrolactones. Synthesis, self-assembling and ionic coupling properties. React. Funct. Polym. 2019, 143, 104316. [Google Scholar] [CrossRef]
  38. Sun, H.; Chang, M.Y.Z.; Cheng, W.-I.; Wang, Q.; Commisso, A.; Capeling, M.; Wu, Y.; Cheng, C. Biodegradable zwitterionic sulfobetaine polymer and its conjugate with paclitaxel for sustained drug delivery. Acta Biomater. 2017, 64, 290–300. [Google Scholar] [CrossRef] [PubMed]
  39. Feng, X.; Xu, W.; Liu, J.; Li, D.; Li, G.; Ding, J.; Chen, X. Polypeptide nanoformulation-induced immunogenic cell death and remission of immunosuppression for enhanced chemoimmunotherapy. Sci. Bull. 2021, 66, 362–373. [Google Scholar] [CrossRef]
  40. Ma, S.; Song, W.; Xu, Y.; Si, X.; Zhang, D.; Lv, S.; Yang, C.; Ma, L.; Tang, Z.; Chen, X. Neutralizing tumor-promoting inflammation with polypeptide-dexamethasone conjugate for microenvironment modulation and colorectal cancer therapy. Biomaterials 2020, 232, 119676. [Google Scholar] [CrossRef]
  41. do Pazo, C.; Nawaz, K.; Webster, R.M. The oncology market for antibody-drug conjugates. Nat. Rev. Drug Discov. 2021, 20, 583–584. [Google Scholar] [CrossRef] [PubMed]
  42. Woodruff, M.A.; Hutmacher, D.W. The return of a forgotten polymer—polycaprolactone in the 21st century. Prog. Polym. Sci. 2010, 35, 1217–1256. [Google Scholar] [CrossRef] [Green Version]
  43. Danhier, F.; Ansorena, E.; Silva, J.M.; Coco, R.; Le Breton, A.; Préat, V. Plga-based nanoparticles: An overview of biomedical applications. J. Control. Release 2012, 161, 505–522. [Google Scholar] [CrossRef] [PubMed]
  44. Nampoothiri, K.M.; Nair, N.R.; John, R.P. An overview of the recent developments in polylactide (pla) research. Bioresour. Technol. 2010, 101, 8493–8501. [Google Scholar] [CrossRef]
  45. Mamidi, N.; Velasco Delgadillo, R.M.; Barrera, E.V. Covalently functionalized carbon nano-onions integrated gelatin methacryloyl nanocomposite hydrogel containing γ-cyclodextrin as drug carrier for high-performance pH-triggered drug release. Pharmaceuticals 2021, 14, 291. [Google Scholar] [CrossRef]
  46. Mamidi, N.; Zuníga, A.E.; Villela-Castrejón, J. Engineering and evaluation of forcespun functionalized carbon nano-onions reinforced poly (ε-caprolactone) composite nanofibers for pH-responsive drug release. Mater. Sci. Eng. C 2020, 112, 110928. [Google Scholar] [CrossRef]
  47. Kumar, R.; Singh, M.; Meena, J.; Singhvi, P.; Thiyagarajan, D.; Saneja, A.; Panda, A.K. Hyaluronic acid-dihydroartemisinin conjugate: Synthesis, characterization and in vitro evaluation in lung cancer cells. Int. J. Biol. Macromol. 2019, 133, 495–502. [Google Scholar] [CrossRef]
  48. Leung, D.; Wurst, J.M.; Liu, T.; Martinez, R.M.; Datta-Mannan, A.; Feng, Y. Antibody conjugates-recent advances and future innovations. Antibodies 2020, 9, 2. [Google Scholar] [CrossRef] [Green Version]
  49. Tsuchikama, K.; An, Z. Antibody-drug conjugates: Recent advances in conjugation and linker chemistries. Protein Cell 2018, 9, 33–46. [Google Scholar] [CrossRef] [Green Version]
  50. Nasiri, H.; Valedkarimi, Z.; Aghebati-Maleki, L.; Majidi, J. Antibody-drug conjugates: Promising and efficient tools for targeted cancer therapy. J. Cell. Physiol. 2018, 233, 6441–6457. [Google Scholar] [CrossRef]
  51. Petersdorf, S.H.; Kopecky, K.J.; Slovak, M.; Willman, C.; Nevill, T.; Brandwein, J.; Larson, R.A.; Erba, H.P.; Stiff, P.J.; Stuart, R.K. A phase 3 study of gemtuzumab ozogamicin during induction and postconsolidation therapy in younger patients with acute myeloid leukemia. Blood J. Am. Soc. Hematol. 2013, 121, 4854–4860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Elias, D.J.; Kline, L.E.; Robbins, B.A.; Johnson Jr, H.; Pekny, K.; Benz, M.; Robb, J.A.; Walker, L.E.; Kosty, M.; Dillman, R.O. Monoclonal antibody ks1/4-methotrexate immunoconjugate studies in non-small cell lung carcinoma. Am. J. Respir. Crit. Care Med. 1994, 150, 1114–1122. [Google Scholar] [CrossRef]
  53. Erickson, H.K.; Park, P.U.; Widdison, W.C.; Kovtun, Y.V.; Garrett, L.M.; Hoffman, K.; Lutz, R.J.; Goldmacher, V.S.; Blaättler, W.A. Antibody-maytansinoid conjugates are activated in targeted cancer cells by lysosomal degradation and linker-dependent intracellular processing. Cancer Res. 2006, 66, 4426–4433. [Google Scholar] [CrossRef] [Green Version]
  54. Ehrlich, P.; Himmelweit, F.; Dale, H.; Marquardt, M. The Collected Papers of Paul Ehrlich; Franklin Book Co., Inc.: New York, NY, USA, 1957. [Google Scholar]
  55. Petersen, B.H.; DeHerdt, S.V.; Schneck, D.W.; Bumol, T.F. The human immune response to ks1/4-desacetylvinblastine (ly256787) and ks1/4-desacetylvinblastine hydrazide (ly203728) in single and multiple dose clinical studies. Cancer Res. 1991, 51, 2286–2290. [Google Scholar]
  56. Kim, E.G.; Kim, K.M. Strategies and advancement in antibody-drug conjugate optimization for targeted cancer therapeutics. Biomol. Ther. 2015, 23, 493. [Google Scholar] [CrossRef] [Green Version]
  57. LoRusso, P.M.; Weiss, D.; Guardino, E.; Girish, S.; Sliwkowski, M.X. Trastuzumab emtansine: A unique antibody-drug conjugate in development for human epidermal growth factor receptor 2-positive cancer. Clin. Cancer Res. 2011, 17, 6437–6447. [Google Scholar] [CrossRef] [Green Version]
  58. Verma, S.; Miles, D.; Gianni, L.; Krop, I.E.; Welslau, M.; Baselga, J.; Pegram, M.; Oh, D.-Y.; Diéras, V.; Guardino, E. Trastuzumab emtansine for her 2-positive advanced breast cancer. N. Engl. J. Med. 2012, 367, 1783–1791. [Google Scholar] [CrossRef] [Green Version]
  59. Beck, A.; Goetsch, L.; Dumontet, C.; Corvaïa, N. Strategies and challenges for the next generation of antibody–drug conjugates. Nat. Rev. Drug Discov. 2017, 16, 315–337. [Google Scholar] [CrossRef]
  60. Dere, R.; Yi, J.-H.; Lei, C.; Saad, O.M.; Huang, C.; Li, Y.; Baudys, J.; Kaur, S. Pk assays for antibody–drug conjugates: Case study with ado-trastuzumab emtansine. Bioanalysis 2013, 5, 1025–1040. [Google Scholar] [CrossRef]
  61. Agarwal, P.; Bertozzi, C.R. Site-specific antibody–drug conjugates: The nexus of bioorthogonal chemistry, protein engineering, and drug development. Bioconj. Chem. 2015, 26, 176–192. [Google Scholar] [CrossRef] [Green Version]
  62. Kung Sutherland, M.S.; Walter, R.B.; Jeffrey, S.; Burke, P.; Yu, C.; Kostner, H.; Stone, I.; Ryan, M.C.; Sussman, D.; Lyon, R.P. Sgn-cd33a: A novel cd33-targeting antibody–drug conjugate using a pyrrolobenzodiazepine dimer is active in models of drug-resistant aml. Blood J. Am. Soc. Hematol. 2013, 122, 1455–1463. [Google Scholar] [CrossRef] [PubMed]
  63. Bargh, J.D.; Isidro-Llobet, A.; Parker, J.S.; Spring, D.R. Cleavable linkers in antibody–drug conjugates. Chem. Soc. Rev. 2019, 48, 4361–4374. [Google Scholar] [CrossRef] [PubMed]
  64. Rosario, G. Decoupling stability and release in disulfide bonds with antibody-small molecule conjugates. Chem. Sci. 2017, 8, 366–370. [Google Scholar]
  65. Gébleux, R.; Wulhfard, S.; Casi, G.; Neri, D. Antibody format and drug release rate determine the therapeutic activity of noninternalizing antibody–drug conjugatesnoninternalizing antibody–drug conjugates. Mol. Cancer Therapeut. 2015, 14, 2606–2612. [Google Scholar] [CrossRef] [Green Version]
  66. Giansanti, F.; Flavell, D.J.; Angelucci, F.; Fabbrini, M.S.; Ippoliti, R. Strategies to improve the clinical utility of saporin-based targeted toxins. Toxins 2018, 10, 82. [Google Scholar] [CrossRef] [Green Version]
  67. Pastan, I.; Hassan, R.; FitzGerald, D.J.; Kreitman, R.J. Immunotoxin therapy of cancer. Nat. Rev. Cancer 2006, 6, 559–565. [Google Scholar] [CrossRef]
  68. Khan, M.; Hossain, M.I.; Hossain, M.K.; Rubel, M.H.K.; Hossain, K.M.; Mahfuz, A.M.U.B.; Anik, M.I. Recent progress in nanostructured smart drug delivery systems for cancer therapy: A review. ACS Appl. Bio. Mater. 2022, 5, 971–1012. [Google Scholar] [CrossRef]
  69. Lyon, R.; Bovee, T.D.; Doronina, S.O.; Burke, P.J.; Hunter, J.H.; Neff-LaFord, H.D.; Jonas, M.; Anderson, M.E.; Setter, J.R.; Senter, P.D. Reducing hydrophobicity of homogeneous antibody-drug conjugates improves pharmacokinetics and therapeutic index. Nat. Biotechnol. 2015, 33, 733–735. [Google Scholar] [CrossRef]
  70. Rudin, C.M.; Pietanza, M.C.; Bauer, T.M.; Ready, N.; Morgensztern, D.; Glisson, B.S.; Byers, L.A.; Johnson, M.L.; Burris, H.A., III; Robert, F. Rovalpituzumab tesirine, a dll3-targeted antibody-drug conjugate, in recurrent small-cell lung cancer: A first-in-human, first-in-class, open-label, phase 1 study. Lancet Oncol. 2017, 18, 42–51. [Google Scholar] [CrossRef] [Green Version]
  71. Hamblett, K.J.; Hammond, P.W.; Barnscher, S.D.; Fung, V.K.; Davies, R.H.; Wickman, G.R.; Hernandez, A.; Ding, T.; Galey, A.S.; Winters, G.C. Zw49, a her2-targeted biparatopic antibody-drug conjugate for the treatment of her2-expressing cancers. Cancer Res. 2018, 78, 3914. [Google Scholar] [CrossRef]
  72. Koopman, L.A.; Janmaat, M.L.; Jacobsen, K.; Terp, M.G.; Heuvel, E.G.V.D.; Forssman, U.; Lingnau, A.; Parren, P.W.; Ditzel, H.; Breij, E.C. An axl-specific antibody-drug conjugate shows preclinical anti-tumor activity in non-small cell lung cancer, including egfr-inhibitor resistant nsclc. Cancer Res. 2018, 78, 832. [Google Scholar] [CrossRef]
  73. Fu, Z.; Li, S.; Han, S.; Shi, C.; Zhang, Y. Antibody drug conjugate: The “biological missile” for targeted cancer therapy. Signal Transduct. Target. Ther. 2022, 7, 93. [Google Scholar] [CrossRef] [PubMed]
  74. Drago, J.Z.; Modi, S.; Chandarlapaty, S. Unlocking the potential of antibody–drug conjugates for cancer therapy. Nat. Rev. Clin. Oncol. 2021, 18, 327–344. [Google Scholar] [CrossRef] [PubMed]
  75. Huang, P.; Wang, D.; Su, Y.; Huang, W.; Zhou, Y.; Cui, D.; Zhu, X.; Yan, D. Combination of small molecule prodrug and nanodrug delivery: Amphiphilic drug–drug conjugate for cancer therapy. J. Am. Chem. Soc. 2014, 136, 11748–11756. [Google Scholar] [CrossRef]
  76. Jaiswal, S.K.; Sarathi, V.; Memon, S.S.; Garg, R.; Malhotra, G.; Verma, P.; Shah, R.; Sehemby, M.K.; Patil, V.A.; Jadhav, S. 177lu-dotatate therapy in metastatic/inoperable pheochromocytoma-paraganglioma. Endocr. Connect. 2020, 9, 864–873. [Google Scholar] [CrossRef]
  77. Zhang, Y.; He, P.; Zhang, P.; Yi, X.; Xiao, C.; Chen, X. Polypeptides–drug conjugates for anticancer therapy. Adv. Healthc. Mater. 2021, 10, 2001974. [Google Scholar] [CrossRef]
  78. Cooper, B.M.; Iegre, J.; O’Donovan, D.H.; Halvarsson, M.Ö.; Spring, D.R. Peptides as a platform for targeted therapeutics for cancer: Peptide–drug conjugates (pdcs). Chem. Soc. Rev. 2021, 50, 1480–1494. [Google Scholar] [CrossRef]
  79. Carvalho, M.R.; Carvalho, C.R.; Maia, F.R.; Caballero, D.; Kundu, S.C.; Reis, R.L.; Oliveira, J.M. Peptide-modified dendrimer nanoparticles for targeted therapy of colorectal cancer. Adv. Ther. 2019, 2, 1900132. [Google Scholar] [CrossRef] [Green Version]
  80. Thorén, P.E.; Persson, D.; Lincoln, P.; Nordén, B. Membrane destabilizing properties of cell-penetrating peptides. Biophys. Chem. 2005, 114, 169–179. [Google Scholar] [CrossRef]
  81. Fukunaga, K.; Tsutsumi, H.; Mihara, H. Self-assembling peptides as building blocks of functional materials for biomedical applications. Bull. Chem. Soc. Jpn. 2019, 92, 391–399. [Google Scholar] [CrossRef] [Green Version]
  82. Tesauro, D.; Accardo, A.; Diaferia, C.; Milano, V.; Guillon, J.; Ronga, L.; Rossi, F. Peptide-based drug-delivery systems in biotechnological applications: Recent advances and perspectives. Molecules 2019, 24, 351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Guidotti, G.; Brambilla, L.; Rossi, D. Cell-penetrating peptides: From basic research to clinics. Trends Pharmacol. Sci. 2017, 38, 406–424. [Google Scholar] [CrossRef] [PubMed]
  84. Reissmann, S. Cell penetration: Scope and limitations by the application of cell-penetrating peptides. J. Pept. Sci. 2014, 20, 760–784. [Google Scholar] [CrossRef] [PubMed]
  85. Jang, S.; Hyun, S.; Kim, S.; Lee, S.; Lee, I.S.; Baba, M.; Lee, Y.; Yu, J. Cell-penetrating, dimeric α-helical peptides: Nanomolar inhibitors of hiv-1 transcription. Angew. Chem. 2014, 126, 10250–10253. [Google Scholar] [CrossRef]
  86. Oh, J.H.; Chong, S.E.; Nam, S.; Hyun, S.; Choi, S.; Gye, H.; Jang, S.; Jang, J.; Hwang, S.W.; Yu, J. Multimeric amphipathic α-helical sequences for rapid and efficient intracellular protein transport at nanomolar concentrations. Adv. Sci. 2018, 5, 1800240. [Google Scholar] [CrossRef]
  87. Hyun, S.; Choi, Y.; Lee, H.N.; Lee, C.; Oh, D.; Lee, D.-K.; Lee, C.; Lee, Y.; Yu, J. Construction of histidine-containing hydrocarbon stapled cell penetrating peptides for in vitro and in vivo delivery of sirnas. Chem. Sci. 2018, 9, 3820–3827. [Google Scholar] [CrossRef] [Green Version]
  88. Dougherty, P.G.; Wen, J.; Pan, X.; Koley, A.; Ren, J.-G.; Sahni, A.; Basu, R.; Salim, H.; Appiah Kubi, G.; Qian, Z. Enhancing the cell permeability of stapled peptides with a cyclic cell-penetrating peptide. J. Med. Chem. 2019, 62, 10098–10107. [Google Scholar] [CrossRef]
  89. LaRochelle, J.R.; Cobb, G.B.; Steinauer, A.; Rhoades, E.; Schepartz, A. Fluorescence correlation spectroscopy reveals highly efficient cytosolic delivery of certain penta-arg proteins and stapled peptides. J. Am. Chem. Soc. 2015, 137, 2536–2541. [Google Scholar] [CrossRef] [Green Version]
  90. Hughes, J.P.; Rees, S.; Kalindjian, S.B.; Philpott, K.L. Principles of early drug discovery. Br. J. Pharmacol. 2011, 162, 1239–1249. [Google Scholar] [CrossRef] [Green Version]
  91. Park, S.E.; El-Sayed, N.S.; Shamloo, K.; Lohan, S.; Kumar, S.; Sajid, M.I.; Tiwari, R.K. Targeted delivery of cabazitaxel using cyclic cell-penetrating peptide and biomarkers of extracellular matrix for prostate and breast cancer therapy. Bioconj. Chem. 2021, 32, 1898–1914. [Google Scholar] [CrossRef]
  92. Tripodi, A.A.P.; Ranđelović, I.; Biri-Kovács, B.; Szeder, B.; Mező, G.; Tóvári, J. In vivo tumor growth inhibition and antiangiogenic effect of cyclic ngr peptide-daunorubicin conjugates developed for targeted drug delivery. Pathol. Oncol. Res. 2020, 26, 1879–1892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Cai, Y.; Shen, H.; Zhan, J.; Lin, M.; Dai, L.; Ren, C.; Shi, Y.; Liu, J.; Gao, J.; Yang, Z. Supramolecular “trojan horse” for nuclear delivery of dual anticancer drugs. J. Am. Chem. Soc. 2017, 139, 2876–2879. [Google Scholar] [CrossRef] [PubMed]
  94. Ma, Y.; Mou, Q.; Zhu, X.; Yan, D. Small molecule nanodrugs for cancer therapy. Mater. Today Chem. 2017, 4, 26–39. [Google Scholar] [CrossRef]
  95. 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] [Green Version]
  96. Li, J.; Li, X.; Liu, P. Doxorubicin-doxorubicin conjugate prodrug as drug self-delivery system for intracellular ph-triggered slow release. Colloids Surf. B Biointerfaces 2020, 185, 110608. [Google Scholar] [CrossRef]
  97. Gao, D.; Guo, X.; Zhang, X.; Chen, S.; Wang, Y.; Chen, T.; Huang, G.; Gao, Y.; Tian, Z.; Yang, Z. Multifunctional phototheranostic nanomedicine for cancer imaging and treatment. Mater. Today Bio. 2020, 5, 100035. [Google Scholar] [CrossRef] [PubMed]
  98. Hou, M.; Li, S.; Xu, Z.; Li, B. A reduction-responsive amphiphilic methotrexate-podophyllotoxin conjugate for targeted chemotherapy. Chem. Asian J. 2019, 14, 3840–3844. [Google Scholar] [CrossRef]
  99. Cheng, C.; Sui, B.; Wang, M.; Hu, X.; Shi, S.; Xu, P. Carrier-free nanoassembly of curcumin–erlotinib conjugate for cancer targeted therapy. Adv. Healthc. Mater. 2020, 9, 2001128. [Google Scholar] [CrossRef]
  100. Liu, Y.; Bhattarai, P.; Dai, Z.; Chen, X. Photothermal therapy and photoacoustic imaging via nanotheranostics in fighting cancer. Chem. Soc. Rev. 2019, 48, 2053–2108. [Google Scholar] [CrossRef]
  101. Beik, J.; Abed, Z.; Ghoreishi, F.S.; Hosseini-Nami, S.; Mehrzadi, S.; Shakeri-Zadeh, A.; Kamrava, S.K. Nanotechnology in hyperthermia cancer therapy: From fundamental principles to advanced applications. J. Control. Release 2016, 235, 205–221. [Google Scholar] [CrossRef]
  102. Abadeer, N.S.; Murphy, C.J. Recent progress in cancer thermal therapy using gold nanoparticles. In Nanomaterials and Neoplasms; Taylor & Francis Group: Oxfordshire, UK, 2021; pp. 143–217. [Google Scholar]
  103. Tan, C.; Cao, X.; Wu, X.-J.; He, Q.; Yang, J.; Zhang, X.; Chen, J.; Zhao, W.; Han, S.; Nam, G.-H. Recent advances in ultrathin two-dimensional nanomaterials. Chem. Rev. 2017, 117, 6225–6331. [Google Scholar] [CrossRef] [PubMed]
  104. Gao, P.; Wang, H.; Cheng, Y. Strategies for efficient photothermal therapy at mild temperatures: Progresses and challenges. Chin. Chem. Lett. 2022, 33, 575–586. [Google Scholar] [CrossRef]
  105. Li, X.; Lovell, J.F.; Yoon, J.; Chen, X. Clinical development and potential of photothermal and photodynamic therapies for cancer. Nat. Rev. Clin. Oncol. 2020, 17, 657–674. [Google Scholar] [CrossRef]
  106. Yang, T.; Wang, Q.; Lv, X.; Song, X.; Ke, H.; Guo, Z.; Huang, X.; Hu, J.; Li, Z.; Yang, P. Albumin-coordinated assembly of clearable platinum nanodots for photo-induced cancer theranostics. Biomaterials 2018, 154, 248–260. [Google Scholar]
  107. Li, X.; Shan, J.; Zhang, W.; Su, S.; Yuwen, L.; Wang, L. Recent advances in synthesis and biomedical applications of two-dimensional transition metal dichalcogenide nanosheets. Small 2017, 13, 1602660. [Google Scholar] [CrossRef]
  108. Gai, S.; Yang, G.; Yang, P.; He, F.; Lin, J.; Jin, D.; Xing, B. Recent advances in functional nanomaterials for light–triggered cancer therapy. Nano Today 2018, 19, 146–187. [Google Scholar] [CrossRef]
  109. Wang, T.; Wang, D.; Yu, H.; Feng, B.; Zhou, F.; Zhang, H.; Zhou, L.; Jiao, S.; Li, Y. A cancer vaccine-mediated postoperative immunotherapy for recurrent and metastatic tumors. Nat. Commun. 2018, 9, 1532. [Google Scholar] [CrossRef] [Green Version]
  110. Wang, P.; Zhang, L.; Zheng, W.; Cong, L.; Guo, Z.; Xie, Y.; Wang, L.; Tang, R.; Feng, Q.; Hamada, Y. Thermo-triggered release of crispr-cas9 system by lipid-encapsulated gold nanoparticles for tumor therapy. Angew. Chem. Int. Ed. 2018, 57, 1491–1496. [Google Scholar] [CrossRef]
  111. Guo, Z.; Zhu, S.; Yong, Y.; Zhang, X.; Dong, X.; Du, J.; Xie, J.; Wang, Q.; Gu, Z.; Zhao, Y. Synthesis of bsa-coated bioi@ bi2s3 semiconductor heterojunction nanoparticles and their applications for radio/photodynamic/photothermal synergistic therapy of tumor. Adv. Mater. 2017, 29, 1704136. [Google Scholar] [CrossRef]
  112. Zhang, D.; Zhang, J.; Li, Q.; Tian, H.; Zhang, N.; Li, Z.; Luan, Y. pH-and enzyme-sensitive IR820–paclitaxel conjugate self-assembled nanovehicles for near-infrared fluorescence imaging-guided chemo–photothermal therapy. ACS Appl. Mater. Interfaces 2018, 10, 30092–30102. [Google Scholar] [CrossRef]
  113. Ao, M.; Yu, F.; Li, Y.; Zhong, M.; Tang, Y.; Yang, H.; Wu, X.; Zhuang, Y.; Wang, H.; Sun, X. Carrier-free nanoparticles of camptothecin prodrug for chemo-photothermal therapy: The making, in vitro and in vivo testing. J. Nanobiotechnol. 2021, 19, 350. [Google Scholar] [CrossRef] [PubMed]
  114. Du, Q.; Qin, X.; Zhang, M.; Zhao, Z.; Li, Q.; Ren, X.; Wang, N.; Luan, Y. A mitochondrial-metabolism-regulatable carrier-free nanodrug to amplify the sensitivity of photothermal therapy. Chem. Commun. 2021, 57, 8993–8996. [Google Scholar] [CrossRef] [PubMed]
  115. Shi, X.; Zhang, C.Y.; Gao, J.; Wang, Z. Recent advances in photodynamic therapy for cancer and infectious diseases. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2019, 11, e1560. [Google Scholar] [CrossRef] [PubMed]
  116. Felsher, D.W. Cancer revoked: Oncogenes as therapeutic targets. Nat. Rev. Cancer 2003, 3, 375–379. [Google Scholar] [CrossRef] [PubMed]
  117. Letokhov, V. Laser biology and medicine. Nature 1985, 316, 325–330. [Google Scholar] [CrossRef]
  118. Rui, L.-L.; Cao, H.-L.; Xue, Y.-D.; Liu, L.-C.; Xu, L.; Gao, Y.; Zhang, W.-A. Functional organic nanoparticles for photodynamic therapy. Chin. Chem. Lett. 2016, 27, 1412–1420. [Google Scholar] [CrossRef]
  119. Maas, A.L.; Carter, S.L.; Wileyto, E.P.; Miller, J.; Yuan, M.; Yu, G.; Durham, A.C.; Busch, T.M. Tumor vascular microenvironment determines responsiveness to photodynamic therapy. Cancer Res. 2012, 72, 2079–2088. [Google Scholar] [CrossRef] [Green Version]
  120. Mroz, P.; Yaroslavsky, A.; Kharkwal, G.B.; Hamblin, M.R. Cell death pathways in photodynamic therapy of cancer. Cancers 2011, 3, 2516–2539. [Google Scholar] [CrossRef] [Green Version]
  121. Preise, D.; Oren, R.; Glinert, I.; Kalchenko, V.; Jung, S.; Scherz, A.; Salomon, Y. Systemic antitumor protection by vascular-targeted photodynamic therapy involves cellular and humoral immunity. Cancer Immunol. Immunother. 2009, 58, 71–84. [Google Scholar] [CrossRef]
  122. Castano, A.P.; Mroz, P.; Wu, M.X.; Hamblin, M.R. Photodynamic therapy plus low-dose cyclophosphamide generates antitumor immunity in a mouse model. Proc. Natl. Acad. Sci. USA 2008, 105, 5495–5500. [Google Scholar] [CrossRef] [Green Version]
  123. Pye, H.; Stamati, I.; Yahioglu, G.; Butt, M.A.; Deonarain, M. Antibody-directed phototherapy (ADP). Antibodies 2013, 2, 270–305. [Google Scholar] [CrossRef]
  124. Hao, Y.; Chen, Y.; He, X.; Yu, Y.; Han, R.; Li, Y.; Yang, C.; Hu, D.; Qian, Z. Polymeric nanoparticles with ros-responsive prodrug and platinum nanozyme for enhanced chemophotodynamic therapy of colon cancer. Adv. Sci. 2020, 7, 2001853. [Google Scholar] [CrossRef] [PubMed]
  125. Ha, S.Y.; Zhou, Y.; Fong, W.-P.; Ng, D.K. Multifunctional molecular therapeutic agent for targeted and controlled dual chemo-and photodynamic therapy. J. Med. Chem. 2020, 63, 8512–8523. [Google Scholar] [CrossRef] [PubMed]
  126. Um, W.; Park, J.; Ko, H.; Lim, S.; Yoon, H.Y.; Shim, M.K.; Lee, S.; Ko, Y.J.; Kim, M.J.; Park, J.H. Visible light-induced apoptosis activatable nanoparticles of photosensitizer-devd-anticancer drug conjugate for targeted cancer therapy. Biomaterials 2019, 224, 119494. [Google Scholar] [CrossRef] [PubMed]
  127. Thankarajan, E.; Tuchinsky, H.; Aviel-Ronen, S.; Bazylevich, A.; Gellerman, G.; Patsenker, L. Antibody guided activatable NIR photosensitizing system for fluorescently monitored photodynamic therapy with reduced side effects. J. Control. Release 2022, 343, 506–517. [Google Scholar] [CrossRef] [PubMed]
  128. Liu, X.; Zheng, C.; Kong, Y.; Wang, H.; Wang, L. An in situ nanoparticle recombinant strategy for the enhancement of photothermal therapy. Chin. Chem. Lett. 2022, 33, 328–333. [Google Scholar] [CrossRef]
  129. Prasher, P.; Sharma, M.; Mudila, H.; Gupta, G.; Sharma, A.K.; Kumar, D.; Bakshi, H.A.; Negi, P.; Kapoor, D.N.; Chellappan, D.K. Emerging trends in clinical implications of bio-conjugated silver nanoparticles in drug delivery. Colloid Interface Sci. Commun. 2020, 35, 100244. [Google Scholar] [CrossRef]
  130. Power, S.; Slattery, M.M.; Lee, M.J. Nanotechnology and its relationship to interventional radiology. Part ii: Drug delivery, thermotherapy, and vascular intervention. Cardiovasc. Interv. Radiol. 2011, 34, 676–690. [Google Scholar] [CrossRef]
  131. Sanna, V.; Pala, N.; Sechi, M. Targeted therapy using nanotechnology: Focus on cancer. Int. J. Nanomed. 2014, 9, 467. [Google Scholar]
  132. Parra-Nieto, J.; Del Cid, M.A.G.; de Cárcer, I.A.; Baeza, A. Inorganic porous nanoparticles for drug delivery in antitumoral therapy. Biotechnol. J. 2021, 16, 2000150. [Google Scholar] [CrossRef]
  133. Abdellatif, A.A.; Khan, R.A.; Alhowail, A.H.; Alqasoumi, A.; Sajid, S.M.; Mohammed, A.M.; Alsharidah, M.; Al Rugaie, O.; Mousa, A.M. Octreotide-conjugated silver nanoparticles for active targeting of somatostatin receptors and their application in a nebulized rat model. Nanotechnol. Rev. 2022, 11, 266–283. [Google Scholar] [CrossRef]
  134. Gormley, A.J.; Larson, N.; Banisadr, A.; Robinson, R.; Frazier, N.; Ray, A.; Ghandehari, H. Plasmonic photothermal therapy increases the tumor mass penetration of hpma copolymers. J. Control. Release 2013, 166, 130–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Vigderman, L.; Khanal, B.P.; Zubarev, E.R. Functional gold nanorods: Synthesis, self-assembly, and sensing applications. Adv. Mater. 2012, 24, 4811–4841. [Google Scholar] [CrossRef] [PubMed]
  136. Kennedy, L.C.; Bickford, L.R.; Lewinski, N.A.; Coughlin, A.J.; Hu, Y.; Day, E.S.; West, J.L.; Drezek, R.A. A new era for cancer treatment: Gold-nanoparticle-mediated thermal therapies. Small 2011, 7, 169–183. [Google Scholar] [CrossRef] [PubMed]
  137. Choi, J.; Kim, S.Y. Photothermally enhanced photodynamic therapy based on glutathione-responsive pheophorbide a-conjugated gold nanorod formulations for cancer theranostic applications. J. Ind. Eng. Chem. 2020, 85, 66–74. [Google Scholar] [CrossRef]
  138. Santos, H.A.; Mäkilä, E.; Airaksinen, A.J.; Bimbo, L.M.; Hirvonen, J. Porous silicon nanoparticles for nanomedicine: Preparation and biomedical applications. Nanomedicine 2014, 9, 535–554. [Google Scholar] [CrossRef]
  139. Duncan, R. Polymer therapeutics: Top 10 selling pharmaceuticals—what next? J. Control. Release 2014, 190, 371–380. [Google Scholar] [CrossRef]
  140. Peng, S.; Zhang, F.; Huang, B.; Wang, J.; Zhang, L. Mesoporous silica nanoprodrug encapsulated with near-infrared absorption dye for photothermal therapy combined with chemotherapy. ACS Appl. Bio. Mater. 2021, 4, 8225–8235. [Google Scholar] [CrossRef]
  141. Thakare, V.S.; Das, M.; Jain, A.K.; Patil, S.; Jain, S. Carbon nanotubes in cancer theragnosis. Nanomedicine 2010, 5, 1277–1301. [Google Scholar] [CrossRef]
  142. Dhar, S.; Liu, Z.; Thomale, J.; Dai, H.; Lippard, S.J. Targeted single-wall carbon nanotube-mediated pt (iv) prodrug delivery using folate as a homing device. J. Am. Chem. Soc. 2008, 130, 11467–11476. [Google Scholar] [CrossRef] [Green Version]
  143. Lotfollahzadeh, S.; Hosseini, E.S.; Mahmoudi Aznaveh, H.; Nikkhah, M.; Hosseinkhani, S. Trail/s-layer/graphene quantum dot nanohybrid enhanced stability and anticancer activity of trail on colon cancer cells. Sci. Rep. 2022, 12, 5851. [Google Scholar] [CrossRef] [PubMed]
  144. Andreou, C.; Weissleder, R.; Kircher, M.F. Multiplexed imaging in oncology. Nat. Biomed. Eng. 2022, 6, 527–540. [Google Scholar] [CrossRef] [PubMed]
  145. Nanayakkara, A.K.; Follit, C.A.; Chen, G.; Williams, N.S.; Vogel, P.D.; Wise, J.G. Targeted inhibitors of p-glycoprotein increase chemotherapeutic-induced mortality of multidrug resistant tumor cells. Sci. Rep. 2018, 8, 967. [Google Scholar] [CrossRef] [Green Version]
  146. Marusyk, A.; Almendro, V.; Polyak, K. Intra-tumour heterogeneity: A looking glass for cancer? Nat. Rev. Cancer 2012, 12, 323–334. [Google Scholar] [CrossRef] [PubMed]
  147. Kerbel, R. Molecular and physiologic mechanisms of drug resistance in cancer: An overview. Cancer Metastasis Rev. 2001, 20, 1. [Google Scholar] [CrossRef]
  148. Wang, F.; Zhang, D.; Zhang, Q.; Chen, Y.; Zheng, D.; Hao, L.; Duan, C.; Jia, L.; Liu, G.; Liu, Y. Synergistic effect of folate-mediated targeting and verapamil-mediated P-gp inhibition with paclitaxel-polymer micelles to overcome multi-drug resistance. Biomaterials 2011, 32, 9444–9456. [Google Scholar] [CrossRef]
  149. Jadia, R.; Scandore, C.; Rai, P. Nanoparticles for effective combination therapy of cancer. Int. J. Nanotechnol. Nanomed. 2016, 1, 1–26. [Google Scholar]
  150. Hu, C.-M.J.; Zhang, L. Nanoparticle-based combination therapy toward overcoming drug resistance in cancer. Biochem. Pharmacol. 2012, 83, 1104–1111. [Google Scholar] [CrossRef]
  151. Chiu, G.N.; Wong, M.-Y.; Ling, L.-U.; Shaikh, I.M.; Tan, K.-B.; Chaudhury, A.; Tan, B.-J. Lipid-based nanoparticulate systems for the delivery of anti-cancer drug cocktails: Implications on pharmacokinetics and drug toxicities. Curr. Drug Metab. 2009, 10, 861–874. [Google Scholar] [CrossRef]
  152. Hu, C.-M.J.; Aryal, S.; Zhang, L. Nanoparticle-assisted combination therapies for effective cancer treatment. Ther. Deliv. 2010, 1, 323–334. [Google Scholar] [CrossRef]
  153. Wang, S.; Deng, H.; Huang, P.; Sun, P.; Huang, X.; Su, Y.; Zhu, X.; Shen, J.; Yan, D. Real-time self-tracking of an anticancer small molecule nanodrug based on colorful fluorescence variations. RSC Adv. 2016, 6, 12472–12478. [Google Scholar] [CrossRef]
  154. Feng, L.; Gao, M.; Tao, D.; Chen, Q.; Wang, H.; Dong, Z.; Chen, M.; Liu, Z. Cisplatin-prodrug-constructed liposomes as a versatile theranostic nanoplatform for bimodal imaging guided combination cancer therapy. Adv. Funct. Mater. 2016, 26, 2207–2217. [Google Scholar] [CrossRef]
  155. Wang, D.; Yu, C.; Xu, L.; Shi, L.; Tong, G.; Wu, J.; Liu, H.; Yan, D.; Zhu, X. Nucleoside analogue-based supramolecular nanodrugs driven by molecular recognition for synergistic cancer therapy. J. Am. Chem. Soc. 2018, 140, 8797–8806. [Google Scholar] [CrossRef]
  156. Sun, M.; Qian, Q.; Shi, L.; Xu, L.; Liu, Q.; Zhou, L.; Zhu, X.; Yue, J.-M.; Yan, D. Amphiphilic drug-drug conjugate for cancer therapy with combination of chemotherapeutic and antiangiogenesis drugs. Sci. China Chem. 2020, 63, 35–41. [Google Scholar] [CrossRef]
  157. Wang, L.; Guo, B.; Wang, R.; Jiang, Y.; Qin, S.; Liang, S.; Zhao, Y.; Guo, W.; Li, K.; Fan, X. Inhibition of cell growth and up-regulation of mad2 in human oesophageal squamous cell carcinoma after treatment with the src/abl inhibitor dasatinib. Clin. Sci. 2012, 122, 13–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Yang, L.; Xu, J.; Xie, Z.; Song, F.; Wang, X.; Tang, R. Carrier-free prodrug nanoparticles based on dasatinib and cisplatin for efficient antitumor in vivo. Asian J. Pharm. Sci. 2021, 16, 762–771. [Google Scholar] [CrossRef] [PubMed]
  159. Shan, X.; Zhao, Z.; Wang, C.; Sun, J.; He, Z.; Luo, C.; Zhang, S. Emerging prodrug-engineered nanomedicines for synergistic chemo-phototherapy. Chem. Eng. J. 2022, 442, 136383. [Google Scholar] [CrossRef]
  160. Yang, Y.; Zhang, Y.; Wang, R.; Rong, X.; Liu, T.; Xia, X.; Fan, J.; Sun, W.; Peng, X. A glutathione activatable pro-drug-photosensitizer for combined chemotherapy and photodynamic therapy. Chin. Chem. Lett. 2022, 33, 4583–4586. [Google Scholar] [CrossRef]
  161. Castano, A.P.; Mroz, P.; Hamblin, M.R. Photodynamic therapy and anti-tumour immunity. Nat. Rev. Cancer 2006, 6, 535–545. [Google Scholar] [CrossRef] [Green Version]
  162. Ji, C.; Gao, Q.; Dong, X.; Yin, W.; Gu, Z.; Gan, Z.; Zhao, Y.; Yin, M. A size-reducible nanodrug with an aggregation-enhanced photodynamic effect for deep chemo-photodynamic therapy. Angew. Chem. 2018, 130, 11554–11558. [Google Scholar] [CrossRef]
  163. Chen, H.; Zeng, X.; Tham, H.P.; Phua, S.Z.F.; Cheng, W.; Zeng, W.; Shi, H.; Mei, L.; Zhao, Y. Nir-light-activated combination therapy with a precise ratio of photosensitizer and prodrug using a host–guest strategy. Angew. Chem. Int. Ed. 2019, 58, 7641–7646. [Google Scholar] [CrossRef] [PubMed]
  164. Lee, J.; Davaa, E.; Jiang, Y.; Shin, K.-J.; Kim, M.H.; An, H.; Kim, J.; Cho, S.K.; Yang, S.-G. Pheophorbide a and sn38 conjugated hyaluronan nanoparticles for photodynamic-and cascadic chemotherapy of cancer stem-like ovarian cancer. Carbohydr. Polym. 2022, 289, 119455. [Google Scholar] [CrossRef] [PubMed]
  165. Xu, S.; Zhu, X.; Zhang, C.; Huang, W.; Zhou, Y.; Yan, D. Oxygen and pt (ii) self-generating conjugate for synergistic photo-chemo therapy of hypoxic tumor. Nat. Commun. 2018, 9, 2053. [Google Scholar] [CrossRef] [PubMed]
  166. Li, X.; Yang, M.; Cao, J.; Gu, H.; Liu, W.; Xia, T.; Sun, W.; Fan, J.; Peng, X. H-aggregates of prodrug-hemicyanine conjugate for enhanced photothermal therapy and sequential hypoxia-activated chemotherapy. ACS Mater. Lett. 2022, 4, 724–732. [Google Scholar] [CrossRef]
  167. Zhou, L.; Du, C.; Zhang, R.; Dong, C. Stimuli-responsive dual drugs-conjugated polydopamine nanoparticles for the combination photothermal-cocktail chemotherapy. Chin. Chem. Lett. 2021, 32, 561–564. [Google Scholar] [CrossRef]
  168. Nam, J.; Son, S.; Park, K.S.; Zou, W.; Shea, L.D.; Moon, J.J. Cancer nanomedicine for combination cancer immunotherapy. Nat. Rev. Mater. 2019, 4, 398–414. [Google Scholar] [CrossRef]
  169. Abdel-Wahab, N.; Alshawa, A.; Suarez-Almazor, M.E. Adverse events in cancer immunotherapy. In Immunotherapy; Springer: Berlin/Heidelberg, Germany, 2017; pp. 155–174. [Google Scholar]
  170. June, C.H.; Warshauer, J.T.; Bluestone, J.A. Is autoimmunity the achilles’ heel of cancer immunotherapy? Nat. Med. 2017, 23, 540–547. [Google Scholar] [CrossRef] [Green Version]
  171. Geng, Z.; Wang, L.; Liu, K.; Liu, J.; Tan, W. Enhancing anti-pd-1 immunotherapy by nanomicelles self-assembled from multivalent aptamer drug conjugates. Angew. Chem. 2021, 133, 15587–15593. [Google Scholar] [CrossRef]
  172. Hu, D.; Zhang, W.; Xiang, J.; Li, D.; Chen, Y.; Yuan, P.; Shao, S.; Zhou, Z.; Shen, Y.; Tang, J. A ros-responsive synergistic delivery system for combined immunotherapy and chemotherapy. Mater. Today Bio. 2022, 14, 100284. [Google Scholar] [CrossRef]
  173. Bai, S.; Yang, L.L.; Wang, Y.; Zhang, T.; Fu, L.; Yang, S.; Wan, S.; Wang, S.; Jia, D.; Li, B. Prodrug-based versatile nanomedicine for enhancing cancer immunotherapy by increasing immunogenic cell death. Small 2020, 16, 2000214. [Google Scholar] [CrossRef]
  174. Libutti, S.K.; Paciotti, G.F.; Byrnes, A.A.; Alexander, H.R.; Gannon, W.E.; Walker, M.; Seidel, G.D.; Yuldasheva, N.; Tamarkin, L. Phase i and pharmacokinetic studies of cyt-6091, a novel pegylated colloidal gold-rhtnf nanomedicinepegylated colloidal gold-rhtnf nanomedicine phase i trial. Clin. Cancer Res. 2010, 16, 6139–6149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Xue, G.; Wang, Z.; Zheng, N.; Fang, J.; Mao, C.; Li, X.; Jin, G.; Ming, X.; Lu, Y. Elimination of acquired resistance to pd-1 blockade via the concurrent depletion of tumour cells and immunosuppressive cells. Nat. Biomed. Eng. 2021, 5, 1306–1319. [Google Scholar] [CrossRef] [PubMed]
  176. Luo, Y.; Chen, Z.; Sun, M.; Li, B.; Pan, F.; Ma, A.; Liao, J.; Yin, T.; Tang, X.; Huang, G. IL-12 nanochaperone-engineered CAR T cell for robust tumor-immunotherapy. Biomaterials 2022, 281, 121341. [Google Scholar] [CrossRef]
  177. Li, Q.; Zhang, Y.; Huang, X.; Yang, D.; Weng, L.; Ou, C.; Song, X.; Dong, X. An NIR-II light responsive antibacterial gelation for repetitious photothermal/thermodynamic synergistic therapy. Chem. Eng. J. 2021, 407, 127200. [Google Scholar] [CrossRef]
  178. Liu, Y.; Tang, P.; Xiao, P.; Luo, S.; Zhang, S.; Zhang, H.; Yang, Y.; Wu, D. Molecular stacking composite nanoparticles of gossypolone and thermodynamic agent for elimination of large tumor in mice via electrothermal-thermodynamic-chemo trimodal combination therapy. Adv. Funct. Mater. 2022, 32, 2201666. [Google Scholar] [CrossRef]
  179. Xia, R.; Zheng, X.; Hu, X.; Liu, S.; Xie, Z. Photothermal-controlled generation of alkyl radical from organic nanoparticles for tumor treatment. ACS Appl. Mater. Interfaces 2019, 11, 5782–5790. [Google Scholar] [CrossRef]
  180. Gao, D.; Chen, T.; Chen, S.; Ren, X.; Han, Y.; Li, Y.; Wang, Y.; Guo, X.; Wang, H.; Chen, X. Targeting hypoxic tumors with hybrid nanobullets for oxygen-independent synergistic photothermal and thermodynamic therapy. Nano-Micro Lett. 2021, 13, 99. [Google Scholar] [CrossRef] [PubMed]
  181. Gao, D.; Shi, Y.; Ni, J.; Chen, S.; Wang, Y.; Zhao, B.; Song, M.; Guo, X.; Ren, X.; Zhang, X. Nir/mri-guided oxygen-independent carrier-free anti-tumor nano-theranostics. Small 2021, 2106000. [Google Scholar] [CrossRef]
Figure 1. Number of publications per year on “conjugate nanomedicine” from 2010 to 2021, based on web of science database.
Figure 1. Number of publications per year on “conjugate nanomedicine” from 2010 to 2021, based on web of science database.
Pharmaceutics 14 01522 g001
Figure 2. Overview of conjugated nanomaterials and representative clinical application of conjugated nanomaterials. (Photodynamic therapy, PDT; Photothermal therapy, PTT; Thermodynamic therapy, TDT).
Figure 2. Overview of conjugated nanomaterials and representative clinical application of conjugated nanomaterials. (Photodynamic therapy, PDT; Photothermal therapy, PTT; Thermodynamic therapy, TDT).
Pharmaceutics 14 01522 g002
Figure 3. Structure and mechanism of action of ADC. Reprinted with permission from Ref. [49]. 2018, Springer.
Figure 3. Structure and mechanism of action of ADC. Reprinted with permission from Ref. [49]. 2018, Springer.
Pharmaceutics 14 01522 g003
Figure 4. (A) Schematic illustration for preparation of dual-drug assemblies and the nuclear drug delivery, (B) Chemical structures of HCPT and HP, TEM image of solution containing 100 μM of (C) Complex 1, and (D) Complex 2, (E) CLSM images of A549 cells treated with HCPT, HP, Complex 1, and Complex 2 (100 μM) for 2 h, and then stained with 1 × Red dot 1. (F) in vivo anticancer efficacy, * p < 0.05 and *** p < 0.001. Reprinted with permission from Ref. [93]. 2017, American Chemical Society.
Figure 4. (A) Schematic illustration for preparation of dual-drug assemblies and the nuclear drug delivery, (B) Chemical structures of HCPT and HP, TEM image of solution containing 100 μM of (C) Complex 1, and (D) Complex 2, (E) CLSM images of A549 cells treated with HCPT, HP, Complex 1, and Complex 2 (100 μM) for 2 h, and then stained with 1 × Red dot 1. (F) in vivo anticancer efficacy, * p < 0.05 and *** p < 0.001. Reprinted with permission from Ref. [93]. 2017, American Chemical Society.
Pharmaceutics 14 01522 g004
Figure 5. (A)The development of erlotinib–PEG–curcumin(EPC) nano-assembly and its characterization. (B) in vivo study of the EPC nano-assembly in a pancreatic xenograft mouse tumor model: quantitative fluorescence intensity of CCM and EPC in different organs and tumor growth profiles. * p < 0.05, ** p < 0.01 and *** p < 0.001. (C) Fluorescence images of drug penetration and cell killing effect in BxPC-3 tumor spheroids. Reprinted with permission from Ref. [99]. 2020, Wiley Online Library. (D) Schematic representation of visible light-induced apoptosis activatable nanoparticles of Ce6–DEVD–MMAE for targeted cancer therapy. (E) In vivo therapeutic efficacy of Ce6–DEVD–MMAE nanoparticles in tumor-bearing mice: Ex vivo apoptosis fluorescence imaging with Annexin V–Cy 5 (green color; Annexin V–Cy 5, blue color; DAPI). Reprinted with permission from Ref. [126]. 2019, Elsevier.
Figure 5. (A)The development of erlotinib–PEG–curcumin(EPC) nano-assembly and its characterization. (B) in vivo study of the EPC nano-assembly in a pancreatic xenograft mouse tumor model: quantitative fluorescence intensity of CCM and EPC in different organs and tumor growth profiles. * p < 0.05, ** p < 0.01 and *** p < 0.001. (C) Fluorescence images of drug penetration and cell killing effect in BxPC-3 tumor spheroids. Reprinted with permission from Ref. [99]. 2020, Wiley Online Library. (D) Schematic representation of visible light-induced apoptosis activatable nanoparticles of Ce6–DEVD–MMAE for targeted cancer therapy. (E) In vivo therapeutic efficacy of Ce6–DEVD–MMAE nanoparticles in tumor-bearing mice: Ex vivo apoptosis fluorescence imaging with Annexin V–Cy 5 (green color; Annexin V–Cy 5, blue color; DAPI). Reprinted with permission from Ref. [126]. 2019, Elsevier.
Pharmaceutics 14 01522 g005
Figure 6. (A) Schematic representation of demonstrating the photothermally enhanced photodynamic therapy of GSH-responsive Pheo-conjugated AuNR. GSH-mediated Pheo release kinetics: (B) in vitro Pheo release profile of FAPAuNR100–Pheo with different GSH concentrations, (C) singlet oxygen generation behavior of free FAPAuNR100–Pheo at intensities of 0.2, 10, 50, and 100 mW/cm2, (D) photothermal effect of FAPAuNR100–Pheo at different 880 nm laser intensities, (E) Assay on the phototoxicity of free Pheo, FAPAuNR100–Pheo and FAPAuNR100–Pheo. Reprinted with permission from Ref. [137]. 2020, Elsevier.
Figure 6. (A) Schematic representation of demonstrating the photothermally enhanced photodynamic therapy of GSH-responsive Pheo-conjugated AuNR. GSH-mediated Pheo release kinetics: (B) in vitro Pheo release profile of FAPAuNR100–Pheo with different GSH concentrations, (C) singlet oxygen generation behavior of free FAPAuNR100–Pheo at intensities of 0.2, 10, 50, and 100 mW/cm2, (D) photothermal effect of FAPAuNR100–Pheo at different 880 nm laser intensities, (E) Assay on the phototoxicity of free Pheo, FAPAuNR100–Pheo and FAPAuNR100–Pheo. Reprinted with permission from Ref. [137]. 2020, Elsevier.
Pharmaceutics 14 01522 g006
Figure 7. (A) Schematic representation of demonstrating combination of therapeutic hyaluronan nanoparticles conjugated with photodynamic pheophorbide A and ROS-cleavable thioketal-SN38 and drug delivery mechanism of nanoparticles. (B) Characterization of PhoeA-SN38-HC NPs: NIR induced singlet oxygen generation from NPs. In the presence of DMA (100 μM), 1 mg/mL NPs were exposed to light, and fluorescence intensity (λ = 420 nm) of DMA was measured by spectrometer, Light induced drug release from NPs depending on light energy. (C) In vivo PDT treatment with HC-PheoA-SN38: body weights and tumor growth curves of HEY-T30 xenograft BALB/C nude mouse, * p < 0.05 and *** p < 0.001. Reprinted with permission from [164]. Copyright@ Elsevier. (D) The preparation of dual drugs-conjugated PDOXCBs nanoparticles. (E) Both PDOXCB18 and PDOXCB27 cytotoxicity against cancer cells. Reprinted with permission from Ref. [167]. 2021, Elsevier.
Figure 7. (A) Schematic representation of demonstrating combination of therapeutic hyaluronan nanoparticles conjugated with photodynamic pheophorbide A and ROS-cleavable thioketal-SN38 and drug delivery mechanism of nanoparticles. (B) Characterization of PhoeA-SN38-HC NPs: NIR induced singlet oxygen generation from NPs. In the presence of DMA (100 μM), 1 mg/mL NPs were exposed to light, and fluorescence intensity (λ = 420 nm) of DMA was measured by spectrometer, Light induced drug release from NPs depending on light energy. (C) In vivo PDT treatment with HC-PheoA-SN38: body weights and tumor growth curves of HEY-T30 xenograft BALB/C nude mouse, * p < 0.05 and *** p < 0.001. Reprinted with permission from [164]. Copyright@ Elsevier. (D) The preparation of dual drugs-conjugated PDOXCBs nanoparticles. (E) Both PDOXCB18 and PDOXCB27 cytotoxicity against cancer cells. Reprinted with permission from Ref. [167]. 2021, Elsevier.
Pharmaceutics 14 01522 g007
Figure 8. (A) Preparation of ApMDC and their self-assembled nanomicelles with tunable surface density of aptamers, initiation of antitumor immune responses of checkpoint blockade therapy by tumor-targeting, yet enhanced, chemotherapy. (B) In vivo antitumor immune responses: quantitative analysis of tumor infiltrating CD4+, CD8+, and level of Ki67 in the tumor-draining lymphoid node after treatment with PBS, free DOX, 0ApMDC, 40ApMDC, 40ApMDC + a-PD1 and a-PD1, respectively, * p < 0.05, ** p < 0.01 and *** p < 0.001. Reprinted with permission from Ref. [171]. 2021, Wiley-VCH GmbH. (C) Schematic illustration of IL-12 nanostimulant-engineered CAR T cells (INS-CAR T) biohybrids with immunofeedback to enhance immunotherapy in solid tumors. The CAR T cell-mediated INS delivery system elicited CAR T cell infiltration and enhanced immune responses. (D) IL-12 accumulation in tumor of NOD/SCID mice at 48 h post last administration and the average CAR T cell number in each field (600 μm × 600 μm) of tumor tissues was calculated from 50 fields, ** p < 0.01 and *** p < 0.001. (E) Representative flow cytometry analysis of the ratios of the percentages of CD8+ CAR T cells and coexpression of CD25 and Foxp3 among CD4+ CAR T cells in tumor tissues. Reprinted with permission from Ref. [176]. 2022, Elsevier.
Figure 8. (A) Preparation of ApMDC and their self-assembled nanomicelles with tunable surface density of aptamers, initiation of antitumor immune responses of checkpoint blockade therapy by tumor-targeting, yet enhanced, chemotherapy. (B) In vivo antitumor immune responses: quantitative analysis of tumor infiltrating CD4+, CD8+, and level of Ki67 in the tumor-draining lymphoid node after treatment with PBS, free DOX, 0ApMDC, 40ApMDC, 40ApMDC + a-PD1 and a-PD1, respectively, * p < 0.05, ** p < 0.01 and *** p < 0.001. Reprinted with permission from Ref. [171]. 2021, Wiley-VCH GmbH. (C) Schematic illustration of IL-12 nanostimulant-engineered CAR T cells (INS-CAR T) biohybrids with immunofeedback to enhance immunotherapy in solid tumors. The CAR T cell-mediated INS delivery system elicited CAR T cell infiltration and enhanced immune responses. (D) IL-12 accumulation in tumor of NOD/SCID mice at 48 h post last administration and the average CAR T cell number in each field (600 μm × 600 μm) of tumor tissues was calculated from 50 fields, ** p < 0.01 and *** p < 0.001. (E) Representative flow cytometry analysis of the ratios of the percentages of CD8+ CAR T cells and coexpression of CD25 and Foxp3 among CD4+ CAR T cells in tumor tissues. Reprinted with permission from Ref. [176]. 2022, Elsevier.
Pharmaceutics 14 01522 g008
Table 1. Nanomedicines for cancer treatment with granted regulatory approval [10,11,12,13,14,15].
Table 1. Nanomedicines for cancer treatment with granted regulatory approval [10,11,12,13,14,15].
Trade NameActive PrincipleNanotechnology PlatformIndicationApproved Status
Doxil/CaelyxDoxorubicinPEGylated liposomesBreast cancer, ovarian cancer, myelomaFDA 1, EMA 2
DaunoXomeDaunorubicinLiposomesKaposi sarcomaFDA
MyocetDoxorubicinLiposomesBreast cancerFDA
LipusuPaclitaxelLiposomesbreast cancer, non-small-cell lung cancerNMPA 3
AbraxanePaclitaxelAlbumin-bound nanoparticlesMetastatic breast cancer, metastatic pancreatic cancer, advanced non-small-cell lung cancerFDA, EMA
Genexol-PMPaclitaxelPolymeric micellesNon-small-cell lung cancerKFDA 4
MEPACTMifamurtideLiposomesOsteosarcomaEMA
MarqiboVincristineLiposomesPhiladelphia chromosome-negative acute lymphoblastic leukemiaFDA
PICNPaclitaxelPolymer/lipid nanoparticlesMetastatic breast cancerIndia
Onivyde (MM-398)IrinotecanPEGylated liposomesMetastatic pancreatic cancerFDA
VYXEOSCytarabine/daunorubicin (5:1)LiposomesAcute myeloid leukaemiaFDA, EMA
PaclicalPaclitaxelPolymeric micellesOvarian cancerRussia
HensifyN/AHafnium oxide NPLocally advanced soft tissue sarcomaEMA
DHP107PaclitaxelLipid nanoparticlesAdvanced gastric cancerKFDA
NanoThermN/A Iron oxide nanoparticlesRecurrent glioblastomaEMA
NanoxelPaclitaxelPolymeric micelles Breast cancer, ovarian cancerIndia
DepocytCytarabineLiposomesAcute Nonlymphocytic Leukemia, Meningeal Leukemia, Lymphomatous MeningitisFDA
1 FDA: Food and Drug Administration. 2 EMA: European Medicines Agency. 3 NMPA: National Medical Products Administration. 4 KFDA: Korean Food and Drug Administration.
Table 2. ADCs on the market.
Table 2. ADCs on the market.
Trade NameADCTarget AntigenIndicationApproved Status
MylotargGemtuzumab ozogamicinCD33CD33 positive AMLFDA
AdcetrisBrentuximab vedotinCD30Hodgkin lymphoma and anaplastic large cell lymphomaFDA
KadcylaAdo-trastuzumab emtansineHER2HER2-positive breast cancerFDA
BesponsaInotuzumab ozogamicinCD22Relapsed or refractory B cell precursor acute lymphoblastic leukemiaFDA
LumoxitiMoxetumomab pasudotoxCD22Relapsed or refractory B-cell precursor acute lymphoblastic leukemiaFDA
PolivyPolatuzumab vedotinCD79bRelapsed or refractory diffuse large B cell lymphomaFDA
PadcevEnfortumab vedotinNectin-4Advanced or metastatic urothelialFDA
EnhertuTrastuzumab deruxtecanHER2Relapsed or refractory diffuse large B cell lymphomaFDA
TrodelvySacituzumab
govitecan
Trop-2HER2-triple-negative breast cancerFDA
BlenrepBelantamab mafodotinBCMARelapsed or refractory multiple myelomaFDA
AkaluxCetuximab IRDye700DXEGFRHead and neck tumors, esophageal tumors, lung tumors, colon cancersPMSB 1
ZvniontaLoncastuximab tesirineCD19Relapsed or refractory diffuse large B cell lymphomaFDA
AidixiDisitamab vedotinHER-2HER-2 positive metastatic gastric cancerNMPA
TivdakTisotumab vedotin-tftvTFRelapsed or metastatic cervical cancerFDA
1 PMSB: Pharmaceutical and Medical Safety Bureau (Japan).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, B.; Chen, S.; Hong, Y.; Jia, L.; Zhou, Y.; He, X.; Wang, Y.; Tian, Z.; Yang, Z.; Gao, D. Research Progress of Conjugated Nanomedicine for Cancer Treatment. Pharmaceutics 2022, 14, 1522. https://doi.org/10.3390/pharmaceutics14071522

AMA Style

Zhao B, Chen S, Hong Y, Jia L, Zhou Y, He X, Wang Y, Tian Z, Yang Z, Gao D. Research Progress of Conjugated Nanomedicine for Cancer Treatment. Pharmaceutics. 2022; 14(7):1522. https://doi.org/10.3390/pharmaceutics14071522

Chicago/Turabian Style

Zhao, Bin, Sa Chen, Ye Hong, Liangliang Jia, Ying Zhou, Xinyu He, Ying Wang, Zhongmin Tian, Zhe Yang, and Di Gao. 2022. "Research Progress of Conjugated Nanomedicine for Cancer Treatment" Pharmaceutics 14, no. 7: 1522. https://doi.org/10.3390/pharmaceutics14071522

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