Next Article in Journal
A Single Dose of a Hybrid hAdV5-Based Anti-COVID-19 Vaccine Induces a Long-Lasting Immune Response and Broad Coverage against VOC
Next Article in Special Issue
Anti-IAPP Monoclonal Antibody Improves Clinical Symptoms in a Mouse Model of Type 2 Diabetes
Previous Article in Journal
Flu and Tdap Maternal Immunization Hesitancy in Times of COVID-19: An Italian Survey on Multiethnic Sample
Previous Article in Special Issue
Human Transbodies to Reverse Transcriptase Connection Subdomain of HIV-1 Gag-Pol Polyprotein Reduce Infectiousness of the Virus Progeny
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antibody-Drug Conjugates: Functional Principles and Applications in Oncology and Beyond

First Department of Medicine, School of Medicine, National and Kapodistrian University of Athens, Laiko General Hospital, 115 27 Athens, Greece
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 25 August 2021 / Revised: 23 September 2021 / Accepted: 24 September 2021 / Published: 29 September 2021
(This article belongs to the Collection Research on Monoclonal Antibodies and Antibody Engineering)

Abstract

:
In the era of precision medicine, antibody-based therapeutics are rapidly enriched with emerging advances and new proof-of-concept formats. In this context, antibody-drug conjugates (ADCs) have evolved to merge the high selectivity and specificity of monoclonal antibodies (mAbs) with the cytotoxic potency of attached payloads. So far, ten ADCs have been approved by FDA for oncological indications and many others are currently being tested in clinical and preclinical level. This paper summarizes the essential components of ADCs, from their functional principles and structure up to their limitations and resistance mechanisms, focusing on all latest bioengineering breakthroughs such as bispecific mAbs, dual-drug platforms as well as novel linkers and conjugation chemistries. In continuation of our recent review on anticancer implication of ADC’s technology, further insights regarding their potential usage outside of the oncological spectrum are also presented. Better understanding of immunoconjugates could maximize their efficacy and optimize their safety, extending their use in everyday clinical practice.

1. Introduction

Antibody-drug conjugates (ADCs) comprise a fast-expanding therapeutic modality designed to target disease cells, sparing the adjacent healthy tissues. The regulatory approval of first-generation ADCs spurred a surge of interest in this biotechnology that has produced a total of ten FDA-approved agents and an ever-increasing panel of candidates in clinical and preclinical level. ADCs are essentially tripartite pro-drugs consisting of an antibody tethered via a chemical linker to a given payload [1]. (Figure 1) After their administration, these agents circulate as inactive assemblies which are eventually catabolized via endogenous cleavage mechanisms at the intracellular compartment of the targeted cell [2]. Exploiting the embedded properties of monoclonal antibodies (mAbs), these immunoconjugates achieve selective delivery and localized release of the attached payload, minimizing the “off-target” effects on normal tissues and improving the therapeutic index [3]. The clinical utility of ADCs has been mainly explored in hematological/oncological indications [1]. However, during the last decade, a constant flow of reports about the implementation of this technology outside the oncological sphere has been also observed. In fact, brentuximab vedotin (BV), ABBV-3373, and DSTA4637S are currently under clinical testing for non-oncological indications, such as autoimmune and infectious diseases [4,5,6]. The discrepancy between the initially large number of ADCs in pharmaceutical pipelines and the small proportion of agents reaching late-stage trials highlights the need for deeper understanding, meticulous selection, and constant optimization of ADC components. In the following sections, we elaborate on the functional principles and the characteristics of these components (Box 1), with emphasis on the latest bioengineering advances, such as bispecific antibodies, multi-drug ADCs, non-internalizing ADCs, and ADC-antibody co-administration. Updating our previous report [1], all current oncological and non-oncological implications of ADC’s technology are also recapped.
Box 1. Essential components and considerations in the design of ADCs.
Essential components of ADCS
1 
ANTIBODIES
TypesConsiderations for AbsConsiderations for target antigen
  • Monoclonal Ab (chimeric, humanized or human, mainly IgG1 isotype)
  • Bispecific Ab
  • Specificity and affinity for the target antigen
  • Internalization
  • Pharmacokinetic properties
  • Effector functions
  • Immunogenicity
  • Tumour-specificity
  • Internalization
  • Expression pattern
  • Ectodomain shedding
2 
LINKERS
TypesConsiderations for linkerConsiderations for conjugation
  • Cleavable linkers
  • Non-cleavable linkers
  • Circulation stability
  • Solubility
  • Aggregation propensity
  • Drug to antibody ratio (DAR)
  • Homogeneity
3 
PAYLOADS
TypesConsiderations for payloadMechanisms of drug resistance
  • Calicheamicins
  • Pyrrolobenzodiazepines (PBDs)
  • Auristatins
  • Maytansinoids
  • Camptothecin (CPT)
  • Half life
  • Bystander killing activity
  • Systemic accumulation
  • Potential for resistance development
  • Antigen-related resistance
  • Deficient lysosomal function
  • Upregulated efflux pumps
  • Survival/apoptotic signaling
  • BSB phenomenon

2. Functional Principles and Essential Components of ADCs

2.1. Antigen Selection and Antibodies

The fundamental step of ADC development remains the selection of target-antigen [7,8]. The optimal target-antigen should be characterized by tumor-specific and homogeneous expression pattern, high levels of expression, rapid internalization, and minimal ectodomain shedding. The exclusive expression of an epitope on the cancer cell surface represents an ideal scenario, since most selected antigens are, in fact, tumor-associated and not tumor-restricted [9]. In general, it is preferable for these tumor-associated antigens to be localized either on tissues resistant to the given payload or on tissues with high regenerative capacity [10]. The level of antigen expression critically affects the therapeutic index of the immunoconjugate as it defines the amount of the cytotoxic payload that will be internalized in the cancer cell [11]. In solid tumor cell lines, this correlation between surface antigen density and intracellular ADC concentration reaches an almost linear relationship (R2 ≥ 0.91) [12]. Regarding the internalization of the ADC-antigen complex, cleavable linker-based ADCs with membrane-permeable payloads seem to be less dependent on the trafficking of the antigen [10]. Such immunoconjugates are able to exert their cytotoxic activity after extracellular cleavage and subsequent local drug diffusion [13,14]. To this direction, several inert constituents of tumor microenvironment have been tested as stromal ADC-targets [15,16] and recently, a non-internalizing ADC was developed detaching its payload after consequent administration of a linker-activator, independently of endogenous cleavage [17]. Notably, this immunoconjugate was proved potent even against murine models insensitive to the FDA-approved ADC, BV. Minimal shedding should be also added in the list of beneficial antigen features, as secreted epitopes can bind to the circulated immunoconjugates and render them ineffective. This off-target antigen-ADC interaction may lessen the portion of the administered drug reaching into the tumor microenvironment and lead to unnecessary dose escalation [18]. The introduction of bispecific antibodies can help to overcome this technical issue [19]. However, in a modelling study, antigen shedding in solid malignancies may act positively by preventing binding site barrier (BSB) phenomenon and facilitating a more homogeneous distribution of ADC [20]. Therefore, these considerations about BSB phenomenon and bystander killing effect are challenging internalization and shedding as strict properties of ADC target [14].
Depending on the selected immunoglobulin subtype, the antibody component of ADC retains both targeted transport and cell-killing potential [21]. The main characteristics of candidate mAbs include high affinity for the target antigen, rapid internalization, favorable pharmacokinetic properties and minimal immunogenicity. The issue of immunogenicity has significantly improved with the introduction of human or humanized immunoglobulins [22]. Moreover, these antibodies interact better with both immune cells and complement system [23]. In ADC-engineering, most mAbs (chimeric, humanized or human) belong to IgG1 isotype while IgG3 isotype is not utilized [24]. The IgG3 isotype is cleared up to three times faster than IgG1, IgG2, and IgG4 (half-life of 7 days compared to 21 days) and is more sensitive to proteolysis, increasing the risk of immunogenicity [25,26]. The IgG1 isotype is usually employed because of its high affinity for all Fc-gamma receptors and its ability to induce secondary immune functions, antibody-dependent cellular cytotoxicity and complement-dependent toxicity [27]. On the other hand, IgG2 and IgG4 isotypes are poor inducers of the complement cascade [28]. The sub-nanomolar levels of mAbs prevent cross-reaction with off-target antigens, limit systemic toxicity and premature elimination [29], while the quick and tight binding to antigen enables rapid internalization and efficient payload delivery. However, similar to high antigen expression, high affinity can increase binding and rapid endocytosis of the prodrug within the first cancer cells, lowering the therapeutic efficacy of ADCs (BSB phenomenon) [30].
The bispecific ADCs (bsADCs) have been manufactured to identify a pair of different antigens or two distinct epitopes on the same antigen (known as biparatopic), conferring a more target-specific drug delivery compared to monospecific ADCs. A wide range of bispecific antibodies compositions have been described, with more than 100 formats reported in the literature [31,32]. Based on their robust selectivity, bsADCs improve the safety profile of conventional ADC formats and upgrade their applicability. Simultaneous engagement of co-expressed target antigens or non-overlapping epitopes allows an accumulated on-target toxicity, and diminishes uptake by adjacent healthy tissues [33]. Furthermore, bsADCs have been manipulated to enhance internalization and redirect lysosomal trafficking and degradation. The bsADCs can utilize strongly internalizing receptors to overcome suboptimal lysosomal uptake and limited drug exposure attributed to increased recycling of prone antigens such as HER2 [34]. Cross-linking between two molecules (e.g., a high-turnover membrane protein and a tumor-marker antigen), irrespective the affinity of mAb can accelerate subsequent downstream cascade. For instance, CD63 was proven to facilitate transmission from membrane to intracellular compartments. The identification of this high-yield molecular “shuttle” led to generation of anti-HER2/CD63, duostatin-3-linked bsADC. This bsADC displayed greater anti-tumor activity compared to monovalent HER2- and CD63 ADCs [35]. Prolactin receptor (PRLR) was pinpointed as another candidate target-antigen due to its rapid internalization and lysosomal delivery. The produced anti-HER2/PRLR bsADC not only boosts trafficking of HER2, but also displays greater activity than single HER2 ADC or PRLR ADC in breast cancer cells with intermediate HER2 and low PRLR levels [36]. Similarly, cross-linking of HER2 with the rapidly internalizing APLP2 receptor in DM1-linked bsADCs has also demonstrated promising findings in terms of potency, compared to single T-DM1 [37]. Another bsADC model was recently developed for CD7+/CD33+ acute myeloid leukemia (AML). Co-targeting of CD7 and CD33 increases the specificity of the immunoconjugate, compared to gemtuzumab ozogamicin (GO, approved anti-CD33 ADC), and augments cytotoxicity against CD7 + CD33+ cells both in vitro and in vivo [38]. Lastly, biparatopic platforms have been used to enhance specificity and maximize internalization and trafficking, as clustering and cross-linking of receptors accelerates intracellular trafficking. For this objective, Li et al. developed a biparatopicADC (MEDI4276), combining two non-overlapping HER2-targeted Abs, with activity across different levels of HER2 expression [39]. MEDI4276 is currently under investigation in a first-in-human clinical trial in patients with HER2+ breast or gastric cancer (NCT02576548) [40].

2.2. Linkers and Conjugation Technologies

Linkers have been designed to tether the cytotoxic molecule to the antibody scaffold, regulating several prodrug parameters such as circulation stability, solubility, and aggregation propensity. These components can generally be categorized into cleavable and non-cleavable ones, depending on whether they can be degraded or not. Cleavable linkers can be (i) acid-labile (e.g., hydrazones), (ii) reducible/glutathione-sensitive (e.g., disulfides), and (iii) protease-sensitive/peptide linkers. As an example of the first subgroup, Gemtuzumab ozogamicin (GO) employs a bifunctional 4-(4-acetylphenoxy) butanoic acid part attached to the calicheamicin payload via hydrazone linkage. This type of linker remains stable at normal blood pH but undergoes hydrolysis under lysosomal and endosomal acidic conditions [41] and possibly elsewhere in the body where pH is also low, resulting in undesired, nonspecific drug release. In 2010, this drawback led to the temporary withdrawal of GO, due to marked toxicity, attributable to linker instability [42]. In the second category, disulfide linkers exploit the transmembrane difference in reductive potential, owing to considerably higher intracellular concentrations of reducing agents [43]. Protease-sensitive/peptide linkers usually consist of oligopeptide substrates, most commonly dipeptide valine-citrulline (Val-Cit) combined with a self-immolative para- amino-benzyloxycarbonyl (PABC) spacer. This type of linker responds to overexpressed lysosomal proteases in cancer cells such as cathepsin B [44]. This last linker category combines well-established release patterns and improved drug control, and is already employed in approved ADC, BV [45,46]. Additional cleavable linker formats include β-glucuronide linkers, which are hydrophilic linkers responsive to β-glucuronidase, present in lysosomes and tumor necrotic areas. The hydrophilic composition of linker provides adequate polarity and stability and solubilizes typically hydrophobic payloads [47]. This masking through hydrophilic monodisperse polysarcosine drug-linker configuration (PSARlink) sensitive to β-glucuronidase, was applied in the production of novel high-DAR ADCs, against T-DM1-resistant cells [48]. Non-cleavable linkers display greater stability than cleavable ones, remaining intact through proteolytic, acidic and reductive conditions. ADCs with non-cleavable linkers depend their cytotoxic effects on the degradation of mAb scaffold. More specifically, the active metabolite is released into cytoplasm upon complete antibody breakdown leaving only a drug-linker-amino acid part [46,49]. Of note, amino acid capping increases hydrophilicity and reduces membrane permeability, influencing the bystander effect [50]. Examples of non-cleavable linkers include non-reducible thioether, as in T-DM1 [29]. Further advances on linker technology have been developed in order to optimize and expand ADCs’ utility. For instance, cleavable pyrophosphate-diester linkers in site-specific, glucocorticoid-bearing ADCs outside of oncological setting [51]; β-galactosidase-cleavable linkers for trastuzumab-MMAE conjugates [52] and dual enzyme-cleavable linkers (e.g., 3-O-sulfo-β-galactose linker), subjects of sequential cleavage by distinct lysosomal enzymes (e.g., arylsulfatase A and β-galactosidase) have demonstrated encouraging findings [53]. In the construction of dual-drug ADCs, flexible linkers contribute in the successful co-delivery of payloads with synergistic cytotoxic mechanisms [54]. Recently, Spangler et al. described a novel linker format, named as Fe(II)-reactive 1,2,4-trioxolane (TRX), which reacts with labile ferrous iron in cancerous tissue to induce a more tumor-selective drug release [55]. This TRX-linker limits the on-target-off-tumor toxicity upon uptake by adjacent healthy cells. A first-in-class platinum (II)-based metal-organic linker (Lx) was designed to surpass conventional linkage pitfalls, such as premature release. In preclinical studies, Lx-based ADCs have shown favorable safety profile and potency [56]. At the end, additional linkers including non-covalent DNA linkers (e.g., based on complementary oligonucleotide hybridization and base-pairing) [57] as well as photo-cleavable linkers on UV light-controlled ADCs [58], are also under clinical testing.
Conjugation affects the ADC stoichiometry and homogeneity, crosslinking the cytotoxic drug-linker moiety to the antibody vehicle. The conjugation strategy dictates the quantity of drug molecules attached per antibody, defined as drug-to-antibody ratio or DAR. Of note, broader DAR distribution produced more heterogenous ADC, which results in product inconsistency and suboptimal efficiency [46]. Conventional conjugation techniques utilize intrinsically nucleophilic side-chain groups of solvent-accessible amino-acid residues in the mAb component, with native lysine and cysteine residues being the most frequently detected. Utilizing native lysine residues can lead to highly heterogenous ADC species because of their relative abundance (>80) in a typical IgG molecule and the wide range of possible conjugation spots. Regarding native cysteines, in IgG1 they form 16 pairs; 4 interchain and 12 intrachain disulfide bridges. Cysteine-based conjugation is based on reduction of the interchain cysteines and can spawn up to eight sulfhydryl(-SH) groups, thus yielding DARs ≤ 8. The nucleophilic sulfhydryl groups consequently can be reacted with electrophilic entities to allow conjugation through various chemical reactions [45,49]. Recently, efforts have been concentrated on more homogenous drug loading and better-controlled DARs. Disulfide re-bridging can be applied without requiring recombinant or enzymatic modifications [59]. Selective mutations of the mAb amino-acid sequence enable site-specific conjugation via recombinant incorporation of reactive handles. THIOMAB™ technology introduces two genetically engineered, unpaired cysteines, spares interchain disulfides, and permits selective attachment. In vivo studies have shown improved therapeutic window and tolerability of developed ADCs with sustained DAR ~ 2 and high homogeneity (>90%) [60]. Another site-specific conjugation approach is based on the introduction of unnatural amino-acids (uAA) with reactive side-chains for chemical tethering. Installation of the non-canonical residues is feasible through recombinant technology [61,62]. Successful enzymatic ligation techniques have also been used for site-selective bioconjugation of native or engineered mAbs with attractive partners. Microbial transglutaminase (mTG) catalyzes the formation of isopeptide bond between linker and glutamine in de-glycosylated mAbs, without modifying native glutamines [63,64]. In addition, short glutamine motif (LLQG) can be inserted into mAbs, rendering them fit for peptide sequence-specific linking via transpeptidation in presence of mTG. This technique generates ADCs with strictly controlled DARs and favorable profiles compared to traditional ADCs [65]. Interestingly, mTG-mediated ligation has also been used for branched linkers in order to flexibly increase payload loading of ADCs, without further intervening in the mAb structure to accommodate multiple individual linkers [66]. Moreover, microbial sortase A (SortA) can recognize a specific pentapeptide tag (LPETG) appended to C-terminus of recombinant mAbs and mediate the conjugation. This sortase-mediated antibody conjugation technology (SMACTM) can be used to efficiently yield site-specific homogenous ADCs with predefined DARs that retain the tumor killing properties of traditional ADCs [67]. SmartTagTM (Specific Modifiable Aldehyde Recombinant Tag) is another enzyme-assisted platform that utilizes a formylglycine-generating enzyme (FGE) to recognize a CxPxR (X: serine, threonine, alanine, or glycine) tag and convert cysteine to formylglycine residues bearing a reactive aldehyde group. Localized bioconjugation is achieved by modifying the mAb and selectively inserting the FGE-recognized sequences at intended sites [68]. Other remodeling techniques are focused on mAbs’ glycosylation. These strategies can modify the conserved N-glycan chain of Fc domain to allow conjugation [69], or can incorporate reaction handles via placement of non-natural saccharides [70].

2.3. Payloads

In cancer settings, the employed warhead is a super-toxic compound potent in sub-nanomolar concentrations and, thus, intolerable if administered unconjugated. The ADCs’ payloads are categorized into two groups: DNA-damaging agents and microtubule-disrupting agents. DNA-damaging drugs are further divided into three subcategories: DNA-double strand break inducing agents, DNA alkylators, and DNA intercalators. Microtubule-disrupting agents are the most widely used cytotoxic agents in ADC technology. Maytansinoids and auristatins are the main representatives. A plethora of other drug classes targeting other cellular processes are investigated as potential ADC payloads. The payload classes of FDA-approved ADCs as well as the attached drugs for non-oncologic ADCs are discussed below.

2.3.1. Calicheamicins

Calicheamicins are a class of natural anticancer antibiotics isolated from the actinomycete Micromonospora echinospora spp. Calichensis [71]. These compounds exhibit site-specific binding in the minor groove of DNA. The primary recognition site is TCCT/AGGA. Subsequently, reductive cleavage by cellular thiols generates a diradical species that abstracts hydrogen atoms from DNA inducing strand scission and cell death [72]. A semisynthetic derivative of chalicheamicin, N-acetyl-gamma calicheamicin 1,2-dimethyl hydrazine, is already used in two FDA-approved ADCs, inotuzumab ozogamicin and gemtuzumab ozogamicin. Although potent, these ADCs have certain limitations such as increased aggregation and shortened half-life, due to the employed linker chemistry. Next-generation calicheamichin-based ADCs exhibit these features to a lesser degree due to novel site-specific conjugation [73].

2.3.2. Pyrrolobenzodiazepines (PBDs)

Pyrrolobenzodiazepines (PBDs) bind on sequence-specific fragments of opposite DNA strands and stop their separation during the cell cycle (e.g., G2/M boundary), inducing cell death [74]. Importantly, PBD dimers can permeate membranes and have a very short half-life. Therefore, these drugs exert bystander killing activity with limited systemic accumulation. Furthermore, they exhibit small tendency for acquired resistance and technically reversible alkylation after DNA digestion and heating [75,76]. The latter characteristic enables the quantitation of PBD compound and isolated DNA, comparing payload exposure in tumor and normal tissue, by the levels of calculated alkylation [62]. This ratio accurately represents the safety and efficacy profile of PBD-containing ADC. The latest FDA-approved ADC, loncastuximab tesirine, employs tesirine as its PBD warhead [77].

2.3.3. Auristatins

Auristatins comprise synthetic analogues of the natural cytotoxic product Dolostatin 10, isolated from Dolabella auricularia. These drugs block mitosis via inhibition of tubulin polymerization, and thus lead to apoptosis [78]. Two synthetic auristatin derivatives, monomethyl auristatin E (MMAE) and monomethyl auristatin F (MMAF), have been studied in great detail. MMAF is less effective compared to MMAE because of the charged phenylalanine in its C-terminal that presumably hinders intracellular access [79]. However, MMAF is highly potent once it has reached its target being one of the most toxic auristatins generated. In contrast to MMAE, MMAF does not exert bystander killing activity. A novel auristatin payload, auristatin F-hydroxypropylamide (AF-HPA), has been recently reported in the context of the pioneering ADC platform [80]. Interestingly, the membrane permeating AF-HPA undergoes intracellular conversion to membrane non-permeating MMAF, resulting in a controlled bystander effect. This characteristic decreases the rate of neutropenia as a dose-limiting toxicity of these ADCs compared to other auristatin platforms [81]. Four FDA-approved ADCs contain auristatins as their attached payload.

2.3.4. Maytansinoids

Maytansinoids are derivatives of maytansine which is a benzoansamacrolide isolated from Maytenus ovatus. Maytansinoids act as antimitotic agents, binding to tubulin at or near the vinca binding site. Tubulin binding destabilizes the microtubule assembly and induces cell cycle arrest at the G2/M phase [2]. Maytansinoids are hydrophobic and upon extracellular or intracellular cleavage they can effectively diffuse into antigen-negative cells. Trastuzumab emtasine (T-DM1) utilizes DM1, a thiol-containing maytansine derivative, as its warhead.

2.3.5. Camptothecin (CPT)

Camptothecin (CPT) is pentacyclic quinoline alkaloid originally isolated from the plant species Camptotheca acuminata. CPT and its analogues inhibit topoisomerase-1 activity. CPT interacts with topoisomerase-1 and DNA interface through hydrogen bonding, and form reversible complexes, blocking DNA replication and inducing cell death [82]. SN-38, an active metabolite of irinotecan, and exatecan, a water-soluble CPT derivative, are the payloads of FDA-approved sacituzumab govitecan and trastuzumab deruxtecan, respectively.

2.3.6. Dual-Drug ADCs

Dual-drug ADCs co-deliver different payloads with complementary or synergistic cytotoxic mechanisms at equimolar concentrations [83]. This co-administration of anti-cancer agents with multi-loading linkers and dual payloads can counteract tumor resistance. Data have shown that cancer cells, resistant to an ADC, remain sensitive to the alternate payload delivered via the same mAb [84]. Levengood et al. reported an homogeneous dual-auristatin ADC carrying two tubulin polymerization inhibitors, MMAE and MMAF [85]. This novel ADC was potent against xenograft models of anaplastic large cell lymphoma refractory to monotherapy with either of the individual drugs, further supporting the development of this concept. To this end, another dual-drug ADC carrying two mechanistically different warheads was subsequently described. This approach combined MMAE and a PBD-dimer and exerted two distinct cytotoxic mechanisms consistent with the tethered payloads [54]. However, this PBD-dimer/MMAE-ADC failed to achieve enhanced cell-killing efficacy compared to the ADC equipped with the PBD-dimer alone, probably because of the great potency of the latter payload. A comparable ADC design carrying a DNA alkylator and a microtubule inhibitor was published by Duvall et al. [86]. Similar to Kumar et al. [54], Nilchan et al. developed a dual-drug ADC (PNU-159682/MMAF) against HER2(+) cell lines that did not exhibit greater potency compared to single-PNU-159682 ADC [87]. MMAE and MMAF were also combined in a dual-drug ADC format for HER2(+) breast cancer lines [88]. Interestingly, this construct was more efficient than both monotherapy with single-drug ADCs and administration of the two variants, as MMAF-induced cell death of HER2(+) cells enabled MMAE bystander killing activity.

3. Bystander Killing and Resistance Phenomena

3.1. Bystander Killing Effect

The bystander killing effect is exhibited when the released cytotoxic drug of ADC is unleashed into surrounding antigen-negative tumor and/or normal cells. The naked payload, in cleavable linkers, or more rarely, the drug-linker-amino acid fraction in non-cleavable linkers, can diffuse through the phospholipid bilayer towards the nearby cells [89]. A critical parameter affecting the extent of the phenomenon is the membrane permeability of the bioactive form of the drug. Charged and hydrophilic ADC drug derivatives are known to demonstrate a minimal bystander effect, while more hydrophobic and neutrally charged payload catabolites experience the effect to a greater degree [50,90]. The bystander activity can be quantified in vitro by incubating the ADC agent with co-cultures of cancer cells with a certain number of antigen negative cells and varying numbers of antigen positive cells. The cytotoxic potency of tethered payload is determined by the number of antigen positive cells required to kill the antigen negatives [91,92]. In an attempt to further quantify and predict the clinical impact of the bystander effect, several pharmacokinetic and pharmacodynamic modelling approaches have been recently proposed [93,94,95], but this issue remains under debate. To a certain degree, the bystander killing effect can be exploited to tackle tumor heterogeneity. In this case, the effect of the ADC-attached drug depends less on the homogenous expression of the target antigen, because its surface expression on all tumor cells is not required [96]. The FDA-approved ADC, Trastuzumab deruxtecan, serves as a great example of how the bystander killing mechanism can be translated into therapeutic benefit [97]. More specifically, in vivo and in vitro preclinical testing in cell lines with varying HER2 expression profile revealed notable antineoplasmatic activity against every antigen expression density [98]. On the other side, the bystander effect when experienced on normal neighboring cells can contribute to undesirable off-target toxicity, hampering the safety profile of administered immunoconjugate. Modern auristatin platforms, such as upifitamab rilsodotin (XMT-1536), with low bystander killing effect are causing less neutropenia compared to traditional auristatin-containing ADC [81]. It is worth-noting that the bystander effect was the functional basis for the development of non-internalizing ADCs [14].

3.2. Drug Resistance

Similar to other anti-cancer drugs, many different resistance mechanisms have been developed by tumor cells to overcome ADC-based treatments and are listed below.

3.2.1. Antigen-Related Resistance

Downregulation of antigen expression is a major mechanism of drug resistance that usually develops in cells chronically exposed to the ligand [99]. Breast cancer cell lines, 361-TM and JIMT1-TM, exposed to multiple cycles of an anti-HER2 trastuzumab–maytansinoid ADC (TM-ADC) expressed 25% and 58% reduction of binding ability, respectively [84]. Masking of the targeted epitope contributes to additional antigen-related resistance [100].

3.2.2. Deficient Lysosomal Function

Impaired lysosomal proteolytic activity delays ADC degradation, limiting its cell-killing efficacy. In T-DM1 insensitive clones, although no alterations were detected in endocytosis and intracellular trafficking, the proteolytic activity within the lysosomes was deficient due to the increased pH [101]. Aberrant vacuolar H+-ATPase (V-ATPase) activity has been observed in resistant cells, while administration of the V-ATPase inhibitor, bafilomycin A1, successfully sensitize again cell lines [102]. Another report proposed the use of photoactivatable nanoparticles to manipulate lysosomal pH levels and restore the anti-tumoral activity of ADCs [103]. At this level, cells’ resistance can also be induced by disruption of transportation through the lysosomal membrane to the cytosol [104]. Especially in cases of non-cleavable linked ADCs, silencing of lysosomal transporters can cause intralysosomal accumulation of the non-cleavable catabolite, decreasing ADC activity [105].

3.2.3. Upregulated Efflux Pumps

The efflux of bioactive payload via ATP binding cassette (ABC) transporters, such as MDR1/PgP, is another mechanism of acquired resistance. Notably, the majority of commonly used ADC payloads are substrates of efflux transporters [100]. Indeed, Chen et al. observed that MMAE, employed by BV, was actively exported in HL cells resistant to BV [106]. Subsequent in vivo studies proved that PgP inhibition could restore sensitivity to BV in BV-resistant HL cell lines [107]. In a phase I clinical trial the co-administration of BV with cyclosporine A (MDR-modulator) achieved responses even in BV-refractory HL patients [107]. In another example, Loganzo et al. detected increased ABCC1 (MRP1) as possible resistance mediator in 361-TM cells and intriguingly showed that switching a non-cleavable linker for a cleavable one could restore sensitivity to TM-ADC [84].

3.2.4. Survival/Apoptotic Signaling

Alterations in survival signaling pathways can also modulate crucially ADC cytotoxicity [99,100,104]. Elevated PI3K/AKT/mTOR pathway activity has been associated with acquired resistance to GO in AML cells [108] via overproduction of anti-apoptotic factors [99]. PI3K/AKT/mTOR pathway inhibition via MK-2206, an AKT inhibitor, or via PP242, a mTOR1/2 inhibitor, resulted in re-sensitization of resistant cells to GO [108,109]. Furthermore, depletion of PTEN expression, a negative PI3K pathway regulator and aberrant activation of STAT3 were also correlated with resistance to T-DM1 [110,111]. Lastly, upregulation of anti-apoptotic proteins BCL-2/BCL-X induces important resistance [112] while their inhibition was shown to restore sensitivity to T-DM1 [113].

3.2.5. Binding-Site Barrier (BSB) Phenomenon

The binding-site barrier (BSB) phenomenon is defined as the binding-dependent penetration and non-uniform distribution of the ADC into the tumor microenvironment [114]. The extensive cellular uptake in the perivascular/peritumoral regions reduces the penetration and homogeneous diffusion into the main mass [115], affecting the cytotoxic effect of particular ADC that do not exhibit effective bystander killing activity [116]. Antigen shedding [20], sizing and affinity alterations [117,118], co-administration of molecules interfering with binding [119] and, more recently, dosing modifications [120] have been described as strategies to attenuate the clinical impact of BSB. Cilliers et al. reported that co-administration of trastuzumab emtasine (T-DM1) with trastuzumab effectively modifies distribution of T-DM1 in HER2(+) NCI-N87 tumor cells. This approach was tested using two different trastuzumab-based ADCs exhibiting different bystander activity, at two different doses and in two different HER2(+) cancer models [114]. Co-administration approach was more efficient in the case of ADC without bystander killing ability in tumor models with very high antigen surface density. In gynecological malignancies, higher doses of mesothelin-targeting ADC in longer intervals were reported to produce better outcomes, overcoming BSB, compared to lower doses administered more frequently [120].

4. Applications of ADCs for Oncological and Non-Oncological Conditions

As previously discussed in detail, ADCs have been incorporated in oncological therapeutic algorithms and have shifted treatment landscape from conventional chemotherapy to the era of molecularly targeted medicine. In just the last two years, six novel ADC agents have been approved by FDA for anticancer indications. Table 1 summarizes all these approved oncological ADCs, presenting their main characteristics. The preclinical and clinical data that drove these ADCs to their regulatory approvals have been published recently by our team [1]. In addition to these indications, many other solid malignancies, including prostate cancer [121], gastric cancer [122], pancreatic cancer [123], and hepatocellular carcinoma [124] have been entered into the focus. Table 2 presents some late-stage trials on currently non-approved oncological ADCs.
Being on preliminary steps, ADC technology is gradually tested into a broader spectrum of diseases beyond the sphere of oncology. For such non-oncological implications, the types of payloads vary from glucocorticoid receptor modulators and kinase inhibitors to antibiotics and siRNA (Table 3). Notable innovative approaches have produced anti-inflammatory ADCs binding dexamethasone or other immunomodulatory drugs to mAbs. In 2012, Graversen et al. reported the development of a biodegradable anti-CD163 dexamethasone conjugate that selectively delivers the glucocorticoid to macrophages [125]. Measuring in vitro the suppression of TNF-a secretion by rat macrophages, the ADC demonstrated approximately 50-fold higher anti-inflammatory activity compared to the unconjugated dexamethasone. The in vitro efficacy was replicated by in vivo findings and the anti-inflammatory activity of this ADC was consistently confirmed by several other preclinical studies [126,127].
ABBV-3373 is a novel antibody-glucocorticoid conjugate that is currently being tested for the treatment of moderate/severe rheumatoid arthritis. ABBV-3373 was constructed by conjugating a glucocorticoid receptor mediator to an anti-TNFa mAb and qualified to clinical evaluation after displaying significant properties in mouse models of arthritis [5]. In a phase II trial (NCT03823391) completed in August 2020, 48 patients were randomly allocated in a 2:1 ratio to receive ABBV-3373 or adalimumab [128]. The primary endpoint, change in Disease Activity Score 28 C-Reactive Protein, was significantly greater in the experimental arm compared to the control arm (−2.65 versus −2.13, respectively, p = 0.022). In terms of safety, the adverse events (AEs) rate was lower in the ADC group (35% vs 71%, respectively).
In the context of infectious diseases, DSTA4637S is the first attempt to develop an antibody-antibiotic conjugate (AAC). DSTA4637S is designed to target and eliminate intracellular reservoirs of Staphylococcus aureus. DSTA4637S comprises an IgG1 mAb (MSTA3852A) against a S. aureus antigen, β-N-acetylglucosamine cell-wall teichoic acid (β-GlcNAc-WTA), conjugated to a rifamycin derivative, 4-dimethylamino piperidino-hydroxybenzoxazino rifamycin (dmDNA31), via a val-cit linker [129]. Its action is based on the opsonization of bacteria and the subsequent endocytosis by the host macrophages. Upon internalization, phagolysosomal process results in linker cleavage and antibiotic release. Once dmDNA31 is liberated, it kills AAC-opsonized and pre-existing bacteria within the phagocytes. DSTA4637S remains stable in circulation after intravenous administration, with minimal antibiotic deconjugation reported [130]. In mouse models, the pharmacokinetic profile was similar between S. aureus infected and non-infected subjects [131]. Notably, a single dose of the AAC resulted in a significant reduction of bacterial load in infected mice for 14 days following the administration. This prolonged bactericidal efficiency is attributed to the extended half-life of the antibiotic once conjugated (unconjugated vs. conjugated half-life: 3–4 h versus 4 days) [131]. In a subsequent study, whole-body bioluminescence imaging was used to examine the antibacterial activity of the conjugate as a monotherapy and in combination with vancomycin in mice injected with luminescent S. aureus. AAC administration yielded persistent bioluminescent intensity reduction, while it also achieved augmented potency in the vancomycin combinatorial regimen [132]. Based on these data, the pharmacokinetic and safety profiles of DSTA4637S were investigated in healthy volunteers in a phase I trial (NCT02596399). No subject withdrawals and no serious AEs were reported upon study completion [6]. Another phase I study (NCT03162250) examining safety and tolerability of DSTA4637S in patients with MRSA (Methicillin-resistant Staphylococcus aureus) and MSSA (Methicillin-susceptible Staphylococcus aureus) bacteremia was recently completed and its results are expected.
Among FDA-approved ADCs, BV is the only agent currently evaluated for a non-oncological condition. BV involves a chimeric anti-CD30 IgG1 tethered via a protease-cleavable linker to the MMAE, bearing DAR = 4 [133]. The efficacy of BV in the treatment of diffuse cutaneous systemic sclerosis is currently tested in two phase II trials (NCT03222492, NCT03198689). According to an interim report of second study, BV has already met the primary endpoint at 24 weeks after treatment initiation (decrease in modified Rodnan skin score of ≥8) [4]. BV has been previously investigated for steroid refractory acute graft versus host disease (NCT01616680) and systemic lupus erythematosus (NCT02533570); however, both trials were discontinued.
Table 3. Summary of ADCs tested for non-oncological indications.
Table 3. Summary of ADCs tested for non-oncological indications.
ADC IndicationAntibodyLinkerDARTesting StatusInitial Publication, YearReference
Anti-E Selectin
dexamethasone
(Dexa–AbhEsel)
Chronic models of inflammationMurine anti-E-selectin mAb (H18/7)Succinate linker2.3In vitro preclinicalJ Immunol, 2002[134]
Anti-CD163 dexamethasone
(Cymac-001)
Chronic models of inflammationMurine anti-CD163 mAb (Ed-2)Hemisuccinate linker~4In vivo preclinicalMol Ther, 2012[125]
Anti-CD74 fluticasone propionate
(Anti-CD74-flu449)
Autoimmune modelsHuman anti-CD74 mAb Pyrophosphate acetal linker≥1.7In vivo preclinicalBioconjug Chem, 2018[135]
Anti-CD70 budesonideChronic models of inflammationMurine anti- CD70 mAb (Bu69)CatPhos linker1.9In vitro preclinicalBioconjug Chem, 2016[136]
Anti-CXCR4 dasatinibAutoimmune and inflammatory modelsHumanized anti-CXCR4 mAb (HLCX)Tetra-poly-ethylene glycol linker~3In vitro preclinicalJ Am Chem Soc, 2015 [137]
Anti-CD11a PDE4 inhibitor Chronic models of inflammationHumanized anti-CD11 mAb PEG4-Phe-Lys~2In vivo preclinicalMol Ther, 2016[138]
Anti-CD11a LXR agonist
AtherosclerosisHumanized anti-CD11 mAb PEG4-Phe-Lys2In vitro preclinicalBioconjug Chem, 2015[139]
Anti-CD71 siRNAMuscular diseasesMurine ant-CD71 mAbMaleimide linkerN/AIn vivo preclinicalJ Control Release, 2016[140]
Anti-TNFRSF13c siRNA Myasthenia gravisMurine anti-TNFRSF13c mAbProtamine linkerN/AIn vivo preclinicalClin Immunol, 2017[141]
Anti-IL-7R MMAE
(A7R-ADC-MMAE)
Steroid-resistant arthritisMurine anti-IL-7R mAbVal-Cit linkerN/AIn vivo preclinicalSci Rep, 2017[142]
Anti-CD30 vedotin
(ADCETRIS)
Systemic sclerosisChimeric anti-CD30 mAb (cAC10, SGN-30)Val-Cit linker~4Phase II clinical trial
(NCT03198689),
(NCT03222492)
Ann Rheum Dis, 2021[4]
Anti-CD117 saporin
(C117-ADC)
Conditioning for HSCTMurine anti-CD117 mAbN/AN/AIn vivo preclinicalNat Commun, 2019[143]
Anti-CD45 saporin
(CD45-SAP)
Conditioning for HSCTMurine anti-CD45 mAbN/AN/AIn vivo preclinicalNat Biotechnol,
2016
[144]
Anti-IL-6 alendronateRheumatoid arthritisHumanized anti-IL-6 mAb (tocilizumab)PDPH-PEG-NHS N/AIn vivo preclinicalBioconjug Chem, 2017[145]
Anti–C5aR1 C5 siRNARheumatoid arthritisMurine anti-C5aR1 mAbProtamine linkerN/AIn vivo preclinicalJ Immunol, 2015[146]
Anti-FRβ Pseudomonas exotoxin A (PE38)Rheumatoid arthritisMurine anti-FRβ mAbN/AN/AIn vivo preclinicalArthritis Rheumatol, 2006[147]
Anti-TNFα glucocorticoid
(ABBV-3373)
Rheumatoid arthritisN/AN/AN/APhase II clinical trial
(NCT03823391)
Ann Rheum Dis, 2021[5]
Anti-S. aureus antibiotic
(DSTA4637S)
S. aureus bacteremiaHuman anti-β-N- acetylglucosamine cell-wall teichoic acid (β-GlcNAc- WTA) mAbMC-Val-Cit-PAB-OH2Phase I clinical trial
(NCT03162250)
Nature, 2015[129]
Abbreviations: DAR: Drug-to-Antibody Ratio; mAb: monoclonal antibody N/A: not available.

5. Conclusions

Capitalizing on the extensive research of last decade, ADCs are entering into a phase of exponential growth. Accumulating clinical and preclinical experience will guide the production of agents with greater potency and better therapeutic window than parental compounds. In the oncological setting, promising strategies such as bispecific antibodies and dual-drug ADCs are expected to overcome limitations of first-generation ADCs. At the same time, the preliminary implications of ADC pioneering technology outside of the oncological sphere are expected to extend this tempered optimism in a variety of other non-oncological diseases.

Author Contributions

C.T. and P.-P.L. reviewed the literature and C.T., P.-P.L., M.S., H.G. and D.C.Z. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Published data supporting this article are included within the reference list. Please contact corresponding author for any further requests or supplementary information.

Conflicts of Interest

HG has received grants and personal fees by Roche, BMS, MSD, Novartis and personal fees by Amgen and Pierre Fabre, outside the submitted work. All other authors declare no conflict of interest.

Abbreviations

ADCsAntibody-drug conjugates
mAbsMonoclonal antibodies
bsADCsBispecific ADCs
BSBBinding-Site Barrier
PRLRProlactin receptor
GOGemtuzumab ozogamicin
BVBrentuximab Vedotin
DARDrug-to-antibody ratio
T-DM1Τrastuzumab emtasine
AMLAcute myeloid leukemia
Val-CitValine-citrulline
PABCPara-amino-benzyloxycarbonyl
PBDsPyrrolobenzodiazepines
TRX1,2,4-trioxolane
uAAUnnatural amino-acids
mTGMicrobial transglutaminase
SortASortase A
FGEFormylglycine-generating enzyme
V-ATPaseVacuolar H+-ATPase
CPTCamptothecin
AF-HPAAuristatin F-hydroxypropylamide
MMAEMonomethyl auristatin E
MMAFMonomethyl auristatin F
TM-ADCTrastuzumab–maytansinoid ADC
AACAntibody-antibiotic conjugate
MRSAMethicillin-resistant Staphylococcus aureus
MSSAMethicillin-susceptible Staphylococcus aureus

References

  1. Theocharopoulos, C.; Lialios, P.-P.; Gogas, H.; Ziogas, D.C. An overview of Antibody-Drug conjugates in oncological practice. Ther. Adv. Med. Oncol. 2020, 12, 1758835920962997. [Google Scholar] [CrossRef] [PubMed]
  2. Peters, C.; Brown, S. Antibody-Drug conjugates as novel anti-cancer chemotherapeutics. Biosci. Rep. 2015, 35, e00225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. 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]
  4. Fernandez-Codina, A.; Nevskaya, T.; Pope, J. OP0172 brentuximab vedontin for skin involvement in refractory diffuse cutaneous systemic sclerosis, interim results of a phase IIb open-label trial. Ann. Rheum. Dis. 2021, 80, 103–104. [Google Scholar] [CrossRef]
  5. Stoffel, B.; McPherson, M.; Hernandez, A.; Goess, C.; Mathieu, S.; Waegell, W.; Bryant, S.; Hobson, A.; Ruzek, M.; Pang, Y.; et al. POS0365 anti-TNF glucocorticoid receptor modulator antibody drug conjugate for the treatment of autoimmune diseases. Ann. Rheum. Dis. 2021, 80, 412–413. [Google Scholar] [CrossRef]
  6. Peck, M.; Rothenberg, M.E.; Deng, R.; Lewin-Koh, N.; She, G.; Kamath, A.V.; Carrasco-Triguero, M.; Saad, O.; Castro, A.; Teufel, L.; et al. A Phase 1, Randomized, Single-Ascending-Dose Study to Investigate the Safety, Tolerability, and Pharmacokinetics of DSTA4637S, an Anti-Staphylococcus aureus Thiomab Antibody-Antibiotic Conjugate, in Healthy Volunteers. Antimicrob. Agents Chemother. 2019, 63, e02588-18. [Google Scholar] [CrossRef] [Green Version]
  7. Harper, J.; Hollingsworth, R. Selecting Optimal Antibody-Drug Conjugate Targets Using Indication-Dependent or Indication-Independent Approaches: Fundamentals, Drug Development, and Clinical Outcomes to Target Cancer; John Wiley & Sons, Inc.: New York, NY, USA, 2016; pp. 33–58. [Google Scholar]
  8. Ritchie, M.; Bloom, L.; Carven, G.; Sapra, P. Selecting an Optimal Antibody for Antibody- Drug Conjugate Therapy. In Antibody-Drug Conjugates; AAPS Advances in the Pharmaceutical Sciences Series; Springer: Cham, Switzerland, 2015; Volume 17, pp. 23–48. [Google Scholar] [CrossRef]
  9. Goldmacher, V.S.; Kovtun, Y.V. Antibody-Drug conjugates: Using monoclonal antibodies for delivery of cytotoxic payloads to cancer cells. Ther. Deliv. 2011, 2, 397–416. [Google Scholar] [CrossRef] [Green Version]
  10. Bander, N.H. Antibody-Drug Conjugate Target Selection: Critical Factors. In Antibody-Drug Conjugates; Methods in Molecular Biology; Humana Press: Totowa, NJ, USA, 2013; Volume 1045, pp. 29–40. [Google Scholar] [CrossRef]
  11. Sharma, S.; Li, Z.; Bussing, D.; Shah, D.K. Evaluation of Quantitative Relationship Between Target Expression and Antibody-Drug Conjugate Exposure Inside Cancer Cells. Drug Metab. Dispos. 2020, 48, 368–377. [Google Scholar] [CrossRef]
  12. Bussing, D.; Sharma, S.; Li, Z.; Meyer, L.F.; Shah, D.K. Quantitative Evaluation of the Effect of Antigen Expression Level on Antibody-Drug Conjugate Exposure in Solid Tumor. AAPS J. 2021, 23, 1–11. [Google Scholar] [CrossRef]
  13. Polson, A.G.; Calemine-Fenaux, J.; Chan, P.; Chang, W.; Christensen, E.; Clark, S.; De Sauvage, F.J.; Eaton, D.; Elkins, K.; Elliott, J.M.; et al. Antibody-Drug Conjugates for the Treatment of Non–Hodgkin’s Lymphoma: Target and Linker-Drug Selection. Cancer Res. 2009, 69, 2358–2364. [Google Scholar] [CrossRef] [Green Version]
  14. Staudacher, A.H.; Brown, M.P. Antibody drug conjugates and bystander killing: Is antigen-dependent internalisation required? Br. J. Cancer 2017, 117, 1736–1742. [Google Scholar] [CrossRef] [PubMed]
  15. Perrino, E.; Steiner, M.; Krall, N.; Bernardes, G.; Pretto, F.; Casi, G.; Neri, D. Curative Properties of Noninternalizing Antibody-Drug Conjugates Based on Maytansinoids. Cancer Res. 2014, 74, 2569–2578. [Google Scholar] [CrossRef] [Green Version]
  16. Gébleux, R.; Stringhini, M.; Casanova, R.; Soltermann, A.; Neri, D. Non-internalizing antibody-drug conjugates display potent anti-cancer activity upon proteolytic release of monomethyl auristatin E in the subendothelial extracellular matrix. Int. J. Cancer 2016, 140, 1670–1679. [Google Scholar] [CrossRef] [Green Version]
  17. Rossin, R.; Versteegen, R.M.; Wu, J.; Khasanov, A.; Wessels, H.J.; Steenbergen, E.J.; Hoeve, W.T.; Janssen, H.M.; Van Onzen, A.H.A.M.; Hudson, P.J.; et al. Chemically triggered drug release from an antibody-drug conjugate leads to potent antitumour activity in mice. Nat. Commun. 2018, 9, 1484. [Google Scholar] [CrossRef] [Green Version]
  18. Zhang, Y.; Pastan, I. High Shed Antigen Levels within Tumors: An Additional Barrier to Immunoconjugate Therapy. Clin. Cancer Res. 2008, 14, 7981–7986. [Google Scholar] [CrossRef] [Green Version]
  19. Bogen, J.P.; Hinz, S.C.; Grzeschik, J.; Ebenig, A.; Krah, S.; Zielonka, S.; Kolmar, H. Dual Function pH Responsive Bispecific Antibodies for Tumor Targeting and Antigen Depletion in Plasma. Front. Immunol. 2019, 10, 1892. [Google Scholar] [CrossRef] [Green Version]
  20. Pak, Y.; Zhang, Y.; Pastan, I.; Lee, B. Antigen Shedding May Improve Efficiencies for Delivery of Antibody-Based Anticancer Agents in Solid Tumors. Cancer Res. 2012, 72, 3143–3152. [Google Scholar] [CrossRef] [Green Version]
  21. Hoffmann, R.M.; Coumbe, B.G.T.; Josephs, D.H.; Mele, S.; Ilieva, K.M.; Cheung, A.; Tutt, A.N.; Spicer, J.; Thurston, D.E.; Crescioli, S.; et al. Antibody structure and engineering considerations for the design and function of Antibody Drug Conjugates (ADCs). OncoImmunology 2017, 7, e1395127. [Google Scholar] [CrossRef]
  22. Harding, F.A.; Stickler, M.M.; Razo, J.; DuBridge, R.B. The immunogenicity of humanized and fully human antibodies: Residual immunogenicity resides in the CDR regions. mAbs 2010, 2, 256–265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Waldmann, H. Human Monoclonal Antibodies: The Benefits of Humanization. In Human Monoclonal Antibodies; Methods in Molecular Biology; Humana Press: New York, NY, USA, 2019; Volume 1904, pp. 1–10. [Google Scholar] [CrossRef]
  24. De Taeye, S.W.; Rispens, T.; Vidarsson, G. The Ligands for Human IgG and Their Effector Functions. Antibodies 2019, 8, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Stewart, R.; Hammond, S.A.; Oberst, M.; Wilkinson, R.W. The role of Fc gamma receptors in the activity of immunomodulatory antibodies for cancer. J. Immunother. Cancer 2014, 2, 29. [Google Scholar] [CrossRef] [Green Version]
  26. Vidarsson, G.; Dekkers, G.; Rispens, T. IgG Subclasses and Allotypes: From Structure to Effector Functions. Front. Immunol. 2014, 5, 520. [Google Scholar] [CrossRef] [Green Version]
  27. Yu, J.; Song, Y.; Tian, W. How to select IgG subclasses in developing anti-tumor therapeutic antibodies. J. Hematol. Oncol. 2020, 13, 45. [Google Scholar] [CrossRef]
  28. Bruhns, P.; Iannascoli, B.; England, P.; Mancardi, D.A.; Fernandez, N.; Jorieux, S.; Daëron, M. Specificity and affinity of human Fcγ receptors and their polymorphic variants for human IgG subclasses. Blood 2009, 113, 3716–3725. [Google Scholar] [CrossRef]
  29. 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]
  30. Ackerman, M.E.; Pawlowski, D.; Wittrup, K.D. Effect of antigen turnover rate and expression level on antibody penetration into tumor spheroids. Mol. Cancer Ther. 2008, 7, 2233–2240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Brinkmann, U.; Kontermann, R.E. The making of bispecific antibodies. mAbs 2016, 9, 182–212. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Q.; Chen, Y.; Park, J.; Liu, X.; Hu, Y.; Wang, T.; McFarland, K.; Betenbaugh, M.J. Design and Production of Bispecific Antibodies. Antibodies 2019, 8, 43. [Google Scholar] [CrossRef] [Green Version]
  33. Comer, F.; Gao, C.; Coats, S. Bispecific and Biparatopic Antibody Drug Conjugates. In Innovations for Next-Generation Antibody-Drug Conjugates; Cancer Drug Discovery and Development; Humana Press: Cham, Switzerland, 2018; pp. 267–280. [Google Scholar]
  34. Maruani, A. Bispecifics and Antibody-Drug conjugates: A positive synergy. Drug Discov. Today Technol. 2018, 30, 55–61. [Google Scholar] [CrossRef] [Green Version]
  35. De Goeij, B.E.; Vink, T.; Napel, H.T.; Breij, E.C.W.; Satijn, D.; Wubbolts, R.; Miao, D.; Parren, P. Efficient Payload Delivery by a Bispecific Antibody-Drug Conjugate Targeting HER2 and CD63. Mol. Cancer Ther. 2016, 15, 2688–2697. [Google Scholar] [CrossRef] [Green Version]
  36. Andreev, J.; Thambi, N.; Bay, A.E.P.; Delfino, F.; Martin, J.; Kelly, M.P.; Kirshner, J.R.; Rafique, A.; Kunz, A.; Nittoli, T.; et al. Bispecific Antibodies and Antibody-Drug Conjugates (ADCs) Bridging HER2 and Prolactin Receptor Improve Efficacy of HER2 ADCs. Mol. Cancer Ther. 2017, 16, 681–693. [Google Scholar] [CrossRef] [Green Version]
  37. Bay, A.P.; Kalsy, A.; Tiwari, S.; Luan, B.; Kunz, A.; Chen, Z.; Zhang, L.; Potocky, T.; Nittoli, T.; Thurston, G.; et al. Abstract 233: Bispecific HER2 ADC: Making more potent HER2 ADC by improving target internalization. In Proceedings of the AACR Annual Meeting 2019, Atlanta, GA, USA, 29 March–3 April 2019; Volume 79. [Google Scholar]
  38. Bethell, R.; Eberlein, C.; Pollard, V.; Georgiou, T.; Orphanides, G.; Mooney, L.; Kendrew, J.; Daniels, T. Abstract 2887: Bispecific antibody drug conjugates targeting CD7 and CD33 for the treatment of acute myeloid leukemia. In Proceedings of the AACR Annual Meeting 2020, Philadelphia, PA, USA, 27–28 April and 22–24 June 2020; Volume 80. [Google Scholar] [CrossRef]
  39. Li, J.Y.; Perry, S.R.; Muniz-Medina, V.; Wang, X.; Wetzel, L.K.; Rebelatto, M.C.; Hinrichs, M.J.M.; Bezabeh, B.Z.; Fleming, R.L.; DiMasi, N.; et al. A Biparatopic HER2-Targeting Antibody-Drug Conjugate Induces Tumor Regression in Primary Models Refractory to or Ineligible for HER2-Targeted Therapy. Cancer Cell 2019, 35, 948–949. [Google Scholar] [CrossRef] [Green Version]
  40. Pegram, M.D.; Hamilton, E.P.; Tan, A.R.; Storniolo, A.M.; Balic, K.; Rosenbaum, A.I.; Liang, M.; He, P.; Marshall, S.; Scheuber, A.; et al. First-in-Human, Phase 1 Dose-Escalation Study of Biparatopic Anti-HER2 Antibody-Drug Conjugate MEDI4276 in Patients with HER2+ Advanced Breast or Gastric Cancer. Mol. Cancer Ther. 2021. [Google Scholar] [CrossRef] [PubMed]
  41. 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]
  42. Ducry, L.; Stump, B. Antibody−Drug Conjugates: Linking Cytotoxic Payloads to Monoclonal Antibodies. Bioconjug. Chem. 2009, 21, 5–13. [Google Scholar] [CrossRef] [PubMed]
  43. Balendiran, G.K.; Dabur, R.; Fraser, D. The role of glutathione in cancer. Cell Biochem. Funct. 2004, 22, 343–352. [Google Scholar] [CrossRef]
  44. Bryden, F.; Martin, C.; Letast, S.; Lles, E.; Viéitez-Villemin, I.; Rousseau, A.; Colas, C.; Brachet-Botineau, M.; Allard-Vannier, E.; Larbouret, C.; et al. Impact of cathepsin B-sensitive triggers and hydrophilic linkers on in vitro efficacy of novel site-specific Antibody-Drug conjugates. Org. Biomol. Chem. 2018, 16, 1882–1889. [Google Scholar] [CrossRef]
  45. Dan, N.; Setua, S.; Kashyap, V.K.; Khan, S.; Jaggi, M.; Yallapu, M.M.; Chauhan, S.C. Antibody-Drug Conjugates for Cancer Therapy: Chemistry to Clinical Implications. Pharmaceuticals 2018, 11, 32. [Google Scholar] [CrossRef] [Green Version]
  46. Tsuchikama, K.; An, Z. Antibody-drug conjugates: Recent advances in conjugation and linker chemistries. Protein Cell 2016, 9, 33–46. [Google Scholar] [CrossRef] [Green Version]
  47. Jeffrey, S.C.; Andreyka, J.B.; Bernhardt, S.X.; Kissler, K.M.; Kline, T.; Lenox, J.S.; Moser, R.F.; Nguyen, M.T.; Okeley, N.M.; Stone, I.J.; et al. Development and Properties of β-Glucuronide Linkers for Monoclonal Antibody−Drug Conjugates. Bioconjug. Chem. 2006, 17, 831–840. [Google Scholar] [CrossRef]
  48. Conilh, L.; Fournet, G.; Fourmaux, E.; Murcia, A.; Matera, E.-L.; Joseph, B.; Dumontet, C.; Viricel, W. Exatecan Antibody Drug Conjugates Based on a Hydrophilic Polysarcosine Drug-Linker Platform. Pharmaceuticals 2021, 14, 247. [Google Scholar] [CrossRef]
  49. Nolting, B. Linker Technologies for Antibody-Drug Conjugates. In Antibody-Drug Conjugates; Methods in Molecular Biology; Humana Press: Totowa, NJ, USA, 2013; Volume 1045, pp. 71–100. [Google Scholar] [CrossRef]
  50. Erickson, H.K.; Widdison, W.C.; Mayo, M.F.; Whiteman, K.; Audette, C.; Wilhelm, S.D.; Singh, R. Tumor Delivery and In Vivo Processing of Disulfide-Linked and Thioether-Linked Antibody−Maytansinoid Conjugates. Bioconjug. Chem. 2009, 21, 84–92. [Google Scholar] [CrossRef]
  51. Kern, J.C.; Cancilla, M.T.; Dooney, D.; Kwasnjuk, K.; Zhang, R.; Beaumont, M.; Figueroa, I.; Hsieh, S.; Liang, L.; Tomazela, D.; et al. Discovery of Pyrophosphate Diesters as Tunable, Soluble, and Bioorthogonal Linkers for Site-Specific Antibody-Drug Conjugates. J. Am. Chem. Soc. 2016, 138, 1430–1445. [Google Scholar] [CrossRef]
  52. Kolodych, S.; Michel, C.; Delacroix, S.; Koniev, O.; Ehkirch, A.; Eberova, J.; Cianférani, S.; Renoux, B.; Krezel, W.; Poinot, P.; et al. Development and evaluation of β-galactosidase-sensitive antibody-drug conjugates. Eur. J. Med. Chem. 2017, 142, 376–382. [Google Scholar] [CrossRef] [PubMed]
  53. Bargh, J.D.; Walsh, S.J.; Ashman, N.; Isidro-Llobet, A.; Carroll, J.S.; Spring, D.R. A dual-enzyme cleavable linker for Antibody-Drug conjugates. Chem. Commun. 2021, 57, 3457–3460. [Google Scholar] [CrossRef]
  54. Kumar, A.; Kinneer, K.; Masterson, L.; Ezeadi, E.; Howard, P.; Wu, H.; Gao, C.; Dimasi, N. Synthesis of a heterotrifunctional linker for the site-specific preparation of antibody-drug conjugates with two distinct warheads. Bioorg. Med. Chem. Lett. 2018, 28, 3617–3621. [Google Scholar] [CrossRef] [PubMed]
  55. Spangler, B.; Kline, T.; Hanson, J.; Li, X.; Zhou, S.; Wells, J.A.; Sato, A.K.; Renslo, A.R. Toward a Ferrous Iron-Cleavable Linker for Antibody-Drug Conjugates. Mol. Pharm. 2018, 15, 2054–2059. [Google Scholar] [CrossRef]
  56. Merkul, E.; Sijbrandi, N.J.; Muns, J.A.; Aydin, I.; Adamzek, K.; Houthoff, H.-J.; Nijmeijer, B.; Van Dongen, G.A. First platinum(II)-based metal-organic linker technology (Lx®) for a plug-and-play development of antibody-drug conjugates (ADCs). Expert Opin. Drug Deliv. 2019, 16, 783–793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Dovgan, I.; Ehkirch, A.; Lehot, V.; Kuhn, I.; Koniev, O.; Kolodych, S.; Hentz, A.; Ripoll, M.; Ursuegui, S.; Nothisen, M.; et al. On the use of DNA as a linker in antibody-drug conjugates: Synthesis, stability and in vitro potency. Sci. Rep. 2020, 10, 7691–7699. [Google Scholar] [CrossRef]
  58. Li, J.; Xiao, D.; Xie, F.; Li, W.; Zhao, L.; Sun, W.; Yang, X.; Zhou, X. Novel antibody-drug conjugate with UV-controlled cleavage mechanism for cytotoxin release. Bioorg. Chem. 2020, 111, 104475. [Google Scholar] [CrossRef]
  59. Bird, M.; Nunes, J.; Frigerio, M. Bridged Cysteine Conjugations. In Antibody-Drug Conjugates; Methods in Molecular Biology; Humana: New York, NY, USA, 2020; Volume 2078, pp. 113–129. [Google Scholar] [CrossRef]
  60. Junutula, J.R.; Raab, H.; Clark, S.; Bhakta, S.; Leipold, D.D.; Weir, S.; Chen, Y.; Simpson, M.; Tsai, S.P.; Dennis, M.S.; et al. Site-specific conjugation of a cytotoxic drug to an antibody improves the therapeutic index. Nat. Biotechnol. 2008, 26, 925–932. [Google Scholar] [CrossRef]
  61. Tian, F.; Lu, Y.; Manibusan, A.; Sellers, A.; Tran, H.; Sun, Y.; Phuong, T.; Barnett, R.; Hehli, B.; Song, F.; et al. A general approach to site-specific antibody drug conjugates. Proc. Natl. Acad. Sci. USA 2014, 111, 1766–1771. [Google Scholar] [CrossRef] [Green Version]
  62. Zimmerman, E.S.; Heibeck, T.H.; Gill, A.; Li, X.; Murray, C.J.; Madlansacay, M.R.; Tran, C.; Uter, N.T.; Yin, G.; Rivers, P.J.; et al. Production of Site-Specific Antibody-Drug Conjugates Using Optimized Non-Natural Amino Acids in a Cell-Free Expression System. Bioconjug. Chem. 2014, 25, 351–361. [Google Scholar] [CrossRef]
  63. Jeger, S.; Zimmermann, K.; Blanc, A.; Grünberg, J.; Honer, M.; Hunziker, P.; Struthers, H.; Schibli, R. Site-Specific and Stoichiometric Modification of Antibodies by Bacterial Transglutaminase. Angew. Chem. Int. Ed. 2010, 49, 9995–9997. [Google Scholar] [CrossRef]
  64. Dennler, P.; Chiotellis, A.; Fischer, E.; Brégeon, D.; Belmant, C.; Gauthier, L.; Lhospice, F.; Romagne, F.; Schibli, R. Transglutaminase-Based Chemo-Enzymatic Conjugation Approach Yields Homogeneous Antibody-Drug Conjugates. Bioconjug. Chem. 2014, 25, 569–578. [Google Scholar] [CrossRef]
  65. Strop, P.; Liu, S.-H.; Dorywalska, M.; Delaria, K.; Dushin, R.G.; Tran, T.-T.; Ho, W.-H.; Farias, S.; Casas, M.G.; Abdiche, Y.; et al. Location Matters: Site of Conjugation Modulates Stability and Pharmacokinetics of Antibody Drug Conjugates. Chem. Biol. 2013, 20, 161–167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Anami, Y.; Xiong, W.; Gui, X.; Deng, M.; Zhang, C.C.; Zhang, N.; An, Z.; Tsuchikama, K. Enzymatic conjugation using branched linkers for constructing homogeneous Antibody-Drug conjugates with high potency. Org. Biomol. Chem. 2017, 15, 5635–5642. [Google Scholar] [CrossRef]
  67. Beerli, R.R.; Hell, T.; Merkel, A.S.; Grawunder, U. Sortase Enzyme-Mediated Generation of Site-Specifically Conjugated Antibody Drug Conjugates with High In Vitro and In Vivo Potency. PLoS ONE 2015, 10, e0131177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Drake, P.M.; Albers, A.E.; Baker, J.; Banas, S.; Barfield, R.M.; Bhat, A.S.; De Hart, G.W.; Garofalo, A.W.; Holder, P.; Jones, L.C.; et al. Aldehyde Tag Coupled with HIPS Chemistry Enables the Production of ADCs Conjugated Site-Specifically to Different Antibody Regions with Distinct in Vivo Efficacy and PK Outcomes. Bioconjug. Chem. 2014, 25, 1331–1341. [Google Scholar] [CrossRef]
  69. Zhou, Q.; Stefano, J.E.; Manning, C.; Kyazike, J.; Chen, B.; Gianolio, D.A.; Park, A.; Busch, M.; Bird, J.; Zheng, X.; et al. Site-Specific Antibody-Drug Conjugation through Glycoengineering. Bioconjug. Chem. 2014, 25, 510–520. [Google Scholar] [CrossRef] [PubMed]
  70. Tang, F.; Shi, W.; Huang, W. Homogeneous Antibody-Drug Conjugates via Glycoengineering. In Bioconjugation; Methods in Molecular Biology; Humana: New York, NY, USA, 2019; Volume 2033, pp. 221–238. [Google Scholar] [CrossRef]
  71. Maiese, W.M.; Lechevalier, M.P.; Lechevalier, H.A.; Korshalla, J.; Kuck, N.; Fantini, A.; Wildey, M.J.; Thomas, J.; Greenstein, M. Calicheamicins, a novel family of antitumor antibiotics. Taxonomy, fermentation and biological properties. J. Antibiot. 1989, 42, 558–563. [Google Scholar] [CrossRef]
  72. Ellestad, G.A. Structural and conformational features relevant to the anti-tumor activity of calicheamicin γ1I? 1I. Chirality 2011, 23, 660–671. [Google Scholar] [CrossRef]
  73. Vollmar, B.S.; Frantz, C.; Schutten, M.M.; Zhong, F.; del Rosario, G.; Go, M.A.T.; Yu, S.-F.; Leipold, D.D.; Kamath, A.V.; Ng, C.; et al. Calicheamicin Antibody-Drug Conjugates with Improved Properties. Mol. Cancer Ther. 2021, 20, 1112–1120. [Google Scholar] [CrossRef]
  74. Hartley, J.A.; Spanswick, V.J.; Brooks, N.; Clingen, P.H.; McHugh, P.J.; Hochhauser, D.; Pedley, R.B.; Kelland, L.R.; Alley, M.C.; Schultz, R.; et al. SJG-136 (NSC 694501), a Novel Rationally Designed DNA Minor Groove Interstrand Cross-Linking Agent with Potent and Broad Spectrum Antitumor Activity: Part 1: Cellular Pharmacology, In vitro and Initial In vivo Antitumor Activity. Cancer Res. 2004, 64, 6693–6699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Hartley, J.A. The development of pyrrolobenzodiazepines as antitumour agents. Expert Opin. Investig. Drugs 2011, 20, 733–744. [Google Scholar] [CrossRef] [PubMed]
  76. Ma, Y.; Khojasteh, S.C.; Hop, C.E.; Erickson, H.K.; Polson, A.; Pillow, T.H.; Yu, S.-F.; Wang, H.; Dragovich, P.S.; Zhang, D. Antibody Drug Conjugates Differentiate Uptake and DNA Alkylation of Pyrrolobenzodiazepines in Tumors from Organs of Xenograft Mice. Drug Metab. Dispos. 2016, 44, 1958–1962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Caimi, P.F.; Ai, W.; Alderuccio, J.P.; Ardeshna, K.M.; Hamadani, M.; Hess, B.; Kahl, B.S.; Radford, J.; Solh, M.; Stathis, A.; et al. Loncastuximab tesirine in relapsed or refractory diffuse large B-cell lymphoma (LOTIS-2): A multicentre, open-label, single-arm, phase 2 trial. Lancet Oncol. 2021, 22, 790–800. [Google Scholar] [CrossRef]
  78. Waight, A.B.; Bargsten, K.; Doronina, S.; Steinmetz, M.; Sussman, D.; Prota, A.E. Structural Basis of Microtubule Destabilization by Potent Auristatin Anti-Mitotics. PLoS ONE 2016, 11, e0160890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Doronina, S.O.; Mendelsohn, B.; Bovee, T.D.; Cerveny, C.G.; Alley, S.C.; Meyer, D.L.; Oflazoglu, E.; Toki, B.E.; Sanderson, R.J.; Zabinski, R.F.; et al. Enhanced Activity of Monomethylauristatin F through Monoclonal Antibody Delivery: Effects of Linker Technology on Efficacy and Toxicity. Bioconjug. Chem. 2005, 17, 114–124. [Google Scholar] [CrossRef] [PubMed]
  80. Yurkovetskiy, A.V.; Bodyak, N.D.; Yin, M.; Thomas, J.D.; Clardy, S.M.; Conlon, P.R.; Stevenson, C.A.; Uttard, A.; Qin, L.; Gumerov, D.R.; et al. Dolaflexin: A Novel Antibody-Drug Conjugate Platform Featuring High Drug Loading and a Controlled Bystander Effect. Mol. Cancer Ther. 2021, 20, 885–895. [Google Scholar] [CrossRef] [PubMed]
  81. Bodyak, N.D.; Mosher, R.; Yurkovetskiy, A.V.; Yin, M.; Bu, C.; Conlon, P.R.; Demady, D.R.; DeVit, M.J.; Gumerov, D.R.; Gurijala, V.R.; et al. The Dolaflexin-based Antibody-Drug Conjugate XMT-1536 Targets the Solid Tumor Lineage Antigen SLC34A2/NaPi2b. Mol. Cancer Ther. 2021, 20, 896–905. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, L.F.; Desai, S.D.; Li, T.-K.; Mao, Y.; Sun, M.; Sim, S.-P. Mechanism of Action of Camptothecin. Ann. N. Y. Acad. Sci. 2000, 922, 1–10. [Google Scholar] [CrossRef]
  83. Levengood, M.R.; Zhang, X.; Hunter, J.H.; Emmerton, K.K.; Miyamoto, J.B.; Lewis, T.S.; Senter, P.D. Orthogonal Cysteine Protection Enables Homogeneous Multi-Drug Antibody-Drug Conjugates. Angew. Chem. Int. Ed. 2016, 56, 733–737. [Google Scholar] [CrossRef] [PubMed]
  84. Loganzo, F.; Tan, X.; Sung, M.; Jin, G.; Myers, J.S.; Melamud, E.; Wang, F.; Diesl, V.; Follettie, M.T.; Musto, S.; et al. Tumor Cells Chronically Treated with a Trastuzumab–Maytansinoid Antibody-Drug Conjugate Develop Varied Resistance Mechanisms but Respond to Alternate Treatments. Mol. Cancer Ther. 2015, 14, 952–963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Levengood, M.R.; Zhang, X.; Emmerton, K.K.; Hunter, J.H.; Senter, P.D. Abstract 982: Development of homogeneous dual-drug ADCs: Application to the co-delivery of auristatin payloads with complementary antitumor activities. In Proceedings of the AACR Annual Meeting 2017, Washington, DC, USA, 1–5 April 2017; Volume 77. [Google Scholar] [CrossRef]
  86. Duvall, J.R.; Damelin, M.; Kozytska, M.V.; Nehilla, B.J.; Protopopova, M.; Conlon, P.R.; Qin, L.; Nazzaro, M.; Thomas, J.D.; Zhang, Q.; et al. Abstract 232: An antibody-drug conjugate carrying a microtubule inhibitor and a DNA alkylator exerts both mechanisms of action on tumor cells. In Proceedings of the AACR Annual Meeting 2019, Atlanta, GA, USA, 29 March–3 April 2019; Volume 79. [Google Scholar] [CrossRef]
  87. Nilchan, N.; Li, X.; Pedzisa, L.; Nanna, A.R.; Roush, W.R.; Rader, C. Dual-mechanistic antibody-drug conjugate via site-specific selenocysteine/cysteine conjugation. Antib. Ther. 2019, 2, 71–78. [Google Scholar] [CrossRef] [Green Version]
  88. Yamazaki, C.M.; Yamaguchi, A.; Anami, Y.; Xiong, W.; Otani, Y.; Lee, J.; Ueno, N.T.; Zhang, N.; An, Z.; Tsuchikama, K. Antibody-drug conjugates with dual payloads for combating breast tumor heterogeneity and drug resistance. Nat. Commun. 2021, 12, 3528. [Google Scholar] [CrossRef] [PubMed]
  89. Li, F.; Emmerton, K.K.; Jonas, M.; Zhang, X.; Miyamoto, J.B.; Setter, J.R.; Nicholas, N.D.; Okeley, N.M.; Lyon, R.P.; Benjamin, D.R.; et al. Intracellular Released Payload Influences Potency and Bystander-Killing Effects of Antibody-Drug Conjugates in Preclinical Models. Cancer Res. 2016, 76, 2710–2719. [Google Scholar] [CrossRef] [Green Version]
  90. McCombs, J.R.; Owen, S.C. Antibody Drug Conjugates: Design and Selection of Linker, Payload and Conjugation Chemistry. AAPS J. 2015, 17, 339–351. [Google Scholar] [CrossRef] [Green Version]
  91. Singh, A.P.; Sharma, S.; Shah, D.K. Quantitative characterization of in vitro bystander effect of antibody-drug conjugates. J. Pharmacokinet. Pharmacodyn. 2016, 43, 567–582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Kovtun, Y.V.; Goldmacher, V.S. Cell killing by Antibody-Drug conjugates. Cancer Lett. 2007, 255, 232–240. [Google Scholar] [CrossRef]
  93. Burton, J.K.; Bottino, D.; Secomb, T.W. A Systems Pharmacology Model for Drug Delivery to Solid Tumors by Antibody-Drug Conjugates: Implications for Bystander Effects. AAPS J. 2019, 22, 12. [Google Scholar] [CrossRef] [PubMed]
  94. Singh, A.P.; Seigel, G.M.; Guo, L.; Verma, A.; Wong, G.G.-L.; Chang, H.-P.; Shah, D.K. Evolution of the Systems Pharmacokinetics-Pharmacodynamics Model for Antibody-Drug Conjugates to Characterize Tumor Heterogeneity and In Vivo Bystander Effect. J. Pharmacol. Exp. Ther. 2020, 374, 184–199. [Google Scholar] [CrossRef] [Green Version]
  95. Khera, E.; Cilliers, C.; Smith, M.D.; Ganno, M.L.; Lai, K.C.; Keating, T.A.; Kopp, A.; Nessler, I.; Abu-Yousif, A.O.; Thurber, G.M. Quantifying ADC bystander payload penetration with cellular resolution using pharmacodynamic mapping. Neoplasia 2020, 23, 210–221. [Google Scholar] [CrossRef] [PubMed]
  96. Chung, S.W.; Choi, J.U.; Cho, Y.S.; Kim, H.R.; Won, T.H.; Dimitrion, P.; Jeon, O.-C.; Kim, S.W.; Kim, I.-S.; Kim, S.Y.; et al. Self-Triggered Apoptosis Enzyme Prodrug Therapy (STAEPT): Enhancing Targeted Therapies via Recurrent Bystander Killing Effect by Exploiting Caspase-Cleavable Linker. Adv. Sci. 2018, 5, 1800368. [Google Scholar] [CrossRef] [PubMed]
  97. Andrikopoulou, A.; Zografos, E.; Liontos, M.; Koutsoukos, K.; Dimopoulos, M.-A.; Zagouri, F. Trastuzumab Deruxtecan (DS-8201a): The Latest Research and Advances in Breast Cancer. Clin. Breast Cancer 2020, 21, e212–e219. [Google Scholar] [CrossRef] [PubMed]
  98. Ogitani, Y.; Hagihara, K.; Oitate, M.; Naito, H.; Agatsuma, T. Bystander killing effect of DS-8201a, a novel anti-human epidermal growth factor receptor 2 Antibody-Drug conjugate, in tumors with human epidermal growth factor receptor 2 heterogeneity. Cancer Sci. 2016, 107, 1039–1046. [Google Scholar] [CrossRef]
  99. Collins, D.; Bossenmaier, B.; Kollmorgen, G.; Niederfellner, G. Acquired Resistance to Antibody-Drug Conjugates. Cancers 2019, 11, 394. [Google Scholar] [CrossRef] [Green Version]
  100. Hochberg, J.; Alexander, S. Resistance to Antibody-Drug Conjugate. In Resistance to Targeted Therapies in Lymphomas; Resistance to Targeted Anti-Cancer Therapeutics; Springer: Cham, Switzerland, 2019; pp. 57–69. [Google Scholar] [CrossRef]
  101. Luci, C.R.; García-Alonso, S.; Díaz-Rodriguez, E.; Nadal-Serrano, M.; Arribas, J.; Ocana, A.; Pandiella, A. Resistance to the Antibody-Drug Conjugate T-DM1 Is Based in a Reduction in Lysosomal Proteolytic Activity. Cancer Res. 2017, 77, 4639–4651. [Google Scholar] [CrossRef] [Green Version]
  102. Wang, H.; Wang, W.; Xu, Y.; Yang, Y.; Chen, X.; Quan, H.; Lou, L. Aberrant intracellular metabolism of T-DM1 confers T-DM1 resistance in human epidermal growth factor receptor 2-positive gastric cancer cells. Cancer Sci. 2017, 108, 1458–1468. [Google Scholar] [CrossRef]
  103. Trudeau, K.M.; Colby, A.H.; Zeng, J.; Las, G.; Feng, J.H.; Grinstaff, M.W.; Shirihai, O.S. Lysosome acidification by photoactivated nanoparticles restores autophagy under lipotoxicity. J. Cell Biol. 2016, 214, 25–34. [Google Scholar] [CrossRef]
  104. García-Alonso, S.; Ocana, A.; Pandiella, A. Resistance to Antibody-Drug Conjugates. Cancer Res. 2018, 78, 2159–2165. [Google Scholar] [CrossRef] [Green Version]
  105. Hamblett, K.J.; Jacob, A.P.; Gurgel, J.L.; Tometsko, M.E.; Rock, B.M.; Patel, S.K.; Milburn, R.R.; Siu, S.; Ragan, S.P.; Rock, D.A.; et al. SLC46A3 Is Required to Transport Catabolites of Noncleavable Antibody Maytansine Conjugates from the Lysosome to the Cytoplasm. Cancer Res. 2015, 75, 5329–5340. [Google Scholar] [CrossRef] [Green Version]
  106. Chen, R.; Hou, J.; Newman, E.; Kim, Y.; Donohue, C.; Liu, X.; Thomas, S.; Forman, S.J.; Kane, S.E. CD30 Downregulation, MMAE Resistance, and MDR1 Upregulation Are All Associated with Resistance to Brentuximab Vedotin. Mol. Cancer Ther. 2015, 14, 1376–1384. [Google Scholar] [CrossRef] [Green Version]
  107. Chen, R.; Herrera, A.F.; Hou, J.; Chen, L.; Wu, J.; Guo, Y.; Synold, T.W.; Ngo, V.; Puverel, S.; Mei, M.; et al. Inhibition of MDR1 Overcomes Resistance to Brentuximab Vedotin in Hodgkin Lymphoma. Clin. Cancer Res. 2019, 26, 1034–1044. [Google Scholar] [CrossRef]
  108. Rosen, D.B.; Harrington, K.H.; Cordeiro, J.A.; Leung, L.Y.; Putta, S.; Lacayo, N.; Laszlo, G.S.; Gudgeon, C.J.; Hogge, D.; Hawtin, R.E.; et al. AKT Signaling as a Novel Factor Associated with In Vitro Resistance of Human AML to Gemtuzumab Ozogamicin. PLoS ONE 2013, 8, e53518. [Google Scholar] [CrossRef] [PubMed]
  109. Maimaitili, Y.; Inase, A.; Miyata, Y.; Kitao, A.; Mizutani, Y.; Kakiuchi, S.; Shimono, Y.; Saito, Y.; Sonoki, T.; Minami, H.; et al. An mTORC1/2 kinase inhibitor enhances the cytotoxicity of gemtuzumab ozogamicin by activation of lysosomal function. Leuk. Res. 2018, 74, 68–74. [Google Scholar] [CrossRef]
  110. Li, G.; Guo, J.; Shen, B.-Q.; Yadav, D.B.; Sliwkowski, M.X.; Crocker, L.M.; Lacap, J.A.; Phillips, G.D.L. Mechanisms of Acquired Resistance to Trastuzumab Emtansine in Breast Cancer Cells. Mol. Cancer Ther. 2018, 17, 1441–1453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Wang, L.; Wang, Q.; Gao, M.; Fu, L.; Li, Y.; Quan, H.; Lou, L. STAT3 activation confers trastuzumab-emtansine (T-DM1) resistance in HER2-positive breast cancer. Cancer Sci. 2018, 109, 3305–3315. [Google Scholar] [CrossRef] [PubMed]
  112. Walter, R.B.; Raden, B.W.; Cronk, M.R.; Bernstein, I.D.; Appelbaum, F.R.; Banker, D.E.; Bello-Fernandez, C.; Stasakova, J.; Renner, A.; Carballido-Perrig, N.; et al. The peripheral benzodiazepine receptor ligand PK11195 overcomes different resistance mechanisms to sensitize AML cells to gemtuzumab ozogamicin. Blood 2004, 103, 4276–4284. [Google Scholar] [CrossRef]
  113. Saatci, Ö.; Borgoni, S.; Akbulut, Ö.; Durmuş, S.; Raza, U.; Eyüpoğlu, E.; Alkan, C.; Akyol, A.; Kutuk, O.; Wiemann, S.; et al. Targeting PLK1 overcomes T-DM1 resistance via CDK1-dependent phosphorylation and inactivation of Bcl-2/xL in HER2-positive breast cancer. Oncogene 2018, 37, 2251–2269. [Google Scholar] [CrossRef] [Green Version]
  114. Singh, A.P.; Guo, L.; Verma, A.; Wong, G.G.-L.; Thurber, G.M.; Shah, D.K. Antibody Coadministration as a Strategy to Overcome Binding-Site Barrier for ADCs: A Quantitative Investigation. AAPS J. 2020, 22, 28. [Google Scholar] [CrossRef]
  115. Rudnick, S.I.; Adams, G.P. Affinity and Avidity in Antibody-Based Tumor Targeting. Cancer Biother. Radiopharm. 2009, 24, 155–161. [Google Scholar] [CrossRef]
  116. Khera, E.; Cilliers, C.; Bhatnagar, S.; Thurber, G.M. Computational transport analysis of antibody-drug conjugate bystander effects and payload tumoral distribution: Implications for therapy. Mol. Syst. Des. Eng. 2017, 3, 73–88. [Google Scholar] [CrossRef]
  117. Muchekehu, R.; Liu, D.; Horn, M.; Campbell, L.; Del Rosario, J.; Bacica, M.; Moskowitz, H.; Osothprarop, T.; Dirksen, A.; Doppalapudi, V.; et al. The Effect of Molecular Weight, PK, and Valency on Tumor Biodistribution and Efficacy of Antibody-Based Drugs. Transl. Oncol. 2013, 6, 562–572. [Google Scholar] [CrossRef] [Green Version]
  118. Adams, G.P.; Schier, R.; McCall, A.M.; Simmons, H.H.; Horak, E.M.; Alpaugh, R.K.; Marks, J.D.; Weiner, L.M. High affinity restricts the localization and tumor penetration of single-chain fv antibody molecules. Cancer Res. 2001, 61, 4750–4755. [Google Scholar]
  119. Lazzerini, L.; Jöhrens, K.; Sehouli, J.; Cichon, G. Favorable therapeutic response after anti-Mesothelin Antibody-Drug conjugate treatment requires high expression of Mesothelin in tumor cells. Arch. Gynecol. Obstet. 2020, 302, 1255–1262. [Google Scholar] [CrossRef]
  120. Cilliers, C.; Guo, H.; Liao, J.; Christodolu, N.; Thurber, G.M. Multiscale Modeling of Antibody-Drug Conjugates: Connecting Tissue and Cellular Distribution to Whole Animal Pharmacokinetics and Potential Implications for Efficacy. AAPS J. 2016, 18, 1117–1130. [Google Scholar] [CrossRef] [Green Version]
  121. Rosellini, M.; Santoni, M.; Mollica, V.; Rizzo, A.; Cimadamore, A.; Scarpelli, M.; Storti, N.; Battelli, N.; Montironi, R.; Massari, F. Treating Prostate Cancer by Antibody-Drug Conjugates. Int. J. Mol. Sci. 2021, 22, 1551. [Google Scholar] [CrossRef] [PubMed]
  122. Koganemaru, S.; Shitara, K. Antibody-Drug conjugates to treat gastric cancer. Expert Opin. Biol. Ther. 2020, 21, 923–930. [Google Scholar] [CrossRef] [PubMed]
  123. Huang, J.; Agoston, A.T.; Guo, P.; Moses, M.A. A Rationally Designed ICAM1 Antibody Drug Conjugate for Pancreatic Cancer. Adv. Sci. 2020, 7, 2002852. [Google Scholar] [CrossRef] [PubMed]
  124. Dahlgren, D.; Lennernäs, H. Antibody-Drug Conjugates and Targeted Treatment Strategies for Hepatocellular Carcinoma: A Drug-Delivery Perspective. Molecules 2020, 25, 2861. [Google Scholar] [CrossRef]
  125. Graversen, J.H.; Svendsen, P.; Dagnæs-Hansen, F.; Dal, J.; Anton, G.; Etzerodt, A.; Petersen, M.D.; Christensen, P.A.; Møller, H.J.; Moestrup, S.K. Targeting the Hemoglobin Scavenger receptor CD163 in Macrophages Highly Increases the Anti-inflammatory Potency of Dexamethasone. Mol. Ther. 2012, 20, 1550–1558. [Google Scholar] [CrossRef] [Green Version]
  126. Møller, L.N.O.; Knudsen, A.R.; Andersen, K.J.; Nyengaard, J.R.; Hamilton-Dutoit, S.; Møller, E.M.O.; Svendsen, P.; Møller, H.J.; Moestrup, S.K.; Graversen, J.H.; et al. Anti-CD163-dexamethasone protects against apoptosis after ischemia/reperfusion injuries in the rat liver. Ann. Med. Surg. 2015, 4, 331–337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Svendsen, P.; Graversen, J.H.; Etzerodt, A.; Hager, H.; Røge, R.; Grønbæk, H.; Christensen, E.I.; Møller, H.J.; Vilstrup, H.; Moestrup, S.K. Antibody-Directed Glucocorticoid Targeting to CD163 in M2-type Macrophages Attenuates Fructose-Induced Liver Inflammatory Changes. Mol. Ther. Methods Clin. Dev. 2017, 4, 50–61. [Google Scholar] [CrossRef] [Green Version]
  128. Buttgereit, F.; Aelion, J.; Rojkovich, B.; Zubrzycka-Sienkiewicz, A.; Radstake, T.; Chen, S.; Arikan, D.; Kupper, H.; Amital, H. OP0115 efficacy and safety of ABBV-3373, a novel anti-tnf glucocorticoid receptor modulator antibody drug conjugate, in patients with moderate to severe rheumatoid arthritis despite methotrexate therapy: A phase 2a proof of concept study. Ann. Rheum. Dis. 2021, 80, 64. [Google Scholar] [CrossRef]
  129. Lehar, S.M.; Pillow, T.H.; Xu, M.; Staben, L.; Kajihara, K.K.; Vandlen, R.; DePalatis, L.; Raab, H.; Hazenbos, W.L.; Morisaki, J.H.; et al. Novel antibody–antibiotic conjugate eliminates intracellular S. aureus. Nature 2015, 527, 323–328. [Google Scholar] [CrossRef]
  130. Cai, H.; Yip, V.; Lee, M.V.; Wong, S.; Saad, O.; Ma, S.; Ljumanovic, N.; Khojasteh, S.C.; Kamath, A.V.; Shen, B.-Q. Characterization of Tissue Distribution, Catabolism, and Elimination of an Anti-Staphylococcus aureus THIOMAB Antibody-Antibiotic Conjugate in Rats. Drug Metab. Dispos. 2020, 48, 1161–1168. [Google Scholar] [CrossRef]
  131. Zhou, C.; Lehar, S.; Gutierrez, J.; Rosenberger, C.M.; Ljumanovic, N.; Dinoso, J.; Koppada, N.; Hong, K.; Baruch, A.; Carrasco-Triguero, M.; et al. Pharmacokinetics and pharmacodynamics of DSTA4637A: A novel THIOMAB™ antibody antibiotic conjugate against Staphylococcus aureus in mice. mAbs 2016, 8, 1612–1619. [Google Scholar] [CrossRef] [Green Version]
  132. Zhou, C.; Cai, H.; Baruch, A.; Lewin-Koh, N.; Yang, M.; Guo, F.; Xu, D.; Deng, R.; Hazenbos, W.; Kamath, A.V. Sustained activity of novel THIOMAB antibody-antibiotic conjugate against Staphylococcus aureus in a mouse model: Longitudinal pharmacodynamic assessment by bioluminescence imaging. PLoS ONE 2019, 14, e0224096. [Google Scholar] [CrossRef] [PubMed]
  133. Deng, C.; Pan, B.; O’Connor, O.A. Brentuximab vedotin. Clin. Cancer Res. 2013, 19, 22–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Everts, M.; Kok, R.J.; Asgeirsdottir, S.A.; Melgert, B.N.; Moolenaar, T.J.; Koning, G.A.; van Luyn, M.J.; Meijer, D.K.; Molema, G. Selective intracellular delivery of dexamethasone into activated endothelial cells using an E-selectin-directed immunoconjugate. J. Immunol. 2002, 168, 883–889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Brandish, P.E.; Palmieri, A.; Antonenko, S.; Beaumont, M.; Benso, L.; Cancilla, M.; Cheng, M.; Fayadat-Dilman, L.; Feng, G.; Figueroa, I.; et al. Development of Anti-CD74 Antibody-Drug Conjugates to Target Glucocorticoids to Immune Cells. Bioconjug. Chem. 2018, 29, 2357–2369. [Google Scholar] [CrossRef]
  136. Kern, J.C.; Dooney, D.; Zhang, R.; Liang, L.; Brandish, P.E.; Cheng, M.; Feng, G.; Beck, A.; Bresson, D.; Firdos, J.; et al. Novel Phosphate Modified Cathepsin B Linkers: Improving Aqueous Solubility and Enhancing Payload Scope of ADCs. Bioconjug. Chem. 2016, 27, 2081–2088. [Google Scholar] [CrossRef]
  137. Wang, R.E.; Liu, T.; Wang, Y.; Cao, Y.; Du, J.; Luo, X.; Deshmukh, V.; Kim, C.H.; Lawson, B.R.; Tremblay, M.S.; et al. An immunosuppressive antibody-drug conjugate. J. Am. Chem. Soc. 2015, 137, 3229–3232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Yu, S.; Pearson, A.D.; Lim, R.K.; Rodgers, D.T.; Li, S.; Parker, H.B.; Weglarz, M.; Hampton, E.N.; Bollong, M.J.; Shen, J.; et al. Targeted Delivery of an Anti-inflammatory PDE4 Inhibitor to Immune Cells via an Antibody-drug Conjugate. Mol. Ther. 2016, 24, 2078–2089. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Lim, R.K.; Yu, S.; Cheng, B.; Li, S.; Kim, N.J.; Cao, Y.; Chi, V.; Kim, J.Y.; Chatterjee, A.K.; Schultz, P.G.; et al. Targeted Delivery of LXR Agonist Using a Site-Specific Antibody-Drug Conjugate. Bioconjug. Chem. 2015, 26, 2216–2222. [Google Scholar] [CrossRef] [PubMed]
  140. Sugo, T.; Terada, M.; Oikawa, T.; Miyata, K.; Nishimura, S.; Kenjo, E.; Ogasawara-Shimizu, M.; Makita, Y.; Imaichi, S.; Murata, S.; et al. Development of antibody-siRNA conjugate targeted to cardiac and skeletal muscles. J Control. Release 2016, 237, 1–13. [Google Scholar] [CrossRef] [PubMed]
  141. Ibtehaj, N.; Huda, R. High-dose BAFF receptor specific mAb-siRNA conjugate generates Fas-expressing B cells in lymph nodes and high-affinity serum autoantibody in a myasthenia mouse model. Clin. Immunol. 2017, 176, 122–130. [Google Scholar] [CrossRef]
  142. Yasunaga, M.; Manabe, S.; Matsumura, Y. Immunoregulation by IL-7R-targeting antibody-drug conjugates: Overcoming steroid-resistance in cancer and autoimmune disease. Sci. Rep. 2017, 7, 10735. [Google Scholar] [CrossRef] [Green Version]
  143. Czechowicz, A.; Palchaudhuri, R.; Scheck, A.; Hu, Y.; Hoggatt, J.; Saez, B.; Pang, W.W.; Mansour, M.K.; Tate, T.A.; Chan, Y.Y.; et al. Selective hematopoietic stem cell ablation using CD117-antibody-drug-conjugates enables safe and effective transplantation with immunity preservation. Nat. Commun. 2019, 10, 617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Palchaudhuri, R.; Saez, B.; Hoggatt, J.; Schajnovitz, A.; Sykes, D.B.; Tate, T.A.; Czechowicz, A.; Kfoury, Y.; Ruchika, F.; Rossi, D.J.; et al. Non-genotoxic conditioning for hematopoietic stem cell transplantation using a hematopoietic-cell-specific internalizing immunotoxin. Nat. Biotechnol. 2016, 34, 738–745. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Lee, H.; Bhang, S.H.; Lee, J.H.; Kim, H.; Hahn, S.K. Tocilizumab-Alendronate Conjugate for Treatment of Rheumatoid Arthritis. Bioconjug. Chem. 2017, 28, 1084–1092. [Google Scholar] [CrossRef] [PubMed]
  146. Mehta, G.; Scheinman, R.I.; Holers, V.M.; Banda, N.K. A New Approach for the Treatment of Arthritis in Mice with a Novel Conjugate of an Anti-C5aR1 Antibody and C5 Small Interfering RNA. J. Immunol. 2015, 194, 5446–5454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Nagai, T.; Tanaka, M.; Tsuneyoshi, Y.; Matsushita, K.; Sunahara, N.; Matsuda, T.; Yoshida, H.; Komiya, S.; Onda, M.; Matsuyama, T. In vitro and in vivo efficacy of a recombinant immunotoxin against folate receptor beta on the activation and proliferation of rheumatoid arthritis synovial cells. Arthritis Rheum. 2006, 54, 3126–3134. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Simplified design and favorable characteristics of ADCs’ components.
Figure 1. Simplified design and favorable characteristics of ADCs’ components.
Vaccines 09 01111 g001
Table 1. Current indications and characteristics of FDA-approved ADCs.
Table 1. Current indications and characteristics of FDA-approved ADCs.
ADCManufacturerYear of Initial FDA ApprovalIndicationsTargetAntibodyPayloadLinkerDARCommon Adverse Events (>10%)
Gemtuzumab Ozogamicin
(Mylotarg®, CMA-676)
Pfizer/Wyeth2000, withdrawn 2010, re-approved 2017Newly diagnosed (de novo) CD33+ AML in adults (as a monotherapy or combined with chemotherapy) and pediatric patients 1 month and older (combined with chemotherapy) and relapsed/refractory CD33+ AML in adults and pediatric patients ≥ 2 years of ageCD33Humanized
IgG4
Calicheamicin derivative
Acid-labile hydrazone linker~2–3Infection, hemorrhage, thrombocytopenia, hypophosphatemia, hypokalemia, hyponatremia, nausea, vomiting, elevated ALP, elevated aminotransferase, fatigue, febrile neutropenia, constipation, abdominal pain, pyrexia, mucositis
Brentuximab Vedotin
(Adcetris®, SGN-35)
Seattle Genetics, Millennium/
Takeda
2011Previously untreated Stage III/IV cHL (combined with chemotherapy), cHL at high risk of relapse or progression as post-auto-HSCT consolidation, cHL after failure of auto-HSCT or after failure of ≥ 2 prior chemotherapy regimens, previously untreated sALCL or other CD30+ peripheral T-cell lymphomas (combined with chemotherapy), relapsed sALCL, relapsed peripheral cutaneous ALCL or CD30+ MFCD30Chimeric IgG1MMAEProtease-cleavable dipeptide
(Val-Cit) linker
~4Neutropenia, peripheral sensory neuropathy, fatigue, upper respiratory tract infection, nausea, diarrhea, anemia, thrombocytopenia, pyrexia, rash, abdominal pain, vomiting, arthralgia, myalgia, pruritus, peripheral motor neuropathy, headache, constipation, dizziness, lymphadenopathy, dyspnea, back pain, anxiety
Ado-trastuzumab emtansine
(T-DM1, Kadcyla®)
Genentech, Roche2013Unresectable locally advanced or metastatic HER2+ breast cancer, previously treated with trastuzumab and a taxane, adjuvant treatment for HER2+ early breast cancer with residual invasive disease after neoadjuvant taxane and trastuzumabHER2/ERB2Humanized IgG1DM1Thioether (non-cleavable) linker3.5 Nausea, constipation, diarrhea, vomiting, abdominal pain, dry mouth, stomatitis, headache, peripheral neuropathy, dizziness, epistaxis, cough, dyspnea, fatigue, musculoskeletal pain, arthralgia, myalgia, pyrexia, thrombocytopenia, anemia, increased aminotransferases, insomnia, rash, hypokalemia
Inotuzumab ozogamicin
(Besponsa®, CMC-544)
Pfizer/Wyeth2017Relapsed or refractory CD22+ B-cell precursor ALL in adultsCD22Humanized IgG4Calicheamicin
derivative
Acid-labile hydrazone linker~4Thrombocytopenia, neutropenia, infection, anemia, leukopenia, nausea, fatigue, hemorrhage, pyrexia, elevated transaminases, febrile neutropenia, elevated gamma-glutamyltransferase, lymphopenia, headache, abdominal pain, diarrhea, constipation, vomiting, stomatitis, elevated ALP
Polatuzumab vedotin-piiq
(Polivy®, DCDS4501A, RG7596)
Genentech, Roche2019Relapsed or refractory diffuse large B-cell lymphoma (combined with bendamustine and rituximab) in adult patients after ≥ 2 prior therapiesCD79bHumanized IgG1MMAEProtease-cleavable dipeptide
(Val-Cit) linker
3.5Neutropenia, thrombocytopenia, anemia, leukopenia, lymphopenia, febrile neutropenia, peripheral neuropathy, dizziness, diarrhea, vomiting, infusion-related reactions, pyrexia, decreased appetite, fatigue, pneumonia, upper respiratory tract infection, decreased weight, hypokalemia, hypoalbuminemia, hypocalcemia
Enfortumab vedotin
(Padcev®, AGS-22M6E, AGS-22CE)
Astellas/
Seattle Genetics
2019Locally advanced or metastatic urothelial cancer in adult patients who had received prior treatment with a PD-1/L1 inhibitor and platinum-based chemotherapy in neoadjuvant/adjuvant settingNectin 4Fully human IgG1MMAEProtease-cleavable dipeptide
(Val-Cit) linker
~3.8 Peripheral neuropathy, dysgeusia, fatigue, decreased appetite, rash, alopecia, dry skin, pruritus, dry eye, nausea, vomiting, constipation
Fam-trastuzumab deruxtecan-nxki
(Enhertu®, DS-8201a, T-DXd)
AstraZeneca/Daiichi Sankyo2019Unresectable or metastatic HER2+ breast cancer in adult patients who have previously received ≥ 2 HER2 blockade regimens in the metastatic setting, locally advanced or metastatic HER2+ gastric or gastroesophageal adenocarcinoma after trastuzumab-based treatmentHER2/ERB2Humanized IgG1DXd
(exatecan derivative)
Protease-cleavable tetrapeptide (Gly-Gly-Phe-Gly) linker7–8Nausea, vomiting, constipation, diarrhea, abdominal pain, stomatitis, dyspepsia, fatigue, alopecia, rash, decreased appetite, hypokalemia, anemia, neutropenia, leukopenia, thrombocytopenia, cough, dyspnea, epistaxis, headache, dizziness, upper respiratory tract infection, dry eye
Sacituzumab govitecan-hziy
(Trodelvy® IMMU-132, HRS7-SN38)
Immunomedics2020Unresectable locally advanced or metastatic triple negative (HR-/HER2-) breast cancer after ≥2 prior systemic therapies, locally advanced or metastatic urothelial carcinoma after platinum-based chemotherapy and either a PD-1 or PD-L1 inhibitor Trop-2Humanized IgG1SN-38HydrolysableCL2A linker7.6Nausea, diarrhea, neutropenia, fatigue, anemia, vomiting, constipation, alopecia, rash, headache, respiratory tract infection, decreased appetite, urinary tract infection, hyperglycemia, arthralgia, dyspnea, dizziness, neuropathy, back pain, edema, thrombocytopenia, hypomagnesemia, hypokalemia, hypophosphatemia, pruritus, mucositis
Belantamab mafadotin-blm

(Blenrep®, GSK2857916)
GlaxoSmithKline2020Relapsed or refractory multiple myeloma in adult patients who have received ≥ 4 therapies, including an anti-CD38 mAb, a proteasome inhibitor and an immunomodulatory agentBCMAAfucosylated Humanized IgG1MMAFMaleimidocaproyl (mc) linker~4Keratopathy, decreased visual acuity, blurred vision, dry eyes, nausea, diarrhea, constipation, blurred vision, pyrexia, infusion-related reactions, arthralgia, back pain, upper respiratory tract infections, decreased appetite, fatigue
Loncastuximab tesirine-lpyl

(Zynlonta®, ADCT-402)
ADC Therapeutics2021Relapsed or refractory large B-cell lymphoma in adult patients after ≥ 2 lines of systemic therapyCD19Humanized IgG1PDB dimer SCX (SG3199) Protease-cleavable valine-alanine linker2.8Fatigue, edema, rash, pruritus, nausea, diarrea, abdominal pain, vomiting, constipation, musculoskeletal pain, decreased appetite, dyspnea, pleural effusion, upper respiratory tract infection
Abbreviations: DAR: drug-to-antibody ratio; hSCT: Hematopoietic Stem Cell Transplant; MMAE: Monomethyl Auristatin E; MMAF: Monomethyl Auristatin F; DM1: maytansine 1 derivative; HER2: human epidermal growth factor receptor 2; ALP: alkaline phosphatase; AML: acute myeloid leukemia; cHL: classic Hodgkin lymphoma; sALCL: systemic anaplastic large cell lymphoma; MF: mycosis fungoides; ALL: acute lymphoblastic lymphoma; HR: hormonal receptor; mAB: monoclonal antibody; PBD: pyrrolobenzodiazepine.
Table 2. Summary of non-approved ADCs that have been granted breakthrough therapy designation, fast-track designation or priority review by the US FDA and their current clinical trial status.
Table 2. Summary of non-approved ADCs that have been granted breakthrough therapy designation, fast-track designation or priority review by the US FDA and their current clinical trial status.
ADC NameFDA StatusTarget/PayloadNCT NumberCurrent Trial StatusIndicationAssigned Interventions
Tisotumab vedotin (TF-011-MMAE)Priority review granted in April 2021Tissue factor/MMAENCT04697628 (innovaTV 301)3Second or Third-line Recurrent or Metastatic Cervical cancer Tisotumab vedotin 2.0 mg/kg IV Q3W
Topotecan 1 or 1.25 mg/m2 IV on D1-5 Q3W
Vinorelbine 30 mg/m2 IV on D1 and 8 Q3W
Gemcitabine 1000 mg/m2 IV on D1 and 8 Q3W
Irinotecan 100 or 125 mg/m2 IV weekly for 28 days, Q6W
Pemetrexed 500 mg/m2 IV on D1 Q3W
Trastuzumab duocarmazine (SYD985)Fast Track designation granted in January 2018HER2/seco-DUBANCT03262935 (TULIP)3HER2+ unresectable locally advanced or metastatic breast cancerTrastuzumab duocarmazine RP2D 1.2 mg/kg IV Q3W
Lapatinib + capecitabine or
Trastuzumab + capecitabine or
Trastuzumab + vinorelbine
Trastuzumab + eribulin
Mirvetuximab soravtansine (IMGN853)Fast Track designation granted in June 2018. Accelerated approval pathway includes pivotal trial SORAYA and confirmatory trial MIRASOL
Folate receptor α/DM4NCT04296890 (SORAYA)3Platinum-resistant advanced high-grade epithelial ovarian, primary peritoneal, or fallopian tube cancer, with high Folate receptor α expressionMirvetuximab soravtasine 6 mg/kg IV Q3W
NCT04209855
(MIRASOL)
3Platinum-resistant advanced high-grade epithelial ovarian, primary peritoneal or fallopian tube cancer with high Folate receptor α expressionMirvetuximab soravtasine 6 mg/kg IV Q3W
Paclitaxel 80 mg/m^2 QW within a 4-week cycle or
Pegylated liposomal doxorubicin 40 mg/m^2 Q4W or
Topotexan 4 mg/m^2 IV either on D1, 8, 15 Q4W or 1.25 mg/m^2 on D1-5 Q3W
Upifitamab rilsodotin (XMT-1536)Fast Track designation granted in August 2020NaPi2b/DolaLock (auristatin F- hydroxypropylamide payload molecules)NCT03319628 (UPLIFT; Pivotal Cohort)1b/2 Platinum-resistant ovarian cancer and non-small cell lung cancer, adenocarcinoma subtype Upifitamab rilsodotin RP2D IV Q4W
Disitamab vedotin (RC48) Breakthrough Therapy designation granted in September 2020HER2/MMAENCT048793292 HER2+ locally advanced or metastatic urothelial carcinoma in second-line treatment of patients pre-treated with platinum-containing chemotherapy Disitamab vedotin 2.0 mg/kg IV once every 2 weeks (maximum dose 200 mg)
IMGN632 Breakthrough Therapy designation granted in October 2020CD123/DNA mono-alkylating payload of the indolinobenzodiazepine pseudodimer (IGN) class.NCT033865131/2Relapsed or refractory or Untreated blastic plasmacytoid dendritic cell neoplasm (BPDCN)IMGN632 IV
ARX788Fast Track designation granted in January 2021HER2/MMAF NCT04829604 (ACE-Breast03)2HER2+ metastatic breast cancer, resistant or refractory to T-DM1, and/or T-DXd, and/or tucatinib-containing regimensARX788 IV Q4W
Abbreviations: MMAE: monomethyl auristatin E; Q3W: every 3 weeks (21 days); IV: intravenously; D: day(s); Q6W: every 6 weeks (42 days); HER2: human epidermal growth factor receptor 2; seco-DUBA: seco-duocarmycin-hydroxybenzamide-azaindole; RP2D: recommended phase 2 dose; QW: once per week; Q4W: every 4 weeks (28 days); NaPi2b: sodium-dependent phosphate transport protein 2B; MMAF: monomethyl auristatin F.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Theocharopoulos, C.; Lialios, P.-P.; Samarkos, M.; Gogas, H.; Ziogas, D.C. Antibody-Drug Conjugates: Functional Principles and Applications in Oncology and Beyond. Vaccines 2021, 9, 1111. https://0-doi-org.brum.beds.ac.uk/10.3390/vaccines9101111

AMA Style

Theocharopoulos C, Lialios P-P, Samarkos M, Gogas H, Ziogas DC. Antibody-Drug Conjugates: Functional Principles and Applications in Oncology and Beyond. Vaccines. 2021; 9(10):1111. https://0-doi-org.brum.beds.ac.uk/10.3390/vaccines9101111

Chicago/Turabian Style

Theocharopoulos, Charalampos, Panagiotis-Petros Lialios, Michael Samarkos, Helen Gogas, and Dimitrios C. Ziogas. 2021. "Antibody-Drug Conjugates: Functional Principles and Applications in Oncology and Beyond" Vaccines 9, no. 10: 1111. https://0-doi-org.brum.beds.ac.uk/10.3390/vaccines9101111

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