Antibody–Drug Conjugates and Small Molecule–Drug Conjugates

Jun 16, 2015 - ... Most Read Articles · Author Index · Cover Art Gallery · Virtual Issues .... A number of independent studies have found that miniant...
1 downloads 3 Views 10MB Size
Subscriber access provided by NEW YORK UNIV

Perspective

Antibody-drug conjugates and small molecule-drug conjugates: opportunities and challenges for the development of selective anti-cancer cytotoxic agents Giulio Casi, and Dario Neri J. Med. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jmedchem.5b00457 • Publication Date (Web): 16 Jun 2015 Downloaded from http://pubs.acs.org on June 22, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Medicinal Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Antibody-drug conjugates and small molecule-drug conjugates: opportunities and challenges for the development of selective anti-cancer cytotoxic agents

Giulio Casi (1) and Dario Neri * (2)

(1) Philochem AG, Libernstrasse 3, CH8112, Otelfingen, Switzerland. (2) Institute of Pharmaceutical Sciences, Department of Chemistry and Applied Biosciences, ETH Zürich, Vladimir-Prelog-Weg 1-5/10, 8093 Zürich, Switzerland

Keywords: Antibody-drug conjugates; small molecule-drug conjugates; tumor targeting; antibody internalization; targeted cytotoxics

Conflict of interest disclosure: G.C. works at Philochem AG, D.N. is a co-founder and shareholder of Philogen SpA.

1 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract Conventional cancer chemotherapy heavily relies on the use of cytotoxic agents, which typically do not preferentially localize at the tumor site and cause toxicity to normal organs, preventing dose escalation to therapeutically active regimens. In principle, antibodies and other ligands could be used for the selective pharmacodelivery of cytotoxic agents to the neoplastic mass. For many years, the availability of ligands, capable of selective internalization into tumor cells, has been considered to be an essential requirement for the development of targeted cytotoxics. This assumption, however, has recently been challenged on the basis of therapeutic data obtained with non-internalizing drug conjugates. Moreover, quantitative evaluations of the tumor targeting properties of antibodies and of small organic ligands have provided new insights for the implementation of optimal strategies for the development of targeted cytotoxics. In this article, we highlight opportunities and challenges associated with the clinical and industrial development of antibody-drug conjugates and small molecule-drug conjugates for cancer therapy.

2 ACS Paragon Plus Environment

Page 2 of 39

Page 3 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Conventional anti-cancer drugs and the first series of antibody-drug conjugates Conventional cancer chemotherapy typically relies on the use of small-molecule drugs (e.g., cytotoxic agents, kinase inhibitors), which inhibit the growth of cells undergoing rapid proliferation (or even kill them), thus offering a potential therapeutic benefit to cancer patients. The combination of highly toxic agents has revolutionized the treatment of many hematological malignancies, some of which are now curable. The induction of cancer cures for disseminated solid tumors by means of cytotoxic agents is more difficult, but some metastatic malignancies (e.g., metastatic testicular cancer)

1

can be eradicated by a combination of chemotherapeutic

drugs.

Unfortunately, the majority of patients with metastatic solid tumors die from the disease, after chemotherapeutic regimens have failed to induce objective response or when resistance develops. 2

One of the most serious limitations of conventional cancer chemotherapy relates to the fact that

most small molecule therapeutic agents (e.g., cytotoxic agents) do not preferentially localize at the tumor site. This pharmacokinetic limitation has long been appreciated in mouse models of cancer.

3, 4

Importantly, recent PET studies with radiolabeled drugs in cancer patients have

unambiguously shown that cytotoxic agents accumulate in certain healthy organs (e.g., those involved in the excretion process; Figure 1), while failing to efficiently target neoplastic masses in vivo. 5, 6

3 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1: PET images at multiple time points of a patient with metastatic mesothelioma, injected with

Page 4 of 39

11

C-docetaxel.

Drug uptake in the neoplastic lesions is not visible, while accumulation in the liver and in some other normal organ structures can be seen at multiple time points. The arrow to the chest indicates the pleural localization of the mesothelium. Reproduced with permission from van der Veldt, A.A.M. et al (2010) Eur. J. Nucl. Med. Mol. Imaging 37, 1950-1958.

Monoclonal antibodies have been considered as possible vehicles for the selective in vivo pharmacodelivery of drugs, under the assumption that the exquisite binding specificity of the antibody molecule would facilitate a selective accumulation at the site of disease. The history of the development of antibody-drug conjugates (ADCs) has extensively been reviewed elsewhere 79

and will not be repeated here. Nonetheless, a few aspects of antibody-based drug delivery

deserve a closer analysis:

4 ACS Paragon Plus Environment

Page 5 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(i)

The real ability of antibody to preferentially localize at the site of disease, which is essential for the implementation of efficient pharmacodelivery strategies in cancer patients, remains largely unexplored in clinical trials. Few monoclonal antibodies have been studied with nuclear medicine procedures and adequate dosimetric analyses in a sufficiently large number of patients.10-19 In those studies, a large variability in tumor uptake, from patient to patient and from lesion to lesion, was observed. Antibodies which display tumor:organ and tumor:blood ratios > 5:1, 24 hours after intravenous injection, represent the exception rather than the rule.

(ii)

The antibody format (scFv, diabody, miniantibody or small immune proteins (SIP), IgG; Figure 2a) greatly impacts on pharmacokinetic properties and on tumor uptake. A number of independent studies have found that mini-antibodies or SIPs may exhibit the best compromise between efficient tumor uptake and a sufficiently rapid clearance from circulation.20-22 Nonetheless, if maximal antibody accumulation at the tumor site is desired, the IgG format may represent the preferred one, as its long circulation time in blood counterbalances the inherently slow extravasation process of proteins and of antibodies.21, 23, 24 However, in humans, typically more than 99% of the antibody administered, does not reach the tumor.

10-19

The persistent high-levels

of drug-conjugate in blood may be responsible for undesired toxicities as a consequence of both premature, and undesired drug release in off-target organs (e.g. endothelium, liver; Figure 2b). (iii)

A microscopic analysis of antibody localization within the tumor mass exhibits a preferential accumulation on perivascular tumor cells.25 This feature, which may hinder certain pharmacodelivery strategies, most probably results from the inefficient

5 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 39

extravasation of antibody molecules and by their trapping by antigen on perivascular tumor cells (the so-called “antigen barrier effect”;26-28 Figure 2c). (iv)

While initially low-potency drugs (e.g., doxorubicin) were used for ADC development programs,29 the field has gradually progressed towards the use of ultrapotent payloads, often capable of killing cells at sub-nanomolar concentrations. The use of highly cytotoxic drugs is also justified by cost-of-goods considerations, as monoclonal antibodies (in IgG format) have a molecular weight of 150’000 Da, while drugs are typically smaller than 1’000 Da. It is probably not economically feasible to administer doses of more than one gram of antibody-drug conjugate to a patient. Less potent drugs may be used for ADC applications, if multivalent (e.g., dendrimeric) linkers are used, leading to high drug:antibody ratios.30

a)

b)

c)

Figure 2: (a) Schematic representation showing domain composition of engineered fragments including scFv (25 kDa), diabody (55 kDa), SIP/minibody (80 kDa) and intact antibodies (150 kDa). (b) Macroscopic distribution of

64

Cu-

DOTA-trastuzumab in breast cancer patients (1, 24 and 48 h) following intravenous administration. The majority of injected antibody molecules do not reach their target in vivo: virtually all of them accumulate (at least transiently) in excretory organs (liver for intact antibodies, kidneys for small antibody fragments)(c) microscopic images of tumor tissue after trastuzumab-FITC conjugate injection (green). Tumor blood vessels (red) and nuclei (blue) were also stained. The image depicts trapping of Trastuzumab in the vicinity of blood vessels (Adapted by permission from the

6 ACS Paragon Plus Environment

Page 7 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

American Association for Cancer Research: Dennis, M. S. et al. Imaging tumors with an albumin-binding Fab, a novel tumor-targeting agent, Cancer Res, January 1, 2007, 67, 254–261, 10.1158/0008-5472.can-06-2531).

Two ADC products are currently available on the market for cancer therapy applications: brentuximab vedotin and trastuzumab emtansine.

Brentuximab vedotin is an anti-CD30 antibody, coupled to the MMAE auristatin, by a cleavable linker comprising a valine-citruline moiety and a self-immolating spacer (Figure 3). The product has been approved for the treatment of last-line Hodgkin lymphoma and anaplastic large cell lymphoma, on the basis of compelling objective responses observed in phase II clinical trials in heavily pretreated patients.31 In particular, brentuximab vedotin (when dosed at 1.8 mg/Kg, every three weeks) mediated 75% overall responses in Hodgkin lymphoma patients that had failed autologous stem cell transplant, with 34% of the patients experiencing complete responses which lasted on average 29 months. In the anaplastic large cell lymphoma settings, brentuximab vedotin provided a substantial benefit to patients that failed at least one multi-agent chemotherapy regimen, with 87% overall response rates, of which 57% were complete responses, and a progression free survival of 12.6 months. Importantly the treatments were well tolerated, with the most common side effects reported as neutropenia, peripheral sensory neuropathy, fatigue, nausea, fever and thrombocytopenia. These encouraging data have recently been complemented by the results of a randomized phase III clinical trial in the consolidation setting for Hodgkin lymphoma patients who had received a bone marrow transplant. A new phase III trials that features a combination of brentuximab vedotin and chemotherapy was recently started in younger patients with newly diagnosed Hodgkin lymphoma (NCT-02166463).

7 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Trastuzumab emtansine consists of the anti-HER2 trastuzumab antibody, armed with the maytansinoid DM1, equipped with a non-cleavable linker (Figure 3). It has been proposed that the drug may be released after product internalization, upon intracellular proteolytic digestion of the antibody moiety.32 33 However, a certain instability of the ADC in blood has been documented, with maleimide elimination occurring slowly under physiological conditions.34 Trastuzumab emtansine has received marketing authorization for the second-line treatment of a subset of patients with metastatic breast cancer. Specifically, it has been approved for the treatment of patients with HER2-positive metastatic breast cancer that has previously been treated with Herceptin and a taxane chemotherapy. Trastuzumab emtansine showed convincing clinical efficacy in a phase III clinical study in 991 patients treated with 3.6 mg /Kg of ADC every three weeks, in comparison to established chemotherapy with lapatinib and capecitabine. Progression free survival (9.6 months versus 6.4 months) and overall survival (30.9 months versus 25.1 months) were improved in the ADC-treatment arm. Safety was significantly improved compared to chemotherapy, with substantially fewer grade 3/4 adverse events.35 In a recent front-line phase III clinical study in breast cancer patients, however, trastuzumab emtansine was not found to be superior to the corresponding naked trastuzumab antibody, used in combination with either docetaxel or paclitaxel chemotherapy.

Figure 3: Structures of the two antibody-drug conjugate products currently available on the market for cancer therapy: Brentuximab vedotin (left) and Trastuzumab emtansine (right).

8 ACS Paragon Plus Environment

Page 8 of 39

Page 9 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Trastuzumab emtansine is used in the clinic at 3.6 mg/Kg doses, while tumor eradications with the same product have been observed in mice at doses of 15 mg/Kg.36 Chemically defined antiHer2 ADCs, with precise control of site and stoichiometry of drug conjugation, displayed improved half-life, efficacy and safety, relative to conventional heterogeneous ADCs, in rodent xenograft models.37-40 Large clinical experience is available from gemtuzumab ozogamicin, a humanized IgG4 kappa CD33 antibody conjugated to the DNA-binder calicheamicin-γ1 via a hydrolysis sensitive hydrazone. Gemtuzumab ozogamicin received accelerated US marketing approval in 2000 for the treatment of patients of 60 years of age and older with CD33+ AML in first relapse, who were not eligible for standard chemotherapy. Subsequent trials failed to confirm therapeutic benefits at doses of 9 mg/m2 when GO was used in combination with chemotherapy. These data and the slight increase in deaths in the group receiving the ADC compared to those receiving chemotherapy, led to voluntary withdrawal of the drug in 2010 from the market. Interestingly the acid-sensitive hydrazone was required to elicit potent antitumor activity in an antigen independent manner on solid tumors in vivo.41 Gemtuzumab ozogamicin is also active in AML patients whose cells apparently do not express CD33.

42

The product remains available for compassionate use,

and recent studies have showed promising efficacy when the ADC was administered either following a fractionated schedule

43, 44

or to patients with more favorable prognosis AML during

standard induction therapy.45-48

Additionally, many promising ADC products can eradicate tumors in mice, at doses which are well tolerated. For example, an anti-CD33 antibody, coupled to a potent DNA alkylating drug of the pyrrolo-benzodiazepine family, has recently been shown to eradicate subcutaneously-grafted AML chloroma lesions in mice at doses as low as 0.1 mg/Kg.49

9 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

It remains to be seen, however, how easily these promising preclinical therapeutic observations can be translated to the human situation. For example, if the tumor targeting performance (e.g., tumor uptake) of an ADC is much better in rodents than in patients, the corresponding therapeutic behavior may be different in the two species.

The need for antibody internalization: fiction or reality? Virtually all small molecule anti-cancer drugs exert their activity by binding to an intracellular target. For this reason, it has been widely assumed that the use of internalizing pharmacodelivery vehicles may represent an absolute requirement for ADC development. Indeed, it has been claimed that “targeting an ADC to a noninternalizing target antigen with the expectation that extracellulary released drug will diffuse into the target cell is not a recipe for a successful ADC“.50 In principle, the ability to deliver cytotoxic agents only to target cells which express the antigen recognized by the cognate antibody, sparing all other cells in the body, would represent an extremely attractive therapeutic feature. In practice, however, several factors limit the performance of internalizing ADCs, including the moderate accumulation of antibody molecules at the tumor site (typically < 0.1 % injected dose per gram of tumor in humans)10-19 and the heterogenous antibody distribution within the tumor mass. In addition, while the antibody internalization process can easily be studied in vitro, the same process often remains a “black box” in vivo, as it is difficult to recover and analyze neoplastic masses from patients.

Recently, an increasing body of experimental evidence has shown that certain ADCs, based on non-internalizing ligands, can efficiently deliver cytotoxic agents to neoplastic masses.51-55 When suitable linker-payload combinations are used, these targeted cytotoxics can liberate their payload

10 ACS Paragon Plus Environment

Page 10 of 39

Page 11 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

at the tumor site and display a potent anti-cancer activity in mouse models of cancer.56-58 In principle, it should be possible to use antibodies for the efficient delivery and release of drugs in the extracellular tumor space (e.g., through proteolytic cleavage or by reduction of disulfide bonds).56-60 The drugs could then internalize into tumor cells or other cellular targets (e.g., tumor endothelial cells, tumor resident leukocytes), causing localized damage. The therapeutic performance of these ADCs is likely to depend on many factors, including the antibody’s tumor targeting efficiency, the kinetics of drug release, as well as the diffusion properties of the drug payload.

Splice isoforms of fibronectin and of tenascin-C are abundantly found in the tumor subendothelial extracellular matrix making them attractive non-internalizing targets for pharmacodelivery applications.51-55 These antigens are more accessible, abundant, and genetically stable than tumorassociated antigens located on the cell membrane, thus facilitating the development of ADC products with long residence time at the tumor site. The ADCs would release their payload in the extracellular space, facilitating a by-stander effect on neighboring tumor cells. As antibody internalization would no longer be needed, biodistribution and imaging data, obtained with noninternalizing antibodies, could be more easily integrated into modeling studies for the prediction of ADC in vivo performance.

Small ligands and peptides as alternative to antibodies for pharmacodelivery applications In principle, small organic ligands and peptides could be used as alternatives to antibodies for tumor targeting applications. While high-affinity human monoclonal antibodies can be raised against virtually all protein targets (for example, by panning combinatorial phage display libraries )61 it is not always possible to isolate small organic or peptidic ligands to proteins of

11 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

pharmaceutical interest. Nonetheless, a number of small ligands to tumor-associated antigens have been characterized in terms of their tumor targeting properties and for pharmacodelivery applications.4, 62 Some of the most promising compounds in this category are depicted in Figure 4. They include folate derivatives (binding to the folate receptor, a protein over-expressed in ovarian cancer and in other malignancies)63, glutamic acid urea derivatives (binding to the prostate specific membrane antigen, a surface marker of prostate cancer cells)64-66, somatostatin analogues (particularly indicated for neuroendocrine tumors)67 and certain aromatic sulfonamides, specific to carbonic anhydrase IX (a marker of hypoxia and of renal cell carcinoma).58, 68-71

Figure 4 Structures of relevant classes of small molecule tumor targeting moieties. For a more comprehensive list of small molecule-drug conjugates (SMDCs) in preclinical and clinical development, see Low et al. (2015) [Ref. 63].

12 ACS Paragon Plus Environment

Page 12 of 39

Page 13 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

The opportunity of using small ligands for pharmacodelivery applications has recently been reviewed in detail.4, 62 There is a strong evidence that ligands with a dissociation constant to their target of 10 nM or better may be needed, in order to achieve an efficient tumor targeting performance.62, 70 In some cases, multivalent binding (“avidity”) may compensate for insufficient affinity,69, 72-74 but almost all products in clinical development feature a single tumor-targeting moiety, coupled to a potent cytotoxic payload.62 Traditionally, the discovery of small ligands to tumor associated antigens has been difficult and has often relied on the synthesis of analogues to naturally-occurring ligands. However, progress in DNA-encoded chemical library technologies promises to facilitate the hit discovery process.4 75-77

Indeed, the tagging of individual compounds with DNA-fragments, serving as amplifiable

identification barcodes, in split&pool synthetic strategies, allows the construction and screening of chemical libraries containing millions of molecules. Similarly, the discovery of high-affinity binding cyclic peptides has been greatly facilitated by the construction and selection of very large combinatorial libraries, featuring the incorporation of non-natural aminoacids 78 or the cyclization of cysteine-containing peptides with reactive chemical moieties serving as structural scaffolds.79

Folic acid derivatives probably represent the first small molecule ligands, which have successfully been used for the selective pharmacodelivery of cytotoxic drugs and other payloads to the tumor environment.62 The folic acid receptor is present at low levels in most normal tissues, with the exception of kidney, spleen and lung tissue.80 By contrast, the same target is strongly upregulated in many epithelial cancers, including tumors of the ovary, breast, lung, colon and kidney.62 Vintafolide (also known as EC145)81 is a conjugate consisting of folic acid as a targeting moiety, connected to desacetylvinblastine as toxic payload through a linker comprising a charged peptide linker and a disulfide bond with a self-immolative spacer for drug release

13 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(Figure 4). Endocyte developed Vintafolide for several indications. Initially the conjugate was studied for the treatment of patients with platinum-resistant ovarian cancer in combination with pegylated lisosomal doxorubicin (PLD) versus PLD alone. The combination extended progression-free survival for approximately 2 months in direct correlation with folate receptor expression.82 Moreover, Endocyte has recently reported promising results of an interim analysis of phase II data, comparing the use of Vintafolide plus docetaxel versus docetaxel alone for the treatment of patients with non-small cell lung cancer. The combination treatment extended progression-free survival by 1.2 months and overall survival by 5.9 months.62, 83 Vintafolide is currently developed in combination with Doxil in a phase III trial for the treatment of platinumresistant ovarian cancer; the enrolment in this trial was suspended in May 2014 following the recommendation from the data safety monitoring board (DSMB) based on futility. The clinical data are currently being reviewed.62 Endocyte and Bristol-Myers Squibb have also investigated conjugates based on other payloads, such as tubulysins or epothilones, in clinical trials.62

Prostate-specific membrane antigen (PSMA) is an accessible cell-surface bound homodimeric endopeptidase, which can be efficiently targeted using urea-based derivatives of glutamic acid (Figure 4). PSMA is usually absent in normal adult tissues, exception made for normal prostate tissue and duodenum.

84, 85

In most cases PSMA-selective imaging agents showed high uptake in

salivary and lacrimal glands, which may reflect either target expression in those tissues or ligand binding to non-cognate proteins.65, 86 By contrast, the antigen is strongly expressed in the majority of localized and metastatic prostate cancer cells.87, 88 PSMA has also been reported to be overexpressed in the neo-vasculature of many malignancies.89, 90 The tumor-targeting performance of

14 ACS Paragon Plus Environment

Page 14 of 39

Page 15 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

various PSMA ligands has been validated in rodents and in cancer patients, using nuclear medicine procedures.65, 86 64, 66

Carbonic anhydrase IX (CAIX) represents an additional example of a tumor-associated antigen, which can be targeted both with antibodies and with small organic ligands. Several types of carbonic anhydrases can be found inside (CAI, CAII, CAIII, CAVII and CAXIII), or on the surface (CAIV, CAIX, CAXII, CAXIV and CAXV) of cells. Interestingly, CAIX is virtually undetectable in normal adult tissues, except for a strong expression in stomach, duodenum and gall bladder.68, 84, 91 By contrast, CAIX is over-expressed in the majority of clear-cell carcinomas, as a result of a von Hippel-Lindau mutation or deletion.92

93

Additionally, the antigen can be

found at sites of hypoxia.94, 95 An ADC product targeting CAIX has been brought to clinical trials by Bayer.96 Certain small aromatic sulfonamides (in particular, acetazolamide; Figure 4) can be used as delivery vehicles for tumor targeting applications.

69, 97

For example, an acetazolamide

moiety coupled to a near-infrared fluorophore has been shown to selectively localize to subcutaneously grafted human renal cell carcinomas.58 The improvement of acetazolamide affinity to its target, using DNA-encoded chemical library technology, led to a further increase in tumor uptake ( Figure 5).70

15 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 39

Figure 5 Evaluation of targeting performance of IRDye 750 conjugates in near-infrared fluorescence imaging of balb/c nu/nu mice bearing SK-RC-52 xenografts. 3 nmol of ligand-IRDye 750 conjugate were injected: untargeted control, targeted based on acetazolamide 1 (b) and targeted based on pharmacophore pair 2 (c). Chemical structures of 1 and 2 are provided in Figure 4. . Reproduced with permission from Wichert, M. et al (2015) Nat. Chem. 7, 241-249.

In our experience, CAIX ligands do not efficiently internalize into cancer cells, but they can be used for the selective delivery of cytotoxic payloads (e.g., DM1, MMAE) to the tumor environment, using disulfide-based or peptide-based cleavable linkers,

58, 69

[Pretto, F. Casi, G.

unpublished data] with promising therapeutic results. A direct comparison of tumor targeting experiments based on two high-affinity CAIX antibodies or on acetazolamide derivatives revealed a superiority of the small ligands, in terms of speed of targeting, tumor uptake and selectivity

70

[N. Krall, F. Pretto, M. Mattarella, D. Neri, manuscript submitted; L. Gualandi, C.

Hutchinson, F. Pretto, D. Neri, S. Wulhfard, unpublished data].

Challenges and opportunities for the development of next-generation cytotoxic agents As mentioned previously, the high efficacy of targeted therapies often seen in rodent models of cancer rarely translate to patients in the clinic. Ideally, one would like to characterize the

16 ACS Paragon Plus Environment

Page 17 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

biodistribution properties of therapeutic agents in both settings, thus learning about tumor targeting performance in mouse and in man. A number of companies and investigators have realized the importance of using companion diagnostics for the development of targeted therapeutics. Nuclear medicine procedures based on radiolabeled products equipped with radionuclides suitable for SPECT or PET detection appear to be the methodology of choice for the quantitative characterization of tumor targeting results. This approach is equally valuable for antibody-based therapeutics 10, 11, 15, 18 and for small molecule-drug conjugates.98, 99 Unfortunately, logistical challenges (e.g., need to work simultaneously with oncology and nuclear medicine units), regulatory hurdles (e.g., requirement to consider the radiolabeled product as a second investigational medicinal agent, with the need to provide adequate documentation and GMP preparation) and radioprotection considerations have so far limited the routine execution of these important investigations.

Nuclear medicine procedures are also ideally suited also for studying the fate of delivery vehicle (e.g., antibody or small organic ligand) and of the therapeutic payload (e.g., cytotoxic drug) in vivo. In a pioneering study, the groups of Guus van Dongen and of Matteo Zanda have recently reported the results of biodistribution studies, featuring the simultaneous labeling of the antibody moiety and of a tubulysin payload with two different radionuclides 100. We anticipate that similar studies, if performed in patients with cancer, could provide a much more detailed understanding of the pharmacodelivery process, than what is currently available. Alternatively, the fate of a targeted cytotoxic agent could be studied in vivo, if biopsies and suitable analytical methodologies are available.

17 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

It is becoming increasingly clear that delivery vehicle (e.g., antibody or small ligand), cleavable linker and payload may all contribute to the anticancer activity of conjugates in vivo. It has become common practice to speak about “linker-payload combinations”, as these chemical moieties both contribute to the success (or failure) of a given therapeutic approach. Cleavable linkers may release the toxic payload directly or indirectly (e.g., after subsequent cleavage of a self-immolative spacer) 101-103. In some cases, a traceless (i.e., linker-less) coupling of drug to the delivery vehicle is possible

104, 105

. Linkers may be cleaved by uncatalyzed hydrolytic processes

(which are not likely to differ between mouse and man)106 or by other procedures (e.g., reduction of disulfide bonds, proteolytic cleavage of peptides) which are greatly influenced by the site of in vivo delivery.56-60 While these “triggered” release processes may provide a highly desirable additional selectivity (e.g., if the trigger is more abundant at the tumor site), clinical translational activities become more challenging, as it is difficult to adequately characterize the drug release process in mouse and man.

Various antibody-drug coupling methodologies can be considered, as the antibody moiety contains various structures that can be used for chemical modification (e.g., primary amines, cysteines, oligosaccharidic moieties). Site-specific conjugation strategies are emerging as important features for the development of a successful ADC product,37, 39, 40, 49, 104, 105, 107 as the corresponding conjugates tend to be more homogenous and to display improved pharmacokinetic profiles. While site-specific drug coupling strategies were not implemented in first-generation ADC products, these features have always been incorporated in the design of small moleculedrug conjugates (SMDCs), as they are assembled by stepwise chemical synthesis processes.

18 ACS Paragon Plus Environment

Page 18 of 39

Page 19 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

One of the most puzzling aspects in the field of targeted cytotoxics relates to the choice between antibodies and small ligands as preferred vehicles for pharmacodelivery applications. On one hand, antibodies can be rapidly generated against virtually all targets of interest, including proteins of exquisite selectivity (e.g., virtually undetectable in normal adult tissues). Knowledge about the blood clearance and tissue distribution properties of antibodies is largely predictable and transferable from one antigen to the other. However, the tumor targeting properties of anticancer antibodies are still largely unexplored, both at a macroscopic and microscopic level, as discussed in a previous section. In addition, for certain antigens, the tumor targeting process may be slow and/or inefficient. By contrast small organic ligands may target tumors rapidly and efficiently. However, certain “off-target” liabilities are often observed in preclinical and clinical imaging studies with SMDCs. Virtually all small ligands characterized so far show some level of kidney uptake, which may be the consequence of a renal clearance process and precipitation (or binding) in renal structures. In addition, an undesired accumulation in normal organs may be observed, for example, when the cognate antigen is expressed at that site. PSMA ligands efficiently accumulate in salivary parotid glands, while CAIX ligands may localize to stomach (where the antigen is expressed) or lung (where other accessible carbonic anhydrase isoforms can be found). Interestingly, some of these “off-target” accumulation features in normal organs can be minimized by the pre-administration of competitors

108

or by the administration of judiciously

chosen doses. Intact monoclonal antibodies in IgG format are cleared via the hepatobiliary route. By contrast, small antibody fragments and most SMDC products studied so far are cleared via the renal route. The development of small drug conjugates, which may preferably be eliminated via the liver, remains an attractive but largely unexplored possibility.

Cancer cures and combination strategies

19 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ADCs and SMDCs are able to cure cancer in certain animal models of the disease, which do not respond to conventional cytotoxic agents. However, cancer cures in patients treated with ADC and SMDC products are rare. It is important to investigate the reasons behind the different activity observed in rodent models of cancer and in patients. As mentioned above, a quantitative characterization of the ligand-based tumor targeting process and of drug release mechanisms in mouse and man will be important. Imaging technologies will play a crucial role for a “Precision Medicine” approach to drug development.

If ADCs and SMDCs fail to cure cancer as single agents, then combination strategies may be recommended. The striking activity observed using brentuximab vedotin and nivolumab as single agents in last-line patients with Hodgkin lymphoma has prompted the clinical investigation of combination strategies. In addition, preclinical studies performed in immunocompetent tumorbearing mice have shown that the combination of ADCs with antibody-cytokine fusion proteins (immunocytokines) or the development of trifunctional antibody products, which simultaneously carry a cytokine moiety and a cytotoxic payload (immunocytokine-drug conjugates, or IDCs) may lead to very potent in vivo activity.109, 110 The antibody-based pharmacodelivery of cytokines and cytotoxic moieties is particularly attractive, due to the orthogonal toxicity profiles of these classes of payloads.

Concluding remarks The development of targeted cytotoxics represents an area of pharmaceutical research in constant evolution. The modular nature of ADC and SMDC products facilitate an “engineering” approach to drug discovery, in which a precise knowledge and a suitable modeling of the pharmacodelivery process allows to predict therapeutic performance.111,

112

In practice, in spite of substantial

20 ACS Paragon Plus Environment

Page 20 of 39

Page 21 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

advances over the last two decades, several aspects of ADC and SMDC development remain insufficiently studied and will require dedicated future investigations. A detailed characterization of the tumor targeting process and of the drug release process will continue to be important. Different types of linker-payload combinations may be required for different types of malignancies. However, an empirical choice determined by success (or failure) in exploratory clinical trials will continue to be problematic, both from an economical viewpoint and in consideration of the patients’ legitimate request for efficacious drugs. Sparing normal organs by the improvement of pharmacokinetic properties or by designing drugs with minimal toxicity to healthy tissue (as a consequence of their intrinsic mechanism of action or of an inactivation process in clearance-associated organs) will probably represent one of the most important challenges for the future. In addition, the marriage of targeted cytotoxics and of immunostimulatory products, which represent two of the most prominent classes of anticancer agents, is likely to bear fruits.

21 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

CORRESPONDING AUTHOR: Prof. Dr. Dario Neri, Professor for Biomacromolecules, ETH Zurich, Institute of Pharmaceutical Sciences, HCI G396.4, Vladimir-Prelog-Weg 1-5/10, CH-8093 Zurich; Tel: +41 44 63 37401; email: [email protected].

ABBREVIATIONS: ADC: antibody-drug conjugates; SMDC: small molecule-drug conjugates; MMAE: monomethyl auristatin-E; AML: acute myeloid leukemia; CAIX; carbonic anhydrase IX; IDC: immunocytokine-drug conjugates; PSMA: prostate specific membrane protein;

BIOGRAPHIES

Giulio Casi Giulio Casi studied chemistry at Florence University (Italy) and ETH Zurich (Switzerland). After postdoctoral research at Roche Penzberg (Germany) in 2009 he returned to Switzerland to work for Philochem AG, where he is currently responsible for the research and development of targeted cytotoxics. Dario Neri Dario Neri studied Chemistry at the Scuola Normale Superiore in Pisa (Italy), obtained his PhD in Chemistry from the ETH Zürich (Switzerland) in 1992, under the supervision of Prof. Kurt Wüthrich. After postdoctoral research at the MRC Centre in Cambridge, under the supervision of Sir Gregory Winter, he became professor at the ETH Zürich in 1996. He is a co-founder of Philogen and Philochem.

REFERENCES:

22 ACS Paragon Plus Environment

Page 22 of 39

Page 23 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

1.

Hanna, N. H.; Einhorn, L. H. Testicular cancer--discoveries and updates. New Engl. J. Med. 2014, 371, 2005-2016.

2.

Szakacs, G.; Paterson, J. K.; Ludwig, J. A.; Booth-Genthe, C.; Gottesman, M. M. Targeting multidrug resistance in cancer. Nat. Rev. Drug Discov. 2006, 5, 219-234.

3.

Bosslet, K.; Straub, R.; Blumrich, M.; Czech, J.; Gerken, M.; Sperker, B.; Kroemer, H. K.; Gesson, J. P.; Koch, M.; Monneret, C. Elucidation of the mechanism enabling tumor selective prodrug monotherapy. Cancer Res 1998, 58, 1195-1201.

4.

Krall, N.; Scheuermann, J.; Neri, D. Small targeted cytotoxics: Current state and promises from DNA-encoded chemical libraries. Angew. Chem. Int. Edit. 2013, 52, 13841402.

5.

van der Veldt, A. A. M.; Hendrikse, N. H.; Smit, E. F.; Mooijer, M. P. J.; Rijnders, A. Y.; Gerritsen, W. R.; van der Hoeven, J. J. M.; Windhorst, A. D.; Lammertsma, A. A.; Lubberink, M. Biodistribution and radiation dosimetry of C-11-labelled docetaxel in cancer patients. Eur. J. Nucl. Med. Mol. Imaging 2010, 37, 1950-1958.

6.

van der Veldt, A. A. M.; Lubberink, M.; Mathijssen, R. H. J.; Loos, W. J.; Herder, G. J. M.; Greuter, H. N.; Comans, E. F. I.; Rutten, H. B.; Eriksson, J.; Windhorst, A. D.; Hendrikse, N. H.; Postmus, P. E.; Smit, E. F.; Lammertsma, A. A. Toward prediction of efficacy of chemotherapy: A proof of concept study in lung cancer patients using C-11 docetaxel and positron emission tomography. Clin. Cancer Res. 2013, 19, 4163-4173.

7.

Chari, R. V. J.; Miller, M. L.; Widdison, W. C. Antibody-drug conjugates: An emerging concept in cancer therapy. Angew. Chem. Int. Edit. 2014, 53, 3796-3827.

8.

Senter, P. D. Potent antibody drug conjugates for cancer therapy. Curr. Opin. Chem. Biol. 2009, 13, 235-244.

9.

Gerber, H.-P.; Senter, P. D.; Grewal, I. S. Antibody drug-conjugates targeting the tumor vasculature. Current and future developments. Mabs 2009, 1, 247-253.

10.

Poli, G. L.; Bianchi, C.; Virotta, G.; Bettini, A.; Moretti, R.; Trachsel, E.; Elia, G.; Giovannoni, L.; Neri, D.; Bruno, A. Radretumab radioimmunotherapy in patients with brain metastasis: A I-124-L19SIP dosimetric PET study. Cancer Immunol. Res. 2013, 1, 134-143.

11.

Erba, P. A.; Sollini, M.; Orciuolo, E.; Traino, C.; Petrini, M.; Paganelli, G.; Bombardieri, E.; Grana, C.; Giovannoni, L.; Neri, D.; Menssen, H. D.; Mariani, G. Radioimmunotherapy with radretumab in patients with relapsed hematologic malignancies. J. Nucl. Med. 2012, 53, 922-927.

12.

Stillebroer, A. B.; Zegers, C. M. L.; Boerman, O. C.; Oosterwijk, E.; Mulders, P. F. A.; O'Donoghue, J. A.; Visser, E. P.; Oyen, W. J. G. Dosimetric analysis of Lu-177-cG250 radioimmunotherapy in renal cell carcinoma patients: Correlation with myelotoxicity and

23 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

pretherapeutic absorbed dose predictions based on In-111-cG250 imaging. J. Nucl. Med. 2012, 53, 82-89. 13.

Sharkey, R. M.; Hajjar, G.; Yeldell, D.; Brenner, A.; Burton, J.; Rubin, A.; Goldenberg, D. M. A phase I trial combining high-dose Y-90-labeled humanized anti-CEA monoclonal antibody with doxorubicin and peripheral blood stem cell rescue in advanced medullary thyroid cancer. J. Nucl. Med. 2005, 46, 620-633.

14.

Carrasquillo, J. A.; Pandit-Taskar, N.; O'Donoghue, J. A.; Humm, J. L.; Zanzonico, P.; Smith-Jones, P. M.; Divgi, C. R.; Pryma, D. A.; Ruan, S. T.; Kemeny, N. E.; Fong, Y. M.; Wong, D.; Jaggi, J. S.; Scheinberg, D. A.; Gonen, M.; Panageas, K. S.; Ritter, G.; Jungbluth, A. A.; Old, L. J.; Larson, S. M. I-124-huA33 antibody PET of colorectal cancer. J. Nucl. Med. 2011, 52, 1173-1180.

15.

Borjesson, P. K. E.; Jauw, Y. W. S.; de Bree, R.; Roos, J. C.; Castelijns, J. A.; Leemans, C. R.; van Dongen, G. A. M. S.; Boellaard, R. Radiation dosimetry of Zr-89-labeled chimeric monoclonal antibody U36 as used for immuno-PET in Head and Neck cancer patients. J. Nucl. Med. 2009, 50, 1828-1836.

16.

Muselaers, C. H. J.; Stillebroer, A. B.; Desar, I. M. E.; Boers-Sonderen, M. J.; van Herpen, C. M. L.; de Weijert, M. C. A.; Langenhuijsen, J. F.; Oosterwijk, E.; Leenders, W. P. J.; Boerman, O. C.; Mulders, P. F. A.; Oyen, W. J. G. Tyrosine kinase inhibitor sorafenib decreases In-111-Girentuximab uptake in patients with clear cell renal cell carcinoma. J. Nucl. Med. 2014, 55, 242-247.

17.

Breitz, H. B.; Tyler, A.; Bjorn, M. J.; Lesley, T.; Weiden, P. L. Clinical experience with Tc-99m nofetumomab merpentan (Verluma) radioimmunoscintigraphy. Clin Nucl Med 1997, 22, 615-620.

18.

Heuveling, D. A.; de Bree, R.; Vugts, D. J.; Huisman, M. C.; Giovannoni, L.; Hoekstra, O. S.; Leemans, C. R.; Neri, D.; van Dongen, G. A. M. S. Phase 0 microdosing PET study using the human mini antibody F16SIP in Head and Neck cancer patients. J. Nucl. Med. 2013, 54, 397-401.

19.

Ychou, M.; Pelegrin, A.; Faurous, P.; Robert, B.; Saccavini, J. C.; Guerreau, D.; Rossi, J. F.; Fabbro, M.; Buchegger, F.; Mach, J. P.; Artus, J. C. Phase-I/II radio-immunotherapy study with Iodine-131-labeled anti-CEA monoclonal antibody F6 F(ab')2 in patients with non-resectable liver metastases from colorectal cancer. Int. J. Cancer 1998, 75, 615-619.

20.

Wu, A. M.; Senter, P. D. Arming antibodies: prospects and challenges for immunoconjugates. Nat Biotechnol 2005, 23, 1137-1146.

21.

Borsi, L.; Balza, E.; Bestagno, M.; Castellani, P.; Carnemolla, B.; Biro, A.; Leprini, A.; Sepulveda, J.; Burrone, O.; Neri, D.; Zardi, L. Selective targeting of tumoral vasculature: Comparison of different formats of an antibody (L19) to the ED-B domain of fibronectin. Int. J. Cancer 2002, 102, 75-85.

24 ACS Paragon Plus Environment

Page 24 of 39

Page 25 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

22.

Berndorff, D.; Borkowski, S.; Sieger, S.; Rother, A.; Friebe, M.; Viti, F.; Hilger, C. S.; Cyr, J. E.; Dinkelborg, L. M. Radioimmunotherapy of solid tumors by targeting extra domain B fibronectin: Identification of the best-suited radioimmunoconjugate. Clin. Cancer Res. 2005, 11, 7053S-7063S.

23.

Moreau, M.; Raguin, O.; Vrigneaud, J. M.; Collin, B.; Bernhard, C.; Tizon, X.; Boschetti, F.; Duchamp, O.; Brunotte, F.; Denat, F. DOTAGA-Trastuzumab. A new antibody conjugate targeting HER2/Neu antigen for diagnostic purposes. Bioconjugate Chem. 2012, 23, 1181-1188.

24.

Brouwers, A. H.; van Eerd, J. E. M.; Frielink, C.; Oosterwijk, E.; Oyen, W. J. G.; Corstens, F. H. M.; Boerman, O. C. Optimization of radioimmunotherapy of renal cell carcinoma: Labeling of monoclonal antibody cG250 with I-131, Y-90, Lu-177, or Re-186. J. Nucl. Med. 2004, 45, 327-337.

25.

Dennis, M. S.; Jin, H.; Dugger, D.; Yang, R.; McFarland, L.; Ogasawara, A.; Williams, S.; Cole, M. J.; Ross, S.; Schwall, R. Imaging tumors with an albumin-binding Fab, a novel tumor-targeting agent. Cancer Res 2007, 67, 254-261.

26.

Saga, T.; Neumann, R. D.; Heya, T.; Sato, J.; Kinuya, S.; Le, N.; Paik, C. H.; Weinstein, J. N. Targeting cancer micrometastases with monoclonal-antibodies - A binding-site barrier. Proc. Natl. Acad. Sci. U.S.A. 1995, 92, 8999-9003.

27.

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.

28.

Rudnick, S. I.; Lou, J. L.; Shaller, C. C.; Tang, Y.; Klein-Szanto, A. J. P.; Weiner, L. M.; Marks, J. D.; Adams, G. P. Influence of affinity and antigen internalization on the uptake and penetration of Anti-HER2 antibodies in solid tumors. Cancer Res. 2011, 71, 22502259.

29.

Trail, P. A.; Willner, D.; Lasch, S. J.; Henderson, A. J.; Hofstead, S.; Casazza, A. M.; Firestone, R. A.; Hellström, I.; Hellström, K. E. Cure of xenografted human carcinomas by BR96-doxorubicin immunoconjugates. Science 1993, 261, 212-215.

30.

Lowinger, T. B. In Applications of biodegradable and biocompatible polyacetal drug conjugates to improve drug delivery, Abstr. Pap. Am. Chem. Soc., Aug 22, 2010; Amer Chemical Soc, 1155 16TH ST, NW, Washington, DC 20036 USA:

31.

de Claro, R. A.; McGinn, K.; Kwitkowski, V.; Bullock, J.; Khandelwal, A.; Habtemariam, B.; Ouyang, Y. L.; Saber, H.; Lee, K.; Koti, K.; Rothmann, M.; Shapiro, M.; Borrego, F.; Clouse, K.; Chen, X. H.; Brown, J.; Akinsanya, L.; Kane, R.; Kaminskas, E.; Farrell, A.; Pazdur, R. U.S. Food and Drug Administration approval summary: Brentuximab Vedotin for the treatment of relapsed Hodgkin Lymphoma or relapsed systemic anaplastic largecell lymphoma. Clin. Cancer Res. 2012, 18, 5845-5849.

25 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

32.

Erickson, H. K.; Park, P. U.; Widdison, W. C.; Kovtun, Y. V.; Garrett, L. M.; Hoffman, K.; Lutz, R. J.; Goldmacher, V. S.; Blättler, W. A. Antibody-maytansinoid conjugates are activated in targeted cancer cells by lysosomal degradation and linker-dependent intracellular processing. Cancer Res. 2006, 66, 4426-4433.

33.

Kovtun, Y. V.; Audette, C. A.; Mayo, M. F.; Jones, G. E.; Doherty, H.; Maloney, E. K.; Erickson, H. K.; Sun, X. X.; Wilhelm, S.; Ab, O.; Lai, K. C.; Widdison, W. C.; Kellogg, B.; Johnson, H.; Pinkas, J.; Lutz, R. J.; Singh, R.; Goldmacher, V. S.; Chari, R. V. J. Antibody-maytansinoid conjugates designed to bypass multidrug resistance. Cancer Res. 2010, 70, 2528-2537.

34.

Chudasama, V. L.; Stark, F. S.; Harrold, J. M.; Tibbitts, J.; Girish, S. R.; Gupta, M.; Frey, N.; Mager, D. E. Semi-mechanistic population pharmacokinetic model of multivalent Trastuzumab emtansine in patients with metastatic breast cancer. Clin Pharmacol Ther 2012, 92, 520-527.

35.

Peddi, P. F.; Hurvitz, S. A. Trastuzumab emtansine: the first targeted chemotherapy for treatment of breast cancer. Future Oncol 2013, 9, 319-326.

36.

Phillips, G. D. L.; Li, G.; Dugger, D. L.; Crocker, L. M.; Parsons, K. L.; Mai, E.; Blattler, W. A.; Lambert, J. M.; Chari, R. V. J.; Lutz, R. J.; Wong, W. L. T.; Jacobson, F. S.; Koeppen, H.; Schwall, R. H.; Kenkare-Mitra, S. R.; Spencer, S. D.; Sliwkowski, M. X. Targeting HER2-positive breast cancer with trastuzumab-DM1, an antibody-cytotoxic drug conjugate. Cancer Res. 2008, 68, 9280-9290.

37.

Tian, F.; Lu, Y. C.; Manibusan, A.; Sellers, A.; Tran, H.; Sun, Y.; Phuong, T.; Barnett, R.; Hehli, B.; Song, F.; DeGuzman, M. J.; Ensari, S.; Pinkstaff, J. K.; Sullivan, L. M.; Biroc, S. L.; Cho, H.; Schultz, P. G.; DiJoseph, J.; Dougher, M.; Ma, D. S.; Dushin, R.; Leal, M.; Tchistiakova, L.; Feyfant, E.; Gerber, H. P.; Sapra, P. A general approach to site-specific antibody drug conjugates. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, 17661771.

38.

Shen, B.-Q.; Xu, K.; Liu, L.; Raab, H.; Bhakta, S.; Kenrick, M.; Parsons-Reponte, K. L.; Tien, J.; Yu, S.-F.; Mai, E.; Li, D.; Tibbitts, J.; Baudys, J.; Saadi, O. M.; Scales, S. J.; McDonald, P. J.; Hass, P. E.; Eigenbrot, C.; Trung, N.; Solis, W. A.; Fuji, R. N.; Flagella, K. M.; Patel, D.; Spencer, S. D.; Khawlil, L. A.; Ebens, A.; Wong, W. L.; Vandlen, R.; Kaur, S.; Sliwkowski, M. X.; Scheller, R. H.; Polakis, P.; Junutula, J. R. Conjugation site modulates the in vivo stability and therapeutic activity of antibody-drug conjugates. Nat. Biotechnol. 2012, 30, 184-189.

39.

Junutula, J. R.; Flagella, K. M.; Graham, R. A.; Parsons, K. L.; Ha, E.; Raab, H.; Bhakta, S.; Nguyen, T.; Dugger, D. L.; Li, G. M.; Mai, E.; Phillips, G. D. L.; Hiraragi, H.; Fuji, R. N.; Tibbitts, J.; Vandlen, R.; Spencer, S. D.; Scheller, R. H.; Polakis, P.; Sliwkowski, M. X. Engineered thio-Trastuzumab-DM1 conjugate with an improved therapeutic index to target human epidermal growth factor receptor 2-positive breast cancer. Clin. Cancer Res. 2010, 16, 4769-4778.

26 ACS Paragon Plus Environment

Page 26 of 39

Page 27 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

40.

Junutula, J. R.; Raab, H.; Clark, S.; Bhakta, S.; Leipold, D. D.; Weir, S.; Chen, Y.; Simpson, M.; Tsai, S. P.; Dennis, M. S.; Lu, Y.; Meng, Y. G.; Ng, C.; Yang, J.; Lee, C. C.; Duenas, E.; Gorrell, J.; Katta, V.; Kim, A.; McDorman, K.; Flagella, K.; Venook, R.; Ross, S.; Spencer, S. D.; Wong, W. L.; Lowman, H. B.; Vandlen, R.; Sliwkowski, M. X.; Scheller, R. H.; Polakis, P.; Mallet, W. Site-specific conjugation of a cytotoxic drug to an antibody improves the therapeutic index. Nat. Biotechnol. 2008, 26, 925-932.

41.

Boghaert, E. R.; Khandke, K.; Sridharan, L.; Armellino, D.; Dougher, M.; DiJoseph, J. F.; Kunz, A.; Hamann, P. R.; Sridharan, A.; Jones, S.; Discafani, C.; Damle, N. K. Tumoricidal effect of calicheamicin immuno-conjugates using a passive targeting strategy. Int. J. Oncol. 2006, 28, 675-684.

42.

Jedema, I.; Barge, R. M. Y.; van der Velden, V. H. J.; Nijmeijer, B. A.; van Dongen, J. J. M.; Willemze, R.; Falkenburg, J. H. F. Internalization and cell cycle-dependent killing of leukemic cells by Gemtuzumab Ozogamicin: rationale for efficacy in CD33-negative malignancies with endocytic capacity. Leukemia 2004, 18, 316-325.

43.

Farhat, H.; Reman, O.; Raffoux, E.; Berthon, C.; Pautas, C.; Kammoun, L.; Chantepie, S.; Gardin, C.; Rousselot, P.; Chevret, S.; Dombret, H.; Castaigne, S. Fractionated doses of gemtuzumab ozogamicin with escalated doses of daunorubicin and cytarabine as first acute myeloid leukemia salvage in patients aged 50-70-year old: A phase 1/2 study of the acute leukemia French association. Am J Hematol 2012, 87, 62-65.

44.

Taksin, A. L.; Legrand, O.; Raffoux, E.; de Revel, T.; Thomas, X.; Contentin, N.; Bouabdallah, R.; Pautas, C.; Turlure, P.; Reman, O.; Gardin, C.; Varet, B.; de Botton, S.; Pousset, F.; Farhat, H.; Chevret, S.; Dombret, H.; Castaigne, S. High efficacy and safety profile of fractionated doses of Mylotarg as induction therapy in patients with relapsed acute myeloblastic leukemia: a prospective study of the alfa group. Leukemia 2007, 21, 66-71.

45.

Burnett, A. K.; Hills, R. K.; Milligan, D.; Kjeldsen, L.; Kell, J.; Russell, N. H.; Yin, J. A. L.; Hunter, A.; Goldstone, A. H.; Wheatley, K. Identification of patients with acute myeloblastic leukemia who benefit from the addition of Gemtuzumab Ozogamicin: Results of the MRC AML15 trial. J. Clin. Oncol. 2011, 29, 369-377.

46.

Burnett, A. K.; Russell, N. H.; Hills, R. K.; Kell, J.; Freeman, S.; Kjeldsen, L.; Hunter, A. E.; Yin, J.; Craddock, C. F.; Dufva, I. H.; Wheatley, K.; Milligan, D. Addition of Gemtuzumab Ozogamicin to induction chemotherapy improves survival in older patients with acute myeloid leukemia. J. Clin. Oncol. 2012, 30, 3924-3931.

47.

Castaigne, S.; Pautas, C.; Terre, C.; Raffoux, E.; Bordessoule, D.; Bastie, J. N.; Legrand, O.; Thomas, X.; Turlure, P.; Reman, O.; de Revel, T.; Gastaud, L.; de Gunzburg, N.; Contentin, N.; Henry, E.; Marolleau, J. P.; Aljijakli, A.; Rousselot, P.; Fenaux, P.; Preudhomme, C.; Chevret, S.; Dombret, H.; Association, A. L. F. Effect of gemtuzumab ozogamicin on survival of adult patients with de-novo acute myeloid leukaemia (ALFA0701): a randomised, open-label, phase 3 study. Lancet 2012, 379, 1508-1516.

27 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

48.

Delaunay, J.; Recher, C.; Pigneux, A.; Witz, F.; Vey, N.; Blanchet, O.; Lefebvre, P.; Luquet, I.; Guillerme, I.; Volteau, C.; Gyan, E.; Lioure, B.; Jourdan, E.; Bouscary, D.; Guieze, R.; Randriamalala, E.; Uribe, M. E. O.; Dreyfus, F.; Lacombe, C.; Bene, M. C.; Cahn, J. Y.; Harousseau, J. L.; Ifrah, N. Addition of Gemtuzumab Ozogamycin to chemotherapy improves event-free survival but not overall survival of AML patients with intermediate cytogenetics not eligible for allogeneic transplantation. Results of the GOELAMS AML 2006 IR study. Blood 2011, 118, 37-38.

49.

Sutherland, M. S. K.; Walter, R. B.; Jeffrey, S. C.; Burke, P. J.; Yu, C. P.; Kostner, H.; Stone, I.; Ryan, M. C.; Sussman, D.; Lyon, R. P.; Zeng, W. P.; Harrington, K. H.; Klussman, K.; Westendorf, L.; Meyer, D.; Bernstein, I. D.; Senter, P. D.; Benjamin, D. R.; Drachman, J. G.; McEarchern, J. A. SGN-CD33A: a novel CD33-targeting antibodydrug conjugate using a pyrrolobenzodiazepine dimer is active in models of drug-resistant AML. Blood 2013, 122, 1455-1463.

50.

Bander, N. H. Antibody–drug conjugate target selection: critical factors. In Methods in Molecular Biology: Antibody-drug conjugates, Ducry, L., Ed. Humana Press: 2013; Vol. 1045, pp 29-40.

51.

Tarli, L.; Balza, E.; Viti, F.; Borsi, L.; Castellani, P.; Berndorff, D.; Dinkelborg, L.; Neri, D.; Zardi, L. A high-affinity human antibody that targets tumoral blood vessels. Blood 1999, 94, 192-198.

52.

Viti, F.; Tarli, L.; Giovannoni, L.; Zardi, L.; Neri, D. Increased binding affinity and valence of recombinant antibody fragments lead to improved targeting of tumoral angiogenesis. Cancer Res. 1999, 59, 347-352.

53.

Pini, A.; Viti, F.; Santucci, A.; Carnemolla, B.; Zardi, L.; Neri, P.; Neri, D. Design and use of a phage display library - Human antibodies with subnanomolar affinity against a marker of angiogenesis eluted from a two-dimensional gel. J. Biol. Chem. 1998, 273, 21769-21776.

54.

Villa, A.; Trachsel, E.; Kaspar, M.; Schliemann, C.; Sommavilla, R.; Rybak, J.-N.; Rösli, C.; Borsi, L.; Neri, D. A high-affinity human monoclonal antibody specific to the alternatively spliced EDA domain of fibronectin efficiently targets tumor neo-vasculature in vivo. Int. J. Cancer 2008, 122, 2405-2413.

55.

Neri, D.; Bicknell, R. Tumour vascular targeting. Nat. Rev. Cancer 2005, 5, 436-446.

56.

Perrino, E.; Steiner, M.; Krall, N.; Bernardes, G. J. L.; Pretto, F.; Casi, G.; Neri, D. Curative properties of noninternalizing antibody-drug conjugates based on maytansinoids. Cancer Res. 2014, 74, 2569-2578.

57.

Bernardes, G. J. L.; Casi, G.; Hartmann, I.; Trüssel, S.; Schwager, K.; Scheuermann, J.; Neri, D. A traceless vascular-targeting antibody-drug conjugate for cancer therapy. Angew. Chem. Int. Ed. 2012, 51, 941-944.

28 ACS Paragon Plus Environment

Page 28 of 39

Page 29 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

58.

Krall, N.; Pretto, F.; Decurtins, W.; Bernardes, G. J. L.; Supuran, C. T.; Neri, D. A smallmolecule drug conjugate for the treatment of carbonic anhydrase IX expressing tumors. Angew. Chem.-Int. Edit. 2014, 53, 4231-4235.

59.

Rothberg, J. M.; Bailey, K. M.; Wojtkowiak, J. W.; Ben-Nun, Y.; Bogyo, M.; Weber, E.; Moin, K.; Blum, G.; Mattingly, R. R.; Gillies, R. J.; Sloane, B. F. Acid-mediated tumor proteolysis: Contribution of cysteine cathepsins. Neoplasia 2013, 15, 1111-1123.

60.

Fuhrmann, K.; Gauthier, M. A.; Leroux, J. C. Targeting of injectable drug nanocrystals. Mol. Pharm. 2014, 11, 1762-1771.

61.

Winter, G.; Griffiths, A. D.; Hawkins, R. E.; Hoogenboom, H. R. Making antibodies by phage display technology. Annu. Rev. Immunol. 1994, 12, 433-455.

62.

Srinivasarao, M.; Galliford, C. V.; Low, P. S. Principles in the design of ligand-targeted cancer therapeutics and imaging agents. Nat Rev Drug Discov 2015, 14, 203-219.

63.

Low, P. S.; Henne, W. A.; Doorneweerd, D. D. Discovery and development of folic-acidbased receptor targeting for imaging and therapy of cancer and inflammatory diseases. Acc. Chem. Res. 2008, 41, 120-129.

64.

Hillier, S. M.; Maresca, K. P.; Femia, F. J.; Marquis, J. C.; Foss, C. A.; Nguyen, N.; Zimmerman, C. N.; Barrett, J. A.; Eckelman, W. C.; Pomper, M. G.; Joyal, J. L.; Babich, J. W. Preclinical evaluation of novel glutamate-urea-lysine analogues that target prostatespecific membrane antigen as molecular imaging pharmaceuticals for prostate cancer. Cancer Res. 2009, 69, 6932-6940.

65.

Vallabhajosula, S.; Nikolopoulou, A.; Babich, J. W.; Osborne, J. R.; Tagawa, S. T.; Lipai, I.; Solnes, L.; Maresca, K. P.; Armor, T.; Joyal, J. L.; Crummet, R.; Stubbs, J. B.; Goldsmith, S. J. Tc-99m-labeled small-molecule inhibitors of prostate-specific membrane antigen: Pharmacokinetics and biodistribution studies in healthy subjects and patients with metastatic prostate cancer. J. Nucl. Med. 2014, 55, 1791-1798.

66.

Hillier, S. M.; Maresca, K. P.; Lu, G.; Merkin, R. D.; Marquis, J. C.; Zimmerman, C. N.; Eckelman, W. C.; Joyal, J. L.; Babich, J. W. Tc-99m-labeled small-molecule inhibitors of prostate-specific membrane antigen for molecular imaging of prostate cancer. J. Nucl. Med. 2013, 54, 1369-1376.

67.

Ginj, M.; Zhang, H.; Waser, B.; Cescato, R.; Wild, D.; Wang, X.; Erchegyi, J.; Rivier, J.; Maecke, H. R.; Reubi, J. C. Radiolabeled somatostatin receptor antagonists are preferable to agonists for in vivo peptide receptor targeting of tumors. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 16436-16441.

68.

Neri, D.; Supuran, C. T. Interfering with pH regulation in tumours as a therapeutic strategy. Nat. Rev. Drug Discov. 2011, 10, 767-777.

29 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

69.

Krall, N.; Pretto, F.; Neri, D. A bivalent small molecule-drug conjugate directed against carbonic anhydrase IX can elicit complete tumour regression in mice. Chem. Sci. 2014, 5, 3640-3644.

70.

Wichert, M.; Krall, N.; Decurtins, W.; Franzini, R. M.; Pretto, F.; Schneider, P.; Neri, D.; Scheuermann, J. Dual-display of small molecules enables the discovery of ligand pairs and facilitates affinity maturation. Nat Chem 2015, 7, 241-249.

71.

Doss, M.; Kolb, H. C.; Walsh, J. C.; Mocharla, V. P.; Zhu, Z. H.; Haka, M.; Alpaugh, R. K.; Chen, D. Y. T.; Yu, J. Q. Biodistribution and radiation dosimetry of the carbonic anhydrase IX imaging agent [(18) F]VM4-037 determined from PET/CT scans in healthy volunteers. Mol Imaging Biol 2014, 16, 739-746.

72.

Lepenies, B.; Lee, J.; Sonkaria, S. Targeting C-type lectin receptors with multivalent carbohydrate ligands. Adv Drug Deliv Rev 2013, 65, 1271-1281.

73.

Chittasupho, C. Multivalent ligand: design principle for targeted therapeutic delivery approach. Ther. Deliv. 2012, 3, 1171-1187.

74.

Guillemard, V.; Nedev, H. N.; Berezov, A.; Murali, R.; Saragovi, H. U. HER2-Mediated internalization of a targeted prodrug cytotoxic conjugate is dependent on the valency of the targeting ligand. DNA Cell Biol 2005, 24, 350-358.

75.

Daguer, J. P.; Zambaldo, C.; Ciobanu, M.; Morieux, P.; Barluenga, S.; Winssinger, N. DNA display of fragment pairs as a tool for the discovery of novel biologically active small molecules. Chem. Sci. 2015, 6, 739-744.

76.

Franzini, R. M.; Neri, D.; Scheuermann, J. DNA-encoded chemical libraries: Advancing beyond conventional small-molecule libraries. Acc. Chem. Res. 2014, 47, 1247-1255.

77.

Chan, A. I.; McGregor, L. M.; Liu, D. R. Novel selection methods for DNA-encoded chemical libraries. Curr Opin Chem Biol 2015, 26C, 55-61.

78.

Morioka, T.; Loik, N. D.; Hipolito, C. J.; Goto, Y.; Suga, H. Selection-based discovery of macrocyclic peptides for the next generation therapeutics. Curr Opin Chem Biol 2015, 26C, 34-41.

79.

Heinis, C.; Rutherford, T.; Freund, S.; Winter, G. Phage-encoded combinatorial chemical libraries based on bicyclic peptides. Nat. Chem. Biol. 2009, 5, 502-507.

80.

Parker, N.; Turk, M. J.; Westrick, E.; Lewis, J. D.; Low, P. S.; Leamon, C. P. Folate receptor expression in carcinomas and normal tissues determined by a quantitative radioligand binding assay. Anal. Biochem. 2005, 338, 284-293.

81.

Morris, R. T.; Joyrich, R. N.; Naumann, R. W.; Shah, N. P.; Maurer, A. H.; Strauss, H. W.; Uszler, J. M.; Symanowski, J. T.; Ellis, P. R.; Harb, W. A. Phase II study of treatment of advanced ovarian cancer with folate-receptor-targeted therapeutic

30 ACS Paragon Plus Environment

Page 30 of 39

Page 31 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(vintafolide) and companion SPECT-based imaging agent (Tc-99m-etarfolatide). Ann Oncol 2014, 25, 852-858. 82.

Naumann, R. W.; Coleman, R. L.; Burger, R. A.; Sausville, E. A.; Kutarska, E.; Ghamande, S. A.; Gabrail, N. Y.; Depasquale, S. E.; Nowara, E.; Gilbert, L.; Gersh, R. H.; Teneriello, M. G.; Harb, W. A.; Konstantinopoulos, P. A.; Penson, R. T.; Symanowski, J. T.; Lovejoy, C. D.; Leamon, C. P.; Morgenstern, D. E.; Messmann, R. A. PRECEDENT: a randomized phase II trial comparing vintafolide (EC145) and pegylated liposomal doxorubicin (PLD) in combination versus PLD alone in patients with platinum-resistant ovarian cancer. J Clin Oncol 2013, 31, 4400-4406.

83.

LoRusso, P. M.; Edelman, M. J.; Bever, S. L.; Forman, K. M.; Pilat, M.; Quinn, M. F.; Li, J.; Heath, E. I.; Malburg, L. M.; Klein, P. J.; Leamon, C. P.; Messmann, R. A.; Sausville, E. A. Phase I study of folate conjugate EC145 (Vintafolide) in patients with refractory solid tumors. J. Clin. Oncol. 2012, 30, 4011-4016.

84.

Uhlen, M.; Fagerberg, L.; Hallstrom, B. M.; Lindskog, C.; Oksvold, P.; Mardinoglu, A.; Sivertsson, A.; Kampf, C.; Sjostedt, E.; Asplund, A.; Olsson, I.; Edlund, K.; Lundberg, E.; Navani, S.; Szigyarto, C. A.; Odeberg, J.; Djureinovic, D.; Takanen, J. O.; Hober, S.; Alm, T.; Edqvist, P. H.; Berling, H.; Tegel, H.; Mulder, J.; Rockberg, J.; Nilsson, P.; Schwenk, J. M.; Hamsten, M.; von Feilitzen, K.; Forsberg, M.; Persson, L.; Johansson, F.; Zwahlen, M.; von Heijne, G.; Nielsen, J.; Ponten, F. Proteomics. Tissue-based map of the human proteome. Science 2015, 347, 394.

85.

Uhlen, M.; Oksvold, P.; Fagerberg, L.; Lundberg, E.; Jonasson, K.; Forsberg, M.; Zwahlen, M.; Kampf, C.; Wester, K.; Hober, S.; Wernerus, H.; Bjorling, L.; Ponten, F. Towards a knowledge-based Human Protein Atlas. Nat. Biotechnol. 2010, 28, 1248-1250.

86.

Zechmann, C. M.; Afshar-Oromieh, A.; Armor, T.; Stubbs, J. B.; Mier, W.; Hadaschik, B.; Joyal, J.; Kopka, K.; Debus, J.; Babich, J. W.; Haberkorn, U. Radiation dosimetry and first therapy results with a I-124/I-131-labeled small molecule (MIP-1095) targeting PSMA for prostate cancer therapy. Eur. J. Nucl. Med. Mol. Imaging 2014, 41, 1280-1292.

87.

Silver, D. A.; Pellicer, I.; Fair, W. R.; Heston, W. D. W.; CordonCardo, C. Prostatespecific membrane antigen expression in normal and malignant human tissues. Clin. Cancer Res. 1997, 3, 81-85.

88.

Israeli, R. S.; Powell, C. T.; Fair, W. R.; Heston, W. D. W. Molecular-cloning of a complementary-DNA encoding a prostate-specific membrane antigen. Cancer Res. 1993, 53, 227-230.

89.

Chang, S. S.; Reuter, V. E.; Heston, W. D. W.; Bander, N. H.; Grauer, L. S.; Gaudin, P. B. Five different anti-prostate-specific membrane antigen (PSMA) antibodies confirm PSMA expression in tumor-associated neovasculature. Cancer Res. 1999, 59, 3192-3198.

90.

Liu, H.; Moy, P.; Kim, S.; Xia, Y.; Rajasekaran, A.; Navarro, V.; Knudsen, B.; Bander, N. H. Monoclonal antibodies to the extracellular domain of prostate-specific membrane antigen also react with tumor vascular endothelium. Cancer Res. 1997, 57, 3629-3634.

31 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

91.

De Simone, G.; Alterio, V.; Supuran, C. T. Exploiting the hydrophobic and hydrophilic binding sites for designing carbonic anhydrase inhibitors. Expert Opin Drug Dis 2013, 8, 793-810.

92.

Thiry, A.; Dogne, J. M.; Masereel, B.; Supuran, C. T. Targeting tumor-associated carbonic anhydrase IX in cancer therapy. Trends Pharmacol Sci 2006, 27, 566-573.

93.

Bui, M. H.; Seligson, D.; Han, K. R.; Pantuck, A. J.; Dorey, F. J.; Huang, Y.; Horvath, S.; Leibovich, B. C.; Chopra, S.; Liao, S. Y.; Stanbridge, E.; Lerman, M. I.; Palotie, A.; Figlin, R. A.; Belldegrun, A. S. Carbonic anhydrase IX is an independent predictor of survival in advanced renal clear cell carcinoma: implications for prognosis and therapy. Clin. Cancer Res. 2003, 9, 802-811.

94.

Svastova, E.; Hulikova, A.; Rafajova, M.; Zat'ovicova, M.; Gibadulinova, A.; Casini, A.; Cecchi, A.; Scozzafava, A.; Supuran, C. T.; Pastorek, J.; Pastorekova, S. Hypoxia activates the capacity of tumor-associated carbonic anhydrase IX to acidify extracellular pH. FEBS Lett. 2004, 577, 439-445.

95.

Scheurer, S. B.; Rybak, J. N.; Rosli, C.; Neri, D.; Elia, G. Modulation of gene expression by hypoxia in human umbilical cord vein endothelial cells: A transcriptomic and proteomic study. Proteomics 2004, 4, 1737-1760.

96.

Petrul, H. M.; Schatz, C. A.; Kopitz, C. C.; Adnane, L.; McCabe, T. J.; Trail, P.; Ha, S.; Chang, Y. S.; Voznesensky, A.; Ranges, G.; Tamburini, P. P. Therapeutic mechanism and efficacy of the antibody-drug conjugate BAY 79-4620 targeting human carbonic anhydrase 9. Mol. Cancer. Ther. 2012, 11, 340-349.

97.

Dubois, L.; Lieuwes, N. G.; Maresca, A.; Thiry, A.; Supuran, C. T.; Scozzafava, A.; Wouters, B. G.; Lambin, P. Imaging of CA IX with fluorescent labelled sulfonamides distinguishes hypoxic and (re)-oxygenated cells in a xenograft tumour model. Radiother. Oncol. 2009, 92, 423-428.

98.

Maurer, A. H.; Elsinga, P.; Fanti, S.; Nguyen, B.; Oyen, W. J. G.; Weber, W. A. Imaging the folate receptor on cancer cells with Tc-99m-Etarfolatide: Properties, clinical use, and future potential of folate receptor imaging. J. Nucl. Med. 2014, 55, 701-704.

99.

Kularatne, S. A.; Wang, K.; Santhapuram, H. K. R.; Low, P. S. Prostate-specific membrane antigen targeted imaging and therapy of prostate cancer using a PSMA inhibitor as a homing ligand. Mol. Pharm. 2009, 6, 780-789.

100.

Cohen, R.; Vugts, D. J.; Visser, G. W. M.; Stigter-Van Walsum, M.; Bolijn, M.; Spiga, M.; Lazzari, P.; Shankar, E.; Sani, M.; Zanda, M.; Van Dongen, G. A. M. S. Development of novel ADCs: Conjugation of tubulysin analogues to trastuzumab monitored by dual radiolabeling. Cancer Res. 2014, 74, 5700-5710.

32 ACS Paragon Plus Environment

Page 32 of 39

Page 33 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

101.

Erez, R.; Segal, E.; Miller, K.; Satchi-Fainaro, R.; Shabat, D. Enhanced cytotoxicity of a polymer-drug conjugate with triple payload of paclitaxel. Biorg. Med. Chem. 2009, 17, 4327-4335.

102.

Ojima, I. Guided molecular missiles for tumor-targeting chemotherapy-case studies using the second-generation taxolds as warheads. Acc. Chem. Res. 2008, 41, 108-119.

103.

Weinstain, R.; Segal, E.; Satchi-Fainaro, R.; Shabat, D. Real-time monitoring of drug release. Chem. Commun. 2010, 46, 553-555.

104.

Bernardes, G. J. L.; Steiner, M.; Hartmann, I.; Neri, D.; Casi, G. Site-specific chemical modification of antibody fragments using traceless cleavable linkers. Nat. Protoc. 2013, 8, 2079-2089.

105.

Casi, G.; Huguenin-Dezot, N.; Zuberbühler, K.; Scheuermann, J.; Neri, D. Site-specific traceless coupling of potent cytotoxic drugs to recombinant antibodies for pharmacodelivery. J. Am. Chem. Soc. 2012, 134, 5887-5892.

106.

Santi, D. V.; Schneider, E. L.; Reid, R.; Robinson, L.; Ashley, G. W. Predictable and tunable half-life extension of therapeutic agents by controlled chemical release from macromolecular conjugates. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 6211-6216.

107.

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.; Zhou, D.; Chandrasekaran, R.; Samain, C.; Loo, C.; Rossi, A.; Rickert, M.; Krimm, S.; Wong, T.; Chin, S. M.; Yu, J.; Dilley, J.; Chaparro-Riggers, J.; Filzen, G. F.; O'Donnell, C. J.; Wang, F.; Myers, J. S.; Pons, J.; Shelton, D. L.; Rajpal, A. Location matters: site of conjugation modulates stability and pharmacokinetics of antibody drug conjugates. Chem Biol 2013, 20, 161-167.

108.

Muller, C.; Schibli, R. Prospects in folate receptor-targeted radionuclide therapy. Front Oncol 2013, 3, 249.

109.

Gutbrodt, K. L.; Casi, G.; Neri, D. Antibody-based delivery of IL2 and cytotoxics eradicates tumors in immunocompetent mice. Mol. Cancer. Ther. 2014, 13, 1772-1776.

110.

List, T.; Casi, G.; Neri, D. A chemically defined trifunctional antibody–cytokine–drug conjugate with potent antitumor activity. Mol. Cancer. Ther. 2014, 13, 2641-2652.

111.

Wittrup, K. D.; Thurber, G. M.; Schmidt, M. M.; Rhoden, J. J. Practical theoretic guidance for the design of tumor-targeting agents. Method Enzymol 2012, 503, 255-268.

112.

Tzeng, A.; Kwan, B. H.; Opel, C. F.; Navaratna, T.; Wittrup, K. D. Antigen specificity can be irrelevant to immunocytokine efficacy and biodistribution. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, 3320-3325.

33 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 39

TABLE OF CONTENTS GRAPHIC

DRUG

DRUG

34 ACS Paragon Plus Environment

Page 35 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45

Journal of Medicinal Chemistry

ACS Paragon Plus Environment

Journal of Medicinal Chemistry

a) 1

b)

H

H

L

SIP / Minibody VH

VL

VL

CH4 CH4

V

H

IgG antibody V

VH

VL CL

H

1 CH

C

H

1

VL

2 3scFv 4 5V V 6 7 8 Diabody 9 10 V V 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

C

L

CH2

CH2 Fc

L

interdomain disulfide bond

CH3

CH3

ACS Paragon Plus Environment

Page 36 of 39

c)

Page 37 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Journal of Medicinal Chemistry

O O

H N

N H

O

O

O N H

O

NH

N

O

O

N

H N O

O N

H N

N OMe O

OH

OMe O

MeO

N Cl

N

O

O

O

O

S O

NH 2

O O

S

N H MeO OH

ACS Paragon Plus Environment

O

O

N

HN O

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

FOLATE RECEPTOR

PSMA O

O N

HN H 2N

N

O

OH

N H

N H

N

Page 38 of 39

OH

O

O O

HO

N H

O

SOMATOSTATIN RECEPTOR

OH

N H

O

CAIX

O

OH

N H

N H

S O S

HO NH

O O

O

O

H N

O HN

N N N

HN HO

H N O

N

HN

O

O

NH

N S

SO2NH 2

1

CH3 O

HN

N

O N N N

NH 2 HN

OH

CO2H O HN

N NH

N S

SO2NH 2

2

CO2H O HN O

ACS Paragon Plus Environment

OH

Page 39 of 39

Untargeted dye mouse #1

a

1h

2h

6h

8h

11 h

23 h

1.0 0.8 0.6 0.4

x 10 9

0.2 Tumour

b

4h

1h

Kidneys 2h

4h

6h

8h

11 h

23 h

0.0 5.0

Epi-fluorescence [radiant efficiency] = p/s/cm2 /sr µW/cm 2

Colour scale Min = 0.00 Max = 1.00 x 10

9

Colour scale Min = 0.00 Max = 5.00 x 10

9

Colour scale Min = 0.00 Max = 5.00 x 10

9

Targeted 1 mouse #1

4.0 3.0 2.0

c

x 10 9

1.0 Tumour 1h

2h

4h

6h

8h

11 h

23 h

0.0 5.0 4.0

Targeted 2 mouse #1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

3.0 2.0 1.0 Tumour

0.0

ACS Paragon Plus Environment

x 10 9