Journal Pre-proof The past, present and metalloproteinase inhibitors
future
perspectives
of
matrix
Kang Li, Franklin R. Tay, Cynthia Kar Yung Yiu PII:
S0163-7258(19)30217-7
DOI:
https://doi.org/10.1016/j.pharmthera.2019.107465
Reference:
JPT 107465
To appear in:
Pharmacology and Therapeutics
Please cite this article as: K. Li, F.R. Tay and C.K.Y. Yiu, The past, present and future perspectives of matrix metalloproteinase inhibitors, Pharmacology and Therapeutics(2019), https://doi.org/10.1016/j.pharmthera.2019.107465
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2019 Published by Elsevier.
Journal Pre-proof
P&T 23637 The Past, Present and Future Perspectives of Matrix Metalloproteinase Inhibitors Article Type: Review Article
Kang Li1 , Franklin R. Tay2 *, Cynthia Kar Yung Yiu1 *
f
1. Paediatric Dentistry and Orthodontics, Faculty of Dentistry, The University of
oo
Hong Kong, Prince Philip Dental Hospital, 34 Hospital Road, Sai Ying Pun, Hong
pr
Kong
Pr
e-
2. College of Graduate Studies, Augusta University, Augusta, GA, USA
Email:
[email protected];
[email protected] (*co-corresponding authors)
al
Tel: +852 28590256
Declaration:
Jo u
rn
Fax: +852 25593803
The article titled “The Past, Present and Future Perspectives of Matrix Metalloproteinase Inhibitors” has not been published and is not under consideration for publication elsewhere.
Journal Pre-proof
Abstract: Matrix metalloproteinases (MMPs) are a large family of enzymes that degrade the extracellular matrix (ECM). Under pathologic conditions, overexpression of MMPs or insufficient control by tissue inhibitors of MMPs (TIMPs) results in the dysregulation of tissue remodeling and causes a variety of diseases such as encephalomyelitis, rheumatoid arthritis, Alzheimer’s disease and tumors. Therefore,
f
the high affinity of MMPs for biomolecules renders them attractive targets for
oo
inhibition when homeostasis breaks down in the ECM. There are 4 generations of
pr
MMP inhibitors (MMPIs), ranging from small molecules or peptides to antibodies and
e-
protein-engineered inhibitors of metalloproteinase. Although a plethora of MMPIs has
Pr
been synthesized, most of them have failed in clinical trials or are still in the laboratory stage of development. The present review summarizes the development of
al
MMPIs, their associated problems and discusses future directions for the development
Jo u
rn
of the future generations of MMPIs.
Key Words: Matrix metalloproteinase; MMP; Inhibitor; MMPI
Journal Pre-proof Table of Contents: 1. Introduction 2. Historical generations of MMPIs 2.1 First generation – hydroxamate-based inhibitors 2.2 Second generation – non-hydroxamate-based inhibitors 2.3 Third generation – catalytic domain (non-zinc binding) inhibitors 2.4 Fourth generation -- allosteric and exosite inhibitors 3. On-going research techniques
3.2 Protein-engineered inhibitors
e-
4.1 Enhancing binding site specificity
pr
4. Future prospects of next generation of MMPIs
oo
f
3.1 Antibody-based inhibitors
4.2 Smart drug delivery systems for achieving selectivity
5. Concluding remarks
rn
References
al
Conflict of interest statement
Pr
4.3 Thorough understanding of MMPs in health and disease
Jo u
Abbreviations: ADAM, a disintegrin and metalloproteinase; ADAMTS, ADAM with thrombospondin motifs; bid, twice daily; COPD, chronic obstructive pulmonary disease; ECM, extracellular matrix; FDA, Food and Drug Administration; GPI, glycophosphatidylinositol; IV, intravenous injection; mAb, monoclonal antibody; MMP, Matrix metalloproteinase; MMPI, Matrix metalloproteinases inhibitor; MSS, musculoskeletal syndrome; MT, membrane-type; NSCLC, non-small-cell lung cancer; N-TIMP, N-terminal peptide of TIMP; PDB, Protein Data Bank; qd, once daily; SC, subcutaneous administration; SCLC, small-cell lung cancer; TDP, time to disease progression; tid, three times daily; TIMP, tissue inhibitors of MMP.
Journal Pre-proof
1. Introduction Matrix metalloproteinases (MMPs) were first discovered in 1962 by Gross and Lapiere in the tail of tadpole during frog metamorphosis (Gross and Lapiere, 1962). These enzymes have been recognized as the major proteolytic enzymes for regulating extracellular matrix (ECM) degradation (Gross, 2004). To date, 23 MMPs have been
f
identified in humans. Apart from their tissue-remodeling function, MMPs also
oo
participate in regulating many non- matrix targets, such as cell surface receptors,
e-
proteinases (Vanlaere and Libert, 2009).
pr
cell-cell adhesion molecules, cytokines, clotting factors, chemokines and other
Pr
Matrix metalloproteinases may be classified in two different ways, based on substrate or domain organization. The substrate classification is more common and
al
categories MMPs into collagenases (MMP-1, MMP-8, and MMP-13), gelatinases
rn
(MMP-2 and MMP-9), stromelysins (MMP-3, MMP-10, and MMP-11), matrilysins
Jo u
(MMP-7 and MMP-26), membrane-type (MT) MMPs (MMP-14, MMP-15, MMP-16, MMP-17, MMP-24, and MMP-25) and others (MMP-12, MMP-19, MMP-20, MMP-21, MMP-23, MMP-27, and MMP-28) (Visse and Nagase, 2003). Many studies have pointed out the limitation of this classification. For example, MMP-1, designated as a collagenase, also cleaves tenascin and aggrecan; MMP-2, a gelatinase, also degrades fibronectin, aggrecan and non-ECM substrates. Even MMP-14, a well-known MT1-MMP, serves as a collagenase (Chakraborti et al., 2003; Tam et al., 2004). The easily misleading names and the unclassified MMPs have limited the accuracy of substrate-based classification.
Journal Pre-proof
An alternative classification of MMPs has been proposed based on their domain organization. In this classification, MMPs may be divided into archetypal MMPs, matrilysins, gelatinases and furin-activable MMPs (Fanjul-Fernández et al., 2010; Vandenbroucke and Libert, 2014) (Fig. 1). All MMPs share some common domain structures (except MMP-7, MMP-23, and MMP-26). The signal peptide mediates
This intramolecular cysteine-switch keeps MMPs in their
oo
with a cysteine residue.
f
MMP secretion. The pro-peptide protects the Zn2+ binding site through interaction
pr
latent, inactive pro- forms until cleavage of the pro-peptide domain. The catalytic
e-
domain, which contains Zn2+, is responsible for substrate hydrolysis. The hemopexin
Jo u
rn
al
Pr
domain, which is linked to catalytic domain via a flexible hinge, is responsible for
Journal Pre-proof
substrate recognition and dimerization (Aureli et al., 2008; Tallant et al., 2010). Fig.1 Structural classification of MMPs based on their domain arrangement and the relationship with other metzincins superfamily members
Apart from the commonly-shared structures, there are features which are characteristic of individual MMPs. These characteristics include: (1) MMP-7 and MMP-26 lack the hinge region and the hemopexin domain; (2) MMP-2 and MMP-9
oo
f
contain three fibronectin type II motifs (aka. collagen binding domain) on the catalytic site; (3) furin-activable MMPs have a furin cleavage motif between the
pr
pro-peptide and catalytic domain; (4) MT-MMPs have a transmembrane domain or
e-
glycophosphatidylinositol (GPI) anchor on their C-terminus; and (5) MMP-23 lacks
Pr
the signal peptide, the cysteine-switch motif and the hemopexin domain. The structure of ECM is not static. Under normal conditions, MMPs and
al
endogenous tissue inhibitors of MMPs (TIMPs) mutually participate in the
rn
homeostasis of the ECM (Overall and López-Otín, 2002; Woolley et al., 1975). The
Jo u
TIMP family consists of four members, namely TIMP1-4, which were first identified in serum and tissue culture (Bauer et al., 1975; Eisen et al., 1971). Each TIMP forms a 1:1 complex with MMP in a non-selective manner (Maskos and Bode, 2003). The three-dimensional structures
of
TIMP-MMP
complexes
stipulate
that
the
wedge-shaped ridge of the N-terminal peptide of TIMP (N-TIMP, consisting of approximately 126 amino acids) interacts with the Zn2+ of MMP active site through the α-amino and carbonyl groups. In addition, the serine/threonine group of N-TIMP interacts with the nucleophilic glutamine in the MMP catalytic site and activates the hydrolysis process (Mohan et al., 2016). Under pathologic conditions, overexpression
Journal Pre-proof
of MMPs or insufficient control by TIMPs results in the dysregulation of tissue remodeling and causes a variety of diseases, such as encephalomyelitis (Mohan et al., 2016), rheumatoid arthritis (Ahrens et al., 1996), Alzheimer’s disease (Peress et al., 1995) and tumors (Coussens et al., 2002). Hence, MMPs are often considered valuable drug targets. A plethora of MMP inhibitors (MMPIs) have been designed for
f
the purpose of restoring tissue homeostasis and curing diseases.
oo
As early as 1988, Reich et al. used a hydroxamic acid compound (SC-44463) to
pr
block collagenase and prevent metastasis in mouse models, which initiated the era of
e-
MMP inhibition therapeutics. Following that seminal study, clinical trials commenced
Pr
with great enthusiasm and expectations on the first generation MMPIs (such as Batimastat, Marimastat, MMI-270 and Prinomastat) (Table 1). However, after nearly
al
30 years of trials and tribulations, only one drug (Periostat®, doxycycline hydrate) had
rn
obtained approval from the US Food and Drug Administration (FDA) for the
Jo u
treatment of periodontal disease (Golub et al., 2001; Preshaw et al., 2004; Wynn, 1999). Accordingly, the objective of the present review is to provide a summary of the historical generations of MMPIs as well as the lessons acquired from those developments. This is followed by reviewing the latest research directions on MMPI synthesis, and finally, perspectives on the development of future generations of MMPIs.
2. Historical generations of MMPIs 2.1 First generation – hydroxamate-based inhibitors
Journal Pre-proof
The first generation MMPIs are mainly small molecules or peptides containing the hydroxamate zinc-binding group (-CONHOH), which shows strong interaction with the Zn2+ ion on the catalytic domain of MMPs (Whittaker et al., 1999). The active
site
of
all
MMPs
shares
a
highly-conserved
sequence
motif
(His-Glu-Xxx-Gly-His-Xxx-Xxx-Gly-Xxx-Xxx-His). Because the three histidine
f
residues form coordinate bonds with the catalytic Zn2+ ion, the hydroxamate group
oo
has the capability of inhibiting a broad-spectrum of MMPs (Bode et al., 1999). For
pr
example, Batimastat, the first MMPI to enter clinical trials for cancer, inhibits MMP-1,
e-
-2, -7, and -9 (Wojtowicz-Praga et al., 1996). However, due to its poor oral
Pr
bioavailability, Batimastat was superseded by Marimastat, which retained the property of its predecessor, but manifested better oral bioavailability (Nemunaitis et al., 1998).
al
Prinomastat, another hydroxamate-based inhibitor, shows high affinity toward
rn
MMP-2, -3, -9, -13, and -14 over MMP-1 (Hande et al., 2004).
Jo u
Although animal-derived pre-clinical results of these hydroxamate-based inhibitors were fairly promising (Davies et al., 1993; Giavazzi et al., 1998), all subsequent clinical trials failed because of several reasons. Firstly, severe side effects such as musculoskeletal syndrome (MSS) and inflammation occurred in patients, which were not observed in the pre-clinical models (Skiles et al., 2004). These side effects were probably attributed to the broad-spectrum inhibiting property of the hydroxamate group. This group also inhibits the ADAMs (a disintegrin and metalloproteinase) family and ADAMTSs (ADAMs with thrombospondin motifs) family, two large related families of MMPs (Fig. 1), which play important roles in
Journal Pre-proof
cellular regulation, cell- matrix interactions and membrane protein shedding (de Arao Tan et al., 2013; Edwards et al., 2008). Secondly, cancer patients who enrolled in the clinical trials were mostly at the advanced stage of the disease. By contrast, animal models that are utilized in the pre-clinical studies manifested only the early stage of cancer. While MMPs produced by stromal cells are important during the early aspects
f
of cancer progression (i.e. local invasion), they may no longer be required once
oo
metastases have been established (Zucker et al., 2000). Thirdly, hydroxamate-based
pr
inhibitors were labile and unstable; their rapid clearance rate and short half- lives
e-
hindered their clinical application (Peng et al., 1999).
Jo u
rn
al
Pr
2.2 Second generation – non-hydroxamate-based inhibitors
Fig. 2 Schematic representation of four zinc-binding groups (a: hydro xamate, b: thiolate, c: carbo xylate, and d: phosphinate) in complex with the zinc ion
To solve the problems associated with first generation MMPIs, the pharmaceutical industry began to shift its attention to non- hydroxamate-based inhibitors such as thiolate, carboxylate and phosphinate (Fig. 2) (Devel et al., 2010).
Journal Pre-proof
The most significant difference between non-hydroxamate-based inhibitors and hydroxamate-based inhibitors is their affinity for the catalytic Zn2+ ion. The hydroxamate is a strong Zn2+-chelating group, which anchors itself tightly to the Zn2+ ion. This strong anchoring mechanism hinders the free orientation of the rest of the compound. On the contrary, the non-hydroxamate group is a relatively weak
f
Zn2+-chelating group. Although there is some reduction of the inhibitor ’s potenc y, this
oo
weak binding property reduces off-target binding to the ADAMs and ADAMTSs
pr
families, thereby preventing side effects to a certain extent (Skiles et al., 2001). As
e-
reported by Overall and Kleifeld (2006a), introduction of the hypophosphite group
Pr
increased the selectivity towards MMP-1 (Ki = 5 nM), compared with MMP-7 (Ki = 78 nM) by one order of magnitude. Likewise, introduction of the carboxylate group
al
increased MMP selectivity by two orders of magnitude (Ki = 2 nM for MMP-1 and Ki
rn
= 850 nM for MMP-7). Despite these improvements, it is now well accepted that an
Jo u
ideal MMP-selective inhibitor should increase the difference in Ki over targets by three log orders. This is difficult to achieve in the catalytic zinc sites because of their highly homogeneous sequences (Overall and Kleifeld, 2006a). Clinical results conducted on non- hydroxamate-based inhibitors were mixed with success and failure. Because of their weak Zn2+-chelating ability, the rates of severe MSS decreased dramatically compared with the hydroxamate-based inhibitors. Although this was extremely encouraging, the invalid and even negative treatment effects of most non- hydroxamate-based inhibitors (except doxycycline) still casted a shadow on the ongoing and future clinical trials (Table 1). For example, Tanomastat,
Journal Pre-proof
which contains a thioether zinc-binding group, had no serious MSS reported, but efficacy was invalid and even negative results were detected compared with the placebo arm in small-cell lung cancer patients (Erlichman et al., 2001; Hirte et al., 2006; Moore et al., 2003; Rigas et al., 2003). Metastat (COL-3), a tetracycline derivative inhibitor, had contradictory treatment effects in different diseases, and the
f
high rates of photosensitivity reaction despite fastidious sun protection hindered the
oo
enthusiasm for pharmaceutical companies to proceed to Phase III trials (Chu et al.,
pr
2007; Cianfrocca et al., 2002; Rudek et al., 2001). Rebimastat (BMS-275291)
e-
(Douillard et al., 2004; Leighl et al., 2005; Miller et al., 2004; Rizvi et al., 2004),
Pr
S-3304 (Chiappori et al., 2007; Van Marle et al., 2005) and AZD1236 (Dahl et al., 2012; Magnussen et al., 2011) all showed good tolerability, but no improved survival
al
and convincing treatment effects were observed (Table 1).
rn
Setting all these failures aside, the only MMPI which gave people relief was
Jo u
doxycycline. As mentioned above, the Periostat® (doxycycline at sub-antimicrobial dose - 20 mg twice daily) is the only MMPI approved by FDA to date. Except for its outstanding performance in treating periodontitis, doxycycline at different dose levels also exhibited acceptable therapeutic effects in treating aortic aneurysms, multiple sclerosis as well as type II diabetes (Baxter et al., 2002; Frankwich et al., 2012; Minagar et al., 2008). Further Phase III trials are needed in larger populations to provide stronger evidence.
2.3 Third generation – catalytic domain (non-zinc binding) inhibitors Learning from the limitations of the former two generations of MMPIs, scientists
Journal Pre-proof
began to shy away from the scope of the zinc binding group and target the catalytic domain of MMPs instead. In addition to the zinc binding site, the catalytic domain contains many other recognition pockets, half of which are on the r ight side of the Zn2+ ion (called primed side, with pockets named S 1 ’, S2 ’, S3 ’) and the other half on the left side (called unprimed side, with pockets named S 1 , S2 , S3 ) (Fig. 3a). The part of substrates or inhibitors that can fit into the S n pocket is named Pn accordingly (for
oo
f
example, P1 ’ fits in S1 ’, P2 fits in S2 ) (Whittaker et al., 1999). Among these
pr
recognition pockets, S1 ’ varies the most among different MMPs in the amino acid
e-
sequence and pocket depth. Consequently, the ranking of the selectivity of the
Pr
recognition pockets toward substrates or inhibitors are: S1 ’ > S2 , S3 , S3 ’ > S1 > S2 ’ (Aureli et al., 2008; Pirard, 2007). Based on the depth of S1 ’ pocket or S1 ’ cavity (Fig.
al
3c), MMPs may be separated into shallow (MMP-1, -7), intermediate (MMP-2, -8, -9),
rn
and deep pocket (MMP-3, -11, -12, -13, -14) (Aureli et al., 2008; Jacobsen et al., 2010;
Jo u
Park et al., 2003). The difference has been utilized in developing selective MMPIs. For example, inhibitors with a long P 1 ’ substituent like aromatic groups can fit in the large deep pocket of MMP-3, but are not accessible to MMP-1 and -7 because of their limited pocket depth (Terp et al., 2002). This method increased the selectivity to some extent; however, the intermediate and other similar deep pocket MMPs cannot be distinguished by simply elongating the P1 ’ substituent. With the development of X-ray crystallography, nuclear magnetic resonance spectroscopy, high-throughput screening and computational methods, structural differences in the S1 ’ pocket depth have been gradually recognized. A variety of
Journal Pre-proof
highly-selective MMPIs with up to three orders of magnitude inhibition capacity have been developed (Matter and Schudok, 2004; Overall and Kleifeld, 2006a; Rao, 2005). For example, Pochetti et al. (2009) reported a selective inhibitor of MMP-8, which
Jo u
rn
al
Pr
e-
pr
oo
f
induces a conformational change of the S 1 ’ loop (Fig. 3b). This, in turn, exposed an
Jo u
rn
al
Pr
e-
pr
oo
f
Journal Pre-proof
extra binding region of the pocket by switching of the Y227 side chain, thereby improving the selectivity of MMP-8 (IC50 = 7.4 nM). Engel et al. (2005) identified another selective inhibitor, the pyrimidine dicarboxamide, which not only binds to the deep S1 ’ pocket of MMP-13, but also protrudes into a side pocket of S 1 ’, a feature that
Journal Pre-proof
has not been observed in other MMPs (Fig. 3d). This observation led to an improved IC50 of 8 nM toward MMP-13 and against MMP-1, -2, -3, -7, -8, -9, -10, -12, -14, and -16 (IC50 > 100 μM). Inspired by the work of Engel et al. (2005), Gege et al. (2012) synthesized a highly-selective MMP-13 inhibitor with IC 50 = 0.03 nM, which is 20,000-fold over other MMPs.
f
Fig. 3 Representation of the X-ray structure of MMP catalytic domain. The Zn 2+ appears as a
oo
purple dot. (a) The general location of S 3 -S3 ’ pockets indicated by circles (PDB code: 2OY4,
pr
uninhibited human MMP-8). (b) The indication of S 1 ’ loop, which is conformational flexible (PDB
e-
code: 3DPE, the complex between MMP-8 and a non-zinc chelating inhibitor). (c) The inhibitor is
Pr
highlighted to indicate the S 1 ’ pocket depth (PDB code: 3DPE, the complex between MMP-8 and a non-zinc chelating inhibitor). (d) The inhibitor is highlighted to indicate its protrusion into S 1 ’
Jo u
rn
Protein Data Bank.
al
side pocket (PDB code: 1XUD, MMP-13 complexed with non-zinc binding inhibitor). PDB:
Apart from the synthesized MMPIs, many natural compounds have also been shown to possess selective inhibition via adaptation with specific MMP pockets. Wang et al. (2012) identified 19 potential MMPIs from 4,000 natural compounds isolated from 100 medicinal plants using structure-based virtual screening. After further testing, two classes of natural compounds (namely, caffeates and flavonoids) were found to have selective inhibition ability against MMP-2 and MMP-9 by occupying the S1 ’ and S3 pockets (Wang et al., 2012). Apart from plant-derived natural compounds, marine natural products are another huge pharmacological
Journal Pre-proof
resource that have great potential in targeting MMPs (Gentile and Liuzzi, 2017). The major MMPIs of the marine environment are derivatives from algae, sponges and cartilage, such as Neovastat, Dieckol, Ageladine A and chitin. The y all manifest anti-angiogenic, anti-proliferative and anti-tumor effects to some extent (Agrawal et al., 2018).
f
Compared with synthetic MMPIs, natural MMPIs are more bio- friendly and less toxic.
oo
However, they still have several drawbacks. Firstly, the effective dosage of most
pr
natural MMPIs lie in micromolar scale, which is thousands of times higher than the
e-
third generation synthetic MMPIs, not to mention the forthcoming protein-engineered
Pr
MMPIs which lie in the picomolar scale (Arkadash et al., 2017; Engel et al., 2005; Gege et al., 2012). Despite the reduced cytotoxicity for natural MMPIs per unit, the
al
overall side-effects on human subjects still require further evaluation via
rn
well-designed clinical trials. Secondly, some natural compounds that possess
Jo u
anti-MMP effects are mixtures derived from plants or the marine environment, such as the grape seed extract and Neovastat (isolated from shark cartilage). They have encountered the same challenges as with Chinese medicine, in that the effective chemical molecule cannot be defined. Consequently, even though the Phase II clinical trials were successful, the inability to set proper surrogate markers prevented the detection of potential clinical benefits in the Phase III clinical trials (Batist et al., 2002; Latreille et al., 2003; C. Lu et al., 2010). Thirdly, because it is difficult to patent these natural compounds, pharmaceutical companies and other investors are reluctant to sponsor large-scale clinical trials. Hence, data collected for potential natural MMPIs
Journal Pre-proof
are inadequate to support their superiority over synthetic MMPIs.
2.4 Fourth generation -- allosteric and exosite inhibitors Apart from the catalytic domain, MMPs contain other distal structures such as the hemopexin domain, the collagen binding domain and the pro-peptide domain. Because these domains are far remote from the catalytic domain, previous research
oo
f
had not devoted much attention to their behaviors. As the full- length MMP structures are gradually unveiled, the functions of these additional domains in substrate
pr
recognition, signal transmission as well as protein-protein interactions are gradually
e-
recognized. These non-catalytic domains may be further utilized in designing
Jo u
rn
al
Pr
selective MMPIs (Sela-Passwell et al., 2010).
Journal Pre-proof Table 1 Summary of MMP Inhibitors Clinical Trials MMPI name Batimastat
Structure Hydroxamate
Year 1996
Disease
Stage
Malignant tumors
Phase I
Treatment
(BB-94)
600,
Marimastat
1,200,
1,800
mg/m
2
via
Results
References
Mainly local toxicities, especially abdominal discomfort.
Wojtowicz-Praga
intraperitoneal injection Hydroxamate
1998
Malignant tumors
Phase I/II
(BB-2516) 2001
Pancreatic cancer
Phase III
et
al., 1996
5 mg to 100 mg bid or 5 mg to 25 mg
MSS increased in dosage over 50 mg bid after 1 month
qd
usage. Recommend 10 and 25 mg bid.
1998
5, 10, 25 mg bid vs. gemcitabine
1-year survival rate: 25 mg = gemcitabine > 10 mg = 5 mg;
Bramhall et al., 2001
f o
ro
Nemunaitis
et
al.,
MSS: 44% of marimastat > 12% of gemcitabine. 2002
Pancreatic cancer
Phase III
2002
Gastric cancer
Phase III
Gemcitabine with 10 mg bid or
-p
placebo
2004
MMI-270
Hydroxamate
2001
Metastatic breast cancer
Advanced solid cancer
Phase III
Prinomastat
Hydroxamate
2004
J
Advanced cancer
(AG3340) 2005
2006
NSCLC (Stage IIIB or IV)
Resectable
esophagus
Phase I
Phase I
Biphenyl,
(Bay 12-9566)
thioether
2001
Solid tumor
tolerated.
al., 2002
Positive treatment effect. Severe MSS observed.
Bramhall,
No difference in treating effects. MSS was associated with
Hallissey,
et al., 2002 Sparano et al., 2004
inferior survival. Rash and MSS observed in doses ≥ 300 mg bid. No
50 mg qd to 600 mg tid
5-fluorouracil
Bramhall, Schulz, et
Levitt et al., 2001
obvious tumor response. The recommended Phase II dose was 300 mg bid. +
folinic
acid
+
Severe MSS observed in 300 mg bid. Six patients (18.2%)
MMI-270 (50 mg qd, 150 mg tid or
had a partial response and 17 patients (51.5%) had stable
300 mg bid)
disease.
1 mg to 100 mg bid
No tumor response. 5-10 mg bid were recommended for
Eatock et al., 2005
Hande et al., 2004
longer treatment. Phase III
Phase II
cancer (Stage II or higher) Tanomastat
al
rn
u o
Colorectal cancer
e r P
10 mg bid vs. placebo
Phase I
(CGS27023A)
2005
10 mg bid vs. placebo
No treatment effect; gemcitabine + marimastat was well
Phase I
Gemcitabine + cisplatin with 15 mg
Terminated early because of lack of treatment efficacy.
bid or placebo
MSS incidence increased.
5-FU + cisplatin + paclitaxel with 15
Early termination due to unexpected thromboembolic
mg bid or placebo
events. No definitive conclusions were obtained.
400 mg qd, 400 mg bid, 400 mg tid
Toxicity was mild. No MSS. No tumor responses. Due to
and 800 mg bid.
the low toxicity, 800 mg bid could be further tested.
Donald et al., 2005
Heath et al., 2006
Erlichman et al., 2001
Journal Pre-proof zinc-binding group 2003
SCLC and NSCLC
Phase III
800 mg bid vs. placebo
For NSCLC, the TDP was significantly increased.
Rigas et al., 2003
However, for SCLC, the TDP was significantly decreased. 2003
Pancreatic cancer without
Phase III
800 mg bid vs. gemcitabine
Early termination due to the poorer survival. Both arms had
prior chemotherapy 2006
Moore et al., 2003
low rates of serious toxicity.
Ovarian cancer with prior
Phase III
f o
Early termination due to Bayer’s decision. Well tolerated
800 mg bid vs. placebo
effective treatment
Hirte et al., 2006
without grade 3 or 4 adverse events, but has no positive
ro
treatment effects. Metastat
Tetracycline
(COL-3)
derivatives
2001
Refractory solid tumors
36, 50, 70, and 98 mg/m 2 qd.
Phase I
AIDS-Related
Kaposi’s
Phase I
25, 50 and 70 mg/m
sarcoma
2007
Cianfrocca
Significant change in MMP-2 levels.
2002
50 mg/m qd, and instructed to apply
Photosensitivity reaction despite fastidious sun protection.
Chu et al., 2007
sun protection procedures.
No obvious treatment effects.
20 mg bid (sub-antimicrobial dose) or
Well tolerability. Outstanding effects without causing
placebo
dysbacteriosis.
20 mg bid (sub-antimicrobial dose) or
Well
placebo
outcomes.
100 mg bid
Well tolerability. No significant change in aneurysms
r P
qd, and
instructed to apply sun protection
Advanced
soft
tissue
sarcoma 2001
Periodontitis
2004
Periodontitis
al
Phase II
n r u
derivatives
Phase II/III
2002
2
Photosensitivity reaction. Overall response rate was 44%.
procedures.
Tetracycline
p e
Rudek et al., 2001
recommended.
2002
Doxycycline
High rates of cutaneous phototoxicity. 36 mg/m2 qd was
o J
Asymptomatic abdominal
Phase III
Phase II
2
aortic aneurysms
tolerability.
Significantly
improvement
in
all
et
al.,
Golub et al., 2001
Preshaw et al., 2004
Baxter et al., 2002
diameter, but significant reduction in plasma MMP-9 level after 6 months treatment.
2008
Multiple sclerosis
Phase II
Interferon beta-1a with doxycycline
The therapy was effective, safe and well-tolerated.
Minagar et al., 2008
Doxycycline decreas ed inflammation and improved insulin
Frankwich
sensitivity.
2012
NCT02774993. No results available, accessed on Sept 4,
NA
100 mg qd 2012
2017
Type II diabetes
Pulmonary Tuberculosis
Phase III
Phase II
100 mg bid or placebo
100 mg bid or placebo
et
al.,
Journal Pre-proof 2019. Rebimastat
Mercaptoacyl,
2004
(BMS-275291)
thiol zinc-binding
mg qd due to free of sheddase inhibition and no arthritis
group
occurred. 2004
Advanced cancer
Phase I
Early stage breast cancer
Phase II
600 to 2,400 mg qd.
(Stage I (T1c) to IIIA)
Conventional
2004
NSCLC in stage IIIB or IV
Phase II
without chemotherapy
Good tolerability. No tumor responses. Recommend 1,200
therapy
with
High dropout rates due to adverse effects forced the clinical
BMS-275291 1,200 mg qd or placebo
trial to terminate early.
Paclitaxel
with
Well tolerated. The overall response rate of BMS-275291
BMS-275291 1,200 mg qd or placebo
arm was lower than that of the placebo arm (21.9% versus
+
carboplatin
2005
NSCLC in stage IIIB or IV
Phase III
without chemotherapy
Paclitaxel
S-3304
Sulfonamide
2005
Health male volunteer
Phase I
AZD1236
Non-hydroxamate
Neovastat
Mixed
(AE-941)
from
extract
Advanced solid tumors
n r u
Moderate to severe COPD
2012
Moderate to severe COPD
2003
shark
o J
Lung cancer
Phase II
Phase II
Phase I/II
with
Higher hypersensitivity and febrile neutropenia rates, but
BMS-275291 1,200 mg qd or placebo
no difference in MSS. No improved survival in advanced
al
Phase I
2011
carboplatin
p e
r P
10-800 mg bid versus placebo
derivatives 2007
+
f o
ro
36.4%).
800-3200 mg bid
Rizvi et al., 2004
Miller et al., 2004
Douillard et al., 2004
Leighl et al., 2005
NSCLC. Good tolerability. Good systemic exposure and free of MSS
Van Marle et al., 2005
at 800 mg bid. Good tolerability at 3200 mg bid. MMP was inhibited at
Chiappori et al., 2007
lower dose.
75 mg bid or placebo
Good tolerability. No significant treatment effects.
75 mg bid or placebo
The rates of adverse events were similar. No clinical
Magnussen
et
2011 Dahl et al., 2012
efficacy was observed. 30 to 240 mL/day bid
Good tolerability. No tumor responses were observed, but a
Latreille et al., 2003
dose-dependent increase in median survival time was
cartilage
found. 2002
Renal cell carcinoma
Phase II
60 mL/day bid or 240 mL/day bid
Good tolerability. 240 mL/day bid increased the median
Batist et al., 2002
survival time. 2010
NSCLC in stage III
Phase III
al.,
Chemotherapy with 120 mL/day bid
Good tolerability. No difference in overall survival, TDP
or placebo
and tumor response.
Lu et al., 2010
Journal Pre-proof
Clinical trials resumed after nearly ten years of trials and tribulations
FP-025
Non-hydroxamate
2017
Healthy male subjects
Phase I
50 to 800 mg qd
NCT02238834. No results available, accessed Sept 4,
f o
2019.
GS-5745
mAb inhibitor
2019
Allergic asthma
Phase II
400 mg bid or placebo
2015
Advanced solid tumors
Phase I
(Andecaliximab)
2016
2018
2018
Active ulcerative colitis
Rheumatoid arthritis
Active Crohn’s disease
NA
1,800 mg IV, followed by 800 mg IV
Manageable safety profile for GS-5745 alone or in
Bendell et al., 2015
with chemotherapy
combination with chemotherapy. More subjects need to be
o r p
e
enrolled.
Either IV infusions (0.3 to 5.0 mg/kg
Good tolerability and no MSS
or placebo) every two weeks; or five
endoscopic and histological responses were positive. The
weekly SC injections (150 mg or
dose 1 mg/kg IV, 2.5 mg/kg IV and 150 mg SC were
placebo)
recommended.
r P
al
Clinical,
n r u
3 times.
hypersensitivity (13.3%).
800 mg GS-5745 + mFOLFOX6 IV
The treatment showed encouraging clinical activity without
every two weeks
additional toxicity, especially in first-line patients with an
Phase II
150 mg SC weekly, 300 mg SC
Good tolerability. No treatment effects were observed after
weekly, 150 mg SC every two weeks,
8 weeks.
Phase I
The highest
reported.
Good tolerability.
o J
Phase I
NCT03858686. Under recruiting, accessed Sept 4, 2019.
400 mg IV every two weeks for total
Gastric/gastroesophageal junction adenocarcinoma
2018
Phase I
NA
adverse event
was
Sandborn et al., 2016
Gossage et al., 2018
Shah et al., 2018
overall response rate at 50%. Schreiber et al., 2018
and placebo 2018
2016
Ulcerative colitis
Gastric/gastroesophageal
Phase
150 mg SC weekly, 150 mg SC every
Good tolerability. Terminated at 8 weeks because of lack of
II/III
two weeks
efficacy.
Phase III
800 mg GS-5745 + mFOLFOX6 IV
NCT02545504. No results available, accessed Sept 4,
every two weeks or placebo
2019.
Various
NCT03631836. Has no recruit yet, accessed Sept 4, 2019.
junction adenocarcinoma 2019
Recurrent glioblastoma
Phase I
doses
of GS-5745 with
Sandborn et al., 2018
Bendell et al., 2016
NA
Journal Pre-proof Bevacizumab 2019
Gastric/gastroesophageal
Phase II
junction adenocarcinoma
Nivolumab + GS-5745 800 mg IV
NCT02864381. Has no recruit yet, accessed Sept 4, 2019.
NA
every two weeks or Nivolumab alone
*Abbreviations: qd once daily; bid twice daily; tid three times daily; TDP time to disease progression; SCLC small-cell lung cancer; NSCL C non-small-cell lung cancer; MSS musculoskeletal syndrome; COPD chronic obstructive pulmonary disease; mAb monoclonal antibody; IV intravenous injection; SC subcutaneous administration
f o
l a n
J
r u o
r P
e
o r p
Journal Pre-proof
The hemopexin domain has a four-blade propeller structure, which is linked to the catalytic domain via a flexible hinge (Fig. 1) (Cha et al., 2002). Research have demonstrated that the hemopexin domain and the catalytic domain are conformationally independent from each other. However, during substrate recognition and collagen hydrolysis, the hemopexin domain can move relative to the catalytic domain and allosterically manipulate enzymatic
f
activity (Bertini et al., 2008; Jozic et al., 2005; Overall and Butler, 2007). The term “allostery”
oo
here refers to the effect exerted by the distal structures, which regulate enzyme activity by
pr
conformational deformation (Sela-Passwell et al., 2010). The predominant advantage of
e-
allosteric drugs is that they provide non-competitive inhibition compared with traditional
Pr
orthosteric compounds through the flexible dynamics of protein structure (Lu et al., 2014; Mitternacht and Berezovsky, 2011). This feature enables an allosteric inhibitor to be free
al
from competition with the well-conserved catalytic site, with the advantage of avoiding
rn
off-target inhibition and preventing the occurrence of clinical side effects (Guarnera and
Jo u
Berezovsky, 2016). For MMPs, the long hinge is the prerequisite for flexible movement of the hemopexin domain. This long hinge regulates various functions of MMPs, including zymogen activation, oligomerization, substrate hydrolysis and inhibition through allosteric control (Sela-Passwell et al., 2010). Accordingly, targeting the hemopexin domain has been perceived as an effective means for designing selective inhibitors. For example, Remacle et al. (2012) reported a small molecule NSC405020 from the library of the DTP National Cancer Institute, which binds selectively to the hemopexin domain of MMP-14. This small molecule inhibits homodimerization of MMP-14 and the interaction between the hemopexin domain with the catalytic domain, thereby preventing degradation of type I collagen and halting
Journal Pre-proof
tumor growth. Apart from allosteric control, exosites provide another alternative for selective MMPIs design. The term “exosite” is a relative concept for the catalytic site, which refers to the alternative binding site in MMP structures (Lauer-Fields et al., 2008). The major difference between allosteric sites and exosites is that the former communicates with the catalytic site
f
through a dynamic conformational change, whereas the latter encompasses a much broader
oo
concept. To date, several exosites have been identified, including the hemopexin domain,
pr
collagen binding domain and pro-peptide domain (Overall, 2002); their corresponding
e-
selective inhibitors had been designed. For example, Dufour et al. (2010) developed a
Pr
structure-based inhibitory peptide targeting the MMP-9 hemopexin domain, which efficiently blocks MMP-9 dimer formation and cell migration. Xu et al. (2007) synthesized a peptide
al
targeting the collagen binding domain of MMP-2, which inhibits hydrolysis of type I and
rn
type IV collagen. Scannevin et al. (2017) identified a highly-selective compound named
Jo u
JNJ0966, which binds to the pro-peptide domain of MMP-9 and inhibits activation of pro-MMP-9 without affecting MMP-1, -2 and -14. Because the binding affinity of most exosites for substrates is typically low (10 -6 - 10-7 M), the IC50 of exosite inhibitors is in the micromolar range. Hence, exosite inhibitors do not manifest outstanding superiority over the allosteric inhibitors (Overall and Kleifeld, 2006a; Scannevin et al., 2017; Xu et al., 2007). It has been suggested recently that each MMP possesses its unique exosites or “hotspots” that may be targeted individually due to difference in their amino acid composition and diversity in geometry (Levin et al., 2017). In the future, highly-selective MMPIs may be synthesized via the combined inhibition of both the active
Journal Pre-proof site and the MMP-specific “hotspots”.
3. On-going research techniques Progress in medicine is always accompanied by technological innovation and knowledge extension. The aforementioned four generations of MMPIs are vivid examples of progressive cognitive advancement, with the inhibition target expanding from the limited small zinc
oo
f
binding site to full- length MMP structures. From a technological perspective, however, there has been little advancement on MMPI development, as they were mainly small molecules or
pr
peptides selected from biochemical or biophysical libraries that lacked direct synthesis based
which have greatly improved MMPI design,
namely
Pr
two on- going techniques,
e-
on specific “hotspots” (Rouanet-Mehouas et al., 2016). In the following part of the review,
al
antibody-based inhibitors and protein-engineering inhibitors, will be presented (Fig. 4).
rn
3.1 Antibody-based inhibitors
Jo u
Antibody, also known as immunoglobulin, is a large Y-shaped protein which binds to an antigen via the Fab variable region (Litman et al., 1993). Monoclonal antibodies (mAb) are antibodies that are produced by identical immune cells that are all clones of a unique parent cell. For almost any substance, it is possible to produce the corresponding mAb via phage display, single B cell culture, single-cell amplification and single plasma cell interrogation technologies. These techniques provide useful and importa nt tools in biochemistry, pharmacy and molecular biology investigations. To date, several mAbs directed against MMPs have been developed, with outstanding selectivity. One of the first MMP-9 antibodies reported was REGA-3G12, a highly-selective
Jo u
rn
al
Pr
e-
pr
oo
f
Journal Pre-proof
antibody against MMP-9, but not MMP-2 (Paemen et al., 1995). REGA-3G12 specifically targets the N-terminal region (Trp116 - Lys214 ) of the catalytic domain of MMP-9, but not the catalytic zinc ion or the collagen binding domain (Martens et al., 2007). DX-2400 is an antibody
isolated
from a phage display
library.
It targets MMP-14
and
the
MMP-14-dependent proMMP-2 processing, decreasing MMP activity up to 70% in animal models (Devy et al., 2009). Unlike DX-2400, the antibody LEM-2/15 only selectively inhibits MMP-14 catalytic activity toward gelatin and collagen type I without affecting
Journal Pre-proof
proMMP-2 activation and MT1-MMP dimerization (Udi et al., 2015). Another antibody of MMP-14, named 9E8, has no effect on MMP-14 hydrolysis, but inhibits proMMP-2 activation (Shiryaev et al., 2013). Although DX-2400, LEM-2/15 and 9E8 are all selective MMP-14 antibody inhibitors, subtle differences exist in their regulation of downstream functions (Winer et al., 2018).
f
Fig. 4 The evolution of MMP inhibitors. Cognitive advancement (blue line) and technique innovation
oo
(orange line) are two pillars that support the progression of MMPIs. They not only move forward
pr
individually, but also synergistically create the first to fourth generation of MMPIs (represented by red dots)
e-
and the potential future generations by their convergence. The rationale of the four generations of MMPIs
Pr
are manifested by four representative pictures (A-D) respectively, and their features are listed on the right. In the close-up views, the catalytic zinc ion is portrayed as a green sphere; the active site His are portrayed
al
as sticks in blue and white; and the zinc ion interactions are represented by purple dash lines. (A) MMP-13
rn
complexed to a hydroxamate-based inhibitor (PDB code: 2D1N, two chelating linkages are formed
Jo u
between Zinc ion and inhibitor). (B) A carboxylic acid inhibitor in complex with MMP-3 (PDB code: 1HY7, one chelating linkage is formed between Zinc ion and inhibitor). (C) Crystal str ucture of the complex between MMP-8 and an inhibitor enters into S 1 ’ pocket (PDB code: 3DPE, no chelating linkage is formed between Zinc ion and inhibitor). (D) Structure of a selective antibody inhibitor (GS-5745) bound to MMP-9 (PDB code: 5TH9, targeting at exosites other than catalytic domain). PDB: Protein Data Bank.
Another MMP-9 monoclonal antibody, GS-5745, is a highly-selective antibody (IC 50 = 0.148 nM) targeting four residues (R162, E111, D113, and I198) near the catalytic zinc ion. GS-5745 exerts allosteric control over tumor growth and metastasis in a colorectal carcinoma
Journal Pre-proof
model (Marshall et al., 2015). To the best of our knowledge, GS-5745 (Andecaliximab) is the only mAb inhibitor that has undergone clinical trials (Table 1). A Phase I trial results suggest that GS-5745 is well tolerated by patients with ulcerative colitis and produces positive clinical, endoscopic and histological outcomes (Sandborn et al., 2016). Another Phase I trial examined the effect of GS-5745 in combination with modified-FOLFOX6 (5- fluorouracil,
f
leucovorin and oxaliplatin) in human subjects with advanced gastric /gastroesophageal
oo
junction adenocarcinoma. The GS-5745 treatment achieved target engagement with no
pr
dose-limiting toxicity, with tuning down of collagen neoepitope levels (MMP-9) that
e-
suggested a therapy-related effect (Shah et al., 2018). The positive results of this Phase I trial
Pr
led to the implementation of a Phase III trial which is currently being conducted with the same treatment on the same disease with a larger anticipated population (NCT02545504,
al
accessed Sept 4, 2019) (Bendell et al., 2016). Except for these ongoing clinical trials, more
rn
GS-5745 trials are now recruiting volunteers to explore its potential effects in different
Jo u
diseases (NCT03631836 and NCT02864381, accessed Sept 4, 2019). Despite the promising effect of antibody-based MMPIs, there are still problems that need to be addressed. For example, binding sites of antibodies are in a planar or concave shape, which do not fit well with the active sites of MMPs (pocket shape) (Levin et al., 2017). To solve this problem, Sircar et al. (2011) suggested replacing the human antibody H3 with the camelid antibody CDR-H3, which is a much longer convex-shaped paratope. After synthesizing a library of CDR-H3 modified antibodies, Nam et al. (2016) evaluated the inhibitory potency of targeting the catalytic domain of MMP-14. The results turned out to be overwhelmingly positive (IC 50 = 9.7 nM), and the hybrid antibody showed higher selectivity
Journal Pre-proof
toward MMP-14 compared with MMP-2 or MMP-9. Despite these positive results, more research is needed to evaluate the biosafety and reproducibility in vivo. Apart from the shape fitness problem, the use of antibodies is limited to intravenous or subcutaneous administration, rather than oral administration, which strictly restricts its application (Vandenbroucke and Libert, 2014). Further studies are required to ameliorate antibody usage
f
in a more friendly and convenient way.
oo
3.2 Protein-engineered inhibitors
pr
Protein-engineering technologies, including computational design and directed evolution,
e-
have been reported to be effective and valuable methods in generating desired proteins with
Pr
novel biologic functions (Chao et al., 2006; Levin et al., 2013). Computational design facilitates the exploration of an astronomical number of protein sequences in silico to screen
al
optimal protein binders, which are then experimentally verified by site-directed mutagenesis
rn
(Mohan et al., 2016). Directed evolution, which is based on gene mutant libraries, screens the
Jo u
desired proteins using high-throughput technology and display systems (phage, bacterial, yeast, etc.) by directed force (Banta et al., 2013). With the combination of these two methods, the protein-engineering technique appears to be the optimal technology for screening highly-selective MMPIs in an economic and efficient manner (Gebauer and Skerra, 2009). The most popular group of MMPIs derived from protein-engineering is a family of engineering N-terminal TIMPs. As previously discussed, there are four endogenous TIMPs (TIMP1-4) in the human body that form a 1:1 complex with MMP in the catalytic domain using their N-terminus peptide non-selectively (Maskos and Bode, 2003). The isolated N-TIMP is a stable peptide that retains inhibitory ability against various MMPs (Murphy et
Journal Pre-proof
al., 1991). Because of their relatively small sizes, N-TIMPs may be generated via microbial cultures to create potent inhibitors for direct delivery into cancer cells (Mohan et al., 2016). Accordingly, N-TIMPs have been perceived as convenient backbones for protein-engineering to increase MMP selectivity. Harnessing on the favorable reports that binding affinity increases up to 10- fold with
f
only one site mutation of the N-TIMP (Grossman et al., 2010; Nagase and Brew, 2002;
oo
Sharabi et al., 2014), Wei et al. (2003) increased the mutant sites to three
pr
(Thr2Ser/Val4Ala/Ser68Tyr) and reported that N-TIMP-1 binds to MMP-3 with picomolar
e-
affinity. With the help of computational design, Arkadash et al. (2017) screened 7 potential
Pr
positions in N-TIMP-2 that may augment binding affinity. With 20 amino acids present in seven positions, a mutant library comprising over 1.3 × 109 combinations is generated. Using evolution
and
yeast
surface
display,
N-TIMP-2
with
five
mutations
al
directed
rn
(Ile35Met/Asn38Asp/Ser68Asn/Val71Gly/His97Arg) has been shown to be the mo st potent
Jo u
inhibitor against MMP-14 (Ki = 0.9 pM).
4. Future prospects of next generation of MMPIs The past decade was a relatively repressive time for the development of MMPIs. Almost no clinical trial was conducted, and increasing doubts have been raised on the feasibility of MMPIs in treating the associated diseases. With the development of X-ray crystallography, nuclear magnetic resonance spectroscopy, high-throughput screening and computational methods, the full- length structures of MMPs are gradually recognized. The S’ pocket, allosteric sites, and exosites brought new innovation to MMPI design. The interaction
Journal Pre-proof between monoclonal antibody with unique “hotspots” and the protein engineering technique provide efficient tools for direct protein synthesis and screening. Phase I clinical trial for GS-5745 had been successfully implemented and Phase III trial is currently being conducted after an intellectual vacuum for almost 10 years (Bendell et al., 2016). All these events have created a new era for MMP inhibitors (Fig. 5). In the following section, we provide possible
f
future directions based on our own perceptions.
Jo u
rn
al
Pr
e-
pr
oo
4.1 Enhancing binding site specificity
Journal Pre-proof
The combination of different inhibition sites and different technologies has been perceived as the most feasible way to increase the selectivity of MMPIs. Udi et al. (2013) proposed the notion of the existence of a unique region in individual MMPs that manifests the highest binding energy for ligand binding. These regions are potential selective “hotspots”
Jo u
rn
al
Pr
e-
pr
oo
f
for drug design. Because of technological limitations, the authors had to synthesize a series of
Journal Pre-proof
polymers with different degrees of branching to screen the most selective entity through time-consuming experiments. With the advent of protein engineering techniques, one can now build a mutant library based on the unique amino acids held by different MMPs in silico, and screen the highly potent selective inhibitors by computational simulation (Arkadash et al.,
Jo u
rn
al
Pr
e-
pr
oo
f
2017). The early MMPIs were small molecules or peptides that could only target one binding
Journal Pre-proof
site, or extended to the nearby side pockets at most. With understanding of the hotspots of each MMP, macromolecular proteins, which inhibit the catalytic site as well as specific hotspots simultaneously, may be the future goal for protein engineering techniques. 4.2 Smart drug delivery systems for achieving selectivity
Jo u
rn
al
Pr
e-
pr
oo
f
Apart from MMP binding site specificity, smart drug delivery systems, which directly
Journal Pre-proof
target the designated position of MMP inhibition, are another important method to reduce off-target side effects. Matrix metalloproteinases play important roles in tumor-associated processes such as metastasis, tumor growth, angiogenesis, apoptosis and immune modulation (Fingleton, 2006). Upregulation of MMPs have been found to be positively-correlated with the development of cancer (Egeblad and Werb, 2002). For example, in benign ovarian tissues, the MMP-2 concentration is around 0.47 μg/mg, whereas the value increases to 0.78-1.2
oo
f
μg/mg in ovarian cancer patients, depending on the cancer stage (Schmalfeldt et al., 2001;
pr
Zucker et al., 1992). Hence, representative MMPs in specific diseases may be utilized as
e-
therapeutic targets for MMP-responsive “smart” drug delivery systems (Yao et al., 2018).
Pr
Currently, three MMP-responsive drug release methods have been proposed, namely, linker cleavage, structure disassembly and membrane “uncorking” (Yao et al., 2018). In these three
al
methods, the drug molecules are either conjugated to the polymer via an MMP-sensitive
rn
linker, or encapsulated in a degradable nanoparticle or opened by a specific MMP. By using
Jo u
the linker cleavage and membrane “uncorking” methods, Purcell et al. (2014) synthesized an injectable and bioresponsive MMP-sensitive hydrogel containing a recombinant tissue inhibitor of MMPs (rTIMP-3) (Purcell et al., 2014). This hydrogel/rTIMP-3 system produces minimal release in non-targeted tissues (that is, lack of MMP activity) and significantly suspends tissue remodeling in myocardial infarction. Further steps may be implemented, such as encapsulating highly-selective mAb inhibitors within MMP-responsive nanoparticles. Such a strategy may change the drug delivery method from parenteral injection to more patient-friendly oral administration without adversely affecting the function of mAb inhibitors.
Journal Pre-proof Fig 5. Timeline listing milestones in the discovery, design and development of MMPIs
4.3 Thorough understanding of MMPs in health and disease A thorough understanding of the functions of different MMPs in different diseases under different spatiotemporal conditions should be the ultimate goal for scientists in this research
oo
f
arena. Numerous reports have indicated that MMPs play different roles in different biological systems, such as substrate recognition, signal transmission and protein-protein interactions.
pr
These substrates and signal molecules together with MMPs form a dynamic web – the
e-
“protease web”, which entangle with each other and maintain homeostasis in a subtle balance
Pr
(Overall and Kleifeld, 2006b). Subsequently, a single MMP can degrade/activate different
al
substrates that may exert opposite effects on physiologic function or disease progression. As
rn
mentioned previously, the three different mAb inhibitors (DX-2400, LEM-2/15 and 9E8) all
Jo u
target at MMP-14 selectively, but their downstream substrates vary from each other (Devy et al., 2009; Shiryaev et al., 2013; Udi et al., 2015). The latest MMP research opens new views in the understanding of previously unrecognized cell physiology. Working with the neamtode C. elegans, Kelley et al. (2019) demonstrated that in the absence of MMPs, invasive cancerous cells can still penetrate the basement membrane via an alternative method. This is achieved via the formation of F-actin-rich protrusions that physically breach and displace the basement membrane (albeit much slower than MMP- mediated degradation). If such a biological process really occurs in human cells, the theory on prevention of cancer metastasis via MMP inhibition will no longer
Journal Pre-proof
be valid (Castro-Castro et al., 2016; Kessenbrock et al., 2015). Although the role of MMPs in cancer metastasis may be mellowed, their regulation on other aspects of cancer progression, as well as the whole “protease web” should not be overlooked. A subtle shift of emphasis from direct blocking of MMPs to targeting the MMP substrate and “after metastasis” processes may be the future direction for developing new generations of MMPIs. This latest
f
research emphasizes that sophisticated and complex cellular pathways mediated by different
oo
MMPs in different diseases need to be further understood. In addition, specific signal
pr
peptides and structure binding sites need to be further characterized, and well-designed
e-
clinical trials need to be further implemented in order for MMP-strategized therapeutics to be
Pr
successful.
5. Concluding remarks
al
It has been nearly five decades since the discovery of MMP in tadpoles. Since then, the
rn
progress in deciphering the roles of MMPs in human physiologic functioning and disease
Jo u
progression has been nothing short of phenomenal. Although a plethora of reviews on MMP inhibitors have recently been published and provided insights on future development of MMP inhibitors (Fields, 2019a; Fischer and Riedl, 2019; Fields, 2019b; Young et al., 2019; Fischer et al., 2019; Cerofolini et al., 2019), this review is original in the following aspects: (1) the historical development of MMPIs was described under two perspectives, cognitive advancement and technique innovations so that the characteristics of different MMPIs and the reasoning behind the supersession could be easily understood; (2) In the previous reviews, the MMPIs were only defined up to the third generation, the catalytic domain inhibitors. In the current review, the fourth generation MMPIs, the allosteric control and exosit es control
Journal Pre-proof
inhibitors, were described and added to cover the full spectrum of MMPIs; (3) Apart from the synthesized MMPIs, the natural MMPIs derived from plants and marine environments were also included; their pros and cons were discussed; (4) After almos t 10 years of intellectual vacuum, the Phase I clinical trial of newly antibody-based inhibitors had been successfully implemented and Phase III trial is currently being conducted. These novel clinical trials have
f
brought new hope for future developments of MMPIs and they have been summarized along
oo
with previous clinical trials in the current review. Although research on MMP inhibitors was
pr
not as successful as anticipated, the fruits derived from previous studies are invaluable for
should
be
cautiously
optimistic
that
structurally-selective
and
Pr
profession
e-
shaping future research. With the development of new experimental techniques, the
al
spatiotemporally-specific MMPIs with minor off-targeting effects will blossom.
rn
Conflict of Interest Statement
References
Jo u
The authors declare that there are no conflicts of interest.
Agrawal, S., Chaugule, S., & Indap, M . (2018). M arine Pharmaceuticals: A New Wave of Anti-angiogenic Drugs. Journal of Oceanography and Marine Research, 6, 2. Ahrens, D., Koch, A. E., Pope, R. M ., Stein ‐ Picarella, M ., & Niedbala, M . J. (1996). Expression of matrix metalloproteinase 9 (96‐ kd gelatinase B) in human rheumatoid arthritis. Arthritis & Rheumatism: Official Journal of the American College of Rheumatology, 39, 1576-1587. Arkadash, V., Yosef, G., Shirian, J., Cohen, I., Horev, Y., Grossman, M ., Sagi, I., Radisky, E. S., Shifman, J. M ., & Papo, N . (2017). Development of high affinity and high specificity inhibitors of matrix metalloproteinase 14 through computational design and directed evolution. Journal of Biological Chemistry, 292, 3481-3495.
Journal Pre-proof Aureli, L., Gioia, M ., Cerbara, I., M onaco, S., Fasciglione, G. F., M arini, S., Ascenzi, P., Topai, A., & Coletta, M . (2008). Structural bases for substrate and inhibitor recognition by matrix metalloproteinases. Current Medicinal Chemistry, 15, 2192-2222. Banta, S., Dooley, K., & Shur, O. (2013). Replacing antibodies: engineering new binding proteins. Annual Review of Biomedical Engineering, 15, 93-113. Batist, G., Patenaude, F., Champagne, P., Croteau, D., Levinton, C., Hariton, C., Escudier, B., & Dupont, E. (2002). Neovastat (AE-941) in refractory renal cell carcinoma patients: report of a phase II trial with two dose levels. Annals of Oncology, 13, 1259-1263.
oo
f
Bauer, E. A., Stricklin, G. P., Jeffrey, J. J., & Eisen, A. Z. (1975). Collagenase production by human skin fibroblasts. Biochemical and Biophysical Research Communications, 64, 232-240.
pr
Baxter, B. T., Pearce, W. H., Waltke, E. A., Littooy, F. N., Hallett Jr, J. W., Kent, K. C., Upchurch Jr, G. R., Chaikof, E. L.,
e-
M ills, J. L., & Fleckten, B. (2002). Prolonged administration of doxycycline in patients with small asymptomatic abdominal aortic aneurysms: report of a prospective (Phase II) multicenter study. Journal of Vascular Surgery, 36,
Pr
1-12.
Bendell, J. C., Starodub, A., Shah, M . A., Sharma, S., Wainberg, Z. A., & Thai, D. L. (2015). Phase I study of GS-5745 alone
rn
33(15_suppl), 4030.
al
and in combination with chemotherapy in patients with advanced solid tumors. Journal of Clinical Oncology,
Bendell, J. C., Starodub, A., Wainberg, Z. A., Wu, M., Werner, D., M altzman, J. D., & Shah, M . A. (2016). A phase 3
Jo u
randomized, double-blind, placebo-controlled study to evaluate the efficacy and safety of GS-5745 combined with mFOLFOX6 as first-line treatment in patients with advanced gastric or gastroesophageal junction adenocarcinoma. Journal of Clinical Oncology, 34(15_suppl), 4132. Bertini, I., Calderone, V., Fragai, M ., Jaiswal, R., Luchinat, C., M elikian, M ., Mylonas, E., & Svergun, D. I. (2008). Evidence of reciprocal reorientation of the catalytic and hemopexin-like domains of full-length MMP-12. Journal of the American Chemical Society, 130, 7011-7021. Bode, W., Fernandez-Catalan, C., Tschesche, H., Grams, F., Nagase, H., & M askos, K. (1999). Structural properties of matrix metalloproteinases. Cellular and Molecular Life Sciences, 55, 639-652. Bramhall, S. R. , Hallissey, M . T. , Whiting, J. , Scholefield, J. , Tierney, G. , & Stuart, R. C. , et al. (2002). M arimastat as maintenance therapy for patients with advanced gastric cancer: a randomised trial. British Journal of Cancer, 86, 1864-1870. Bramhall, S. R. , Rosemurgy, A. , Brown, P. D. , Bowry, C. , & Buckels, J. A. C. . (2001). M arimastat as first-line therapy for
Journal Pre-proof patients with unresectable pancreatic cancer: a randomized trial. Journal of Clinical Oncology, 19, 3447-3455. Bramhall, S. R. , Schulz, J. , Nemunaitis, J. , Brown, P. D. , Baillet, M . , & Buckels, J. A. C. . (2002). A double-blind placebo-controlled, randomised study comparing gemcitabine and marimastat with gemcitabine and placebo as first line therapy in patients with advanced pancreatic cancer. British Journal of Cancer, 87, 161-167. Castro-Castro, A., M archesin, V., M onteiro, P., Lodillinsky, C., Rossé, C., & Chavrier, P. (2016). Cellular and molecular mechanisms of M T1-MMP-dependent cancer cell invasion. Annual Review of Cell and Developmental Biology, 32, 555-576. Cerofolini, L., Fragai, M ., & Luchinat, C. (2019). M echanism and Inhibition of M atrix M etalloproteinases. Current
f
Chemistry, 26, 2609-2633.
oo
Medicinal
Cha, H., Kopetzki, E., Huber, R., Lanzendörfer, M ., & Brandstetter, H. (2002). Structural basis of the adaptive molecular
pr
recognition by M M P9. Journal of Molecular Biology, 320, 1065-1079.
e-
Chakraborti, S., M andal, M ., Das, S., M andal, A., & Chakraborti, T. (2003). Regulation of matrix metalloproteinases: an overview. Molecular and Cellular Biochemistry, 253, 269-285.
Pr
Chao, G., Lau, W. L., Hackel, B. J., Sazinsky, S. L., Lippow, S. M ., & Wittrup, K. D. (2006). Isolating and engineering human antibodies using yeast surface display. Nature Protocols, 1, 755.
al
Chiappori, A. A., Eckhardt, S. G., Bukowski, R., Sullivan, D. M ., Ikeda, M ., Yano, Y., Yamada-Sawada, T., Kambayashi, Y.,
rn
Tanaka, K., & Javle, M . M . (2007). A phase I pharmacokinetic and pharmacodynamic study of s -3304, a novel matrix metalloproteinase inhibitor, in patients with advanced and refractory solid tumors. Clinical Cancer
Jo u
Research, 13, 2091-2099.
Chu, Q. S., Forouzesh, B., Syed, S., M ita, M ., Schwartz, G., Copper, J., Curtright, J., & Rowinsky, E. K. (2007). A phase II and pharmacological study of the matrix metalloproteinase inhibitor (MMPI) COL-3 in patients with advanced soft tissue sarcomas. Investigational New Drugs, 25, 359. Cianfrocca, M ., Cooley, T. P., Lee, J. Y., Rudek, M . A., Scadden, D. T., Ratner, L., Pluda, J. M ., Figg, W. D., Krown, S. E., & Dezube, B. J. (2002). M atrix metalloproteinase inhibitor COL-3 in the treatment of AIDS-related Kaposi’s sarcoma: a phase I AIDS malignancy consortium study. Journal of Clinical Oncology, 20, 153-159. Coussens, L. M ., Fingleton, B., & M atrisian, L. M . (2002). M atrix metalloproteinase inhibitors and cancer—trials and tribulations. Science, 295, 2387-2392. Dahl, R., Titlestad, I., Lindqvist, A., Wielders, P., Wray, H., Wang, M ., Samuelsson, V., M o, J., & Holt, A. (2012). Effects of an oral MM P-9 and-12 inhibitor, AZD1236, on biomarkers in moderate/severe COPD: a randomised controlled trial. Pulmonary Pharmacology & Therapeutics, 25, 169-177.
Journal Pre-proof Davies, B., Brown, P. D., East, N., Crimmin, M . J., & Balkwill, F. R. (1993). A synthetic matrix metalloproteinase inhibitor decreases tumor burden and prolongs survival of mice bearing human ovarian carcinoma xenografts. Cancer Research, 53, 2087-2091. de Arao Tan, I., Ricciardelli, C., & Russell, D. L. (2013). The metalloproteinase ADAM TS1: a comprehensive review of its role in tumorigenic and metastatic pathways. International Journal of Cancer, 133, 2263-2276. Devel, L., Czarny, B., Beau, F., Georgiadis, D., Stura, E., & Dive, V. (2010). Third generation of matrix metalloprotease inhibitors: Gain in selectivity by targeting the depth of the S1′ cavity. Biochimie, 92, 1501-1508. Devy, L., Huang, L., Naa, L., Yanamandra, N., Pieters, H., Frans, N., Chang, E., Tao, Q., Vanhove, M ., & Lejeune, A. (2009).
oo
f
Selective inhibition of matrix metalloproteinase-14 blocks tumor growth, invasion, and angiogenesis. Cancer Research, 69, 1517-1526.
pr
Donald, B., O'Byrne, K. J., Pawel, J., Von, Ulrich, G., Allan, P., M arianne, N., Richard, M ., Elva, M ., Carol, P., & M in, H.,
e-
Zhang. (2005). Phase III study of matrix metalloproteinase inhibitor prinomastat in non-small-cell lung cancer. Journal of Clinical Oncology Official Journal of the American Society of Clinical Oncology, 23, 842-849.
Pr
Douillard, J.-Y., Peschel, C., Shepherd, F., Paz-Ares, L., Arnold, A., Davis, M ., Tonato, M ., Smylie, M ., Tu, D., & Voi, M . (2004). Randomized phase II feasibility study of combining the matrix metalloproteinase inhibitor BM S-275291
al
with paclitaxel plus carboplatin in advanced non-small cell lung cancer. Lung Cancer, 46, 361-368.
rn
Dufour, A., Zucker, S., Sampson, N. S., Kuscu, C., & Cao, J. (2010). Role of matrix metalloproteinase-9 dimers in cell migration design of inhibitory peptides. Journal of Biological Chemistry, 285, 35944-35956.
Jo u
Eatock, M ., Cassidy, J., Johnson, J., M orrison, R., Devlin, M ., Blackey, R., Owen, S., Choi, L., & Twelves, C. (2005). A dose-finding and pharmacokinetic study of the matrix metalloproteinase inhibitor MM I270 (previously termed CGS27023A) with 5-FU and folinic acid. Cancer Chemotherapy & Pharmacology, 55, 39. Edwards, D. R., Handsley, M. M ., & Pennington, C. J. (2008). The ADAM metalloprot einases. Molecular aspects of medicine, 29, 258-289. Egeblad, M ., & Werb, Z. (2002). New functions for the matrix metalloproteinases in cancer progression. Nature Reviews Cancer, 2, 161. Eisen, A. Z., Bauer, E. A., & Jeffrey, J. J. (1971). Human skin colla genase. The role of serum alpha-globulins in the control of activity in vivo and in vitro. Proceedings of the National Academy of Sciences, 68, 248-251. Engel, C. K., Pirard, B., Schimanski, S., Kirsch, R., Habermann, J., Klingler, O., Schlotte, V., Weithmann, K. U., & Wendt, K. U. (2005). Structural basis for the highly selective inhibition of M M P-13. Chemistry & Biology, 12, 181-189. Erlichman, C., Adjei, A., Alberts, S. R., Sloan, J. A., Goldberg, R., Pitot, H. C., Rubin, J., Atherton, P., Klee, G., &
Journal Pre-proof Humphrey, R. (2001). Phase I study of the matrix metal loproteinase inhibitor, BAY 12–9566. Annals of Oncology, 12, 389-395. Fanjul-Fernández, M ., Folgueras, A. R., Cabrera, S., & López -Otín, C. (2010). M atrix metalloproteinases: evolution, gene regulation and functional analysis in mouse models. Biochimica et Biophysica Acta (BBA)-Molecular Cell Research, 1803, 3-19. Fields, G.B. (2019a). The rebirth of matrix metalloproteinase inhibitors: moving beyond the dogma. Cells, 8, 984. Fields, G.B. (2019b). M echanisms of Action of Novel Drugs Targeting Angiogenesis-Promoting M atrix M etalloproteinases. Frontiers in Immunology, 10,1278.
oo
f
Fingleton, B. (2006). M atrix metalloproteinases: roles in cancer and metastasis. Frontiers in Bioscience, 11, 91. Fischer, T., & Riedl, R. (2019). Inhibitory antibodies designed for matrix metalloproteinase modulation. Molecules, 24,
pr
2265.
Chemistry-A European Journal, 25, 7960-7980.
e-
Fischer, T., Senn, N., & Riedl, R. (2019). Design and Structural Evolution of M atrix M etalloproteinase Inhibitors.
Pr
Frankwich, K., Tibble, C., Torres-Gonzalez, M ., Bonner, M ., Lefkowitz, R., Tyndall, M ., Schmid-Schönbein, G. W., Villarreal, F., Heller, M ., & Herbst, K. (2012). Proof of concept: matrix metalloproteinase inhibitor decreases
al
inflammation and improves muscle insulin sensitivity in people with type 2 diabetes. Journal of Inflammation, 9,
rn
35.
Gebauer, M ., & Skerra, A. (2009). Engineered protein scaffolds as next-generation antibody therapeutics. Current Opinion in
Jo u
Chemical Biology, 13, 245-255.
Gege, C., Bao, B., Bluhm, H., Boer, J. r., Gallagher Jr, B. M ., Korniski, B., Powers, T. S., Steeneck, C., Taveras, A. G., & Baragi, V. M . (2012). Discovery and evaluation of a non-Zn chelating, selective matrix metalloproteinase 13 (MMP-13) inhibitor for potential intra-articular treatment of osteoarthritis. Journal of Medicinal Chemistry, 55, 709-716. Gentile, E., & Liuzzi, G. M . (2017). M arine pharmacology: therapeutic targeting of matrix metalloproteinases in neuroinflammation. Drug Discovery Today, 22, 299-313. Giavazzi, R., Garofalo, A., Ferri, C., Lucchini, V., Bone, E. A., Chiari, S., Brown, P. D., Nicoletti, M . I., & Taraboletti, G. (1998). Batimastat, a synthetic inhibitor of matrix metalloproteinases, potentiates the antitumor activity of cisplatin in ovarian carcinoma xenografts. Clinical Cancer Research, 4, 985-992. Golub, L. M ., M cNamara, T. F., Ryan, M . E., Kohut, B., Blieden, T., Payonk, G., Sipos, T., & Baron, H. J. (2001). Adjunctive treatment with subantimicrobial doses of doxycycline: effects on gingival fluid collagenase activity and
Journal Pre-proof attachment loss in adult periodontitis. Journal of Clinical Periodontology, 28, 146-156. Gossage, D. L., Cieslarová, B., Ap, S., Zheng, H., Xin, Y., Lal, P., Chen, G., Smith, V., & Sundy, J. S. (2018). Phase 1b Study of the Safety, Pharmacokinetics, and Disease-related Outcomes of the M atrix M etalloproteinase-9 Inhibitor Andecaliximab in Patients With Rheumatoid Arthritis. Clinical Therapeutics, 40, 156-165. Gross, J. (2004). How tadpoles lose their tails: path to discovery of the first matrix metalloproteinase. Matrix Biology, 1, 3-13. Gross, J., & Lapiere, C. M . (1962). Collagenolytic activity in amphibian tissues: a tissue culture assay. Proceedings of the National Academy of Sciences of the United States of America, 48, 1014.
oo
f
Grossman, M ., Tworowski, D., Dym, O., Lee, M .-H., Levy, Y., M urphy, G., & Sagi, I. (2010). The intrinsic protein flexibility of endogenous protease inhibitor TIM P-1 controls its binding interface and affects its function. Biochemistry, 49,
pr
6184-6192.
e-
Guarnera, E., & Berezovsky, I. N. (2016). Allosteric sites: remote control in regulation of protein activity. Current Opinion in Structural Biology, 37, 1-8.
Pr
Hande, K. R., Collier, M ., Paradiso, L., Stuart-Smith, J., Dixon, M ., Clendeninn, N., Yeun, G., Alberti, D., Binger, K., & Wilding, G. (2004). Phase I and pharmacokinetic study of prinomastat, a matrix metalloprotease inhibitor. Clinical
al
Cancer Research, 10, 909-915.
rn
Heath, E. I., Burtness, B. A., Lawrence, K., Salem, R. R., Yang, S. C., Heitmiller, R. F., Canto, M . I., Knisely, J. P. S., M ark, T., & Elizabeth, M . (2006). Phase II, parallel-design study of preoperative combined modality therapy and the
Jo u
matrix metalloprotease (mmp) inhibitor prinomastat in patients with esophageal adenocarcinoma. Investigational New Drugs, 24, 135.
Hirte, H., Vergote, I., Jeffrey, J., Grimshaw, R., Coppieters, S., Schwartz, B., Tu, D., Sadura, A., Brundage, M ., & Seymour, L. (2006). A phase III randomized trial of BAY 12-9566 (tanomastat) as maintenance therapy in patients with advanced ovarian cancer responsive to primary surgery and paclitaxel/platinum containing chemotherapy: a National Cancer Institute of Canada Clinical Trials Group Study. Gynecologic Oncology, 102, 300-308. Jacobsen, J. A., Jourden, J. L. M ., M iller, M . T., & Cohen, S. M . (2010). To bind zinc or not to bind zinc: an examination of innovative approaches to improved metalloproteinase inhibition. Biochimica et Biophysica Acta (BBA)-Molecular Cell Research, 1803, 72-94. Jozic, D., Bourenkov, G., Lim, N.-H., Visse, R., Nagase, H., Bode, W., & M askos, K. (2005). X-ray Structure of Human proMMP-1 new insights into procollagenase activation and collagen binding. Journal of Biological Chemistry, 280, 9578-9585.
Journal Pre-proof Kelley, L. C., Chi, Q., Cáceres, R., Hastie, E., Schindler, A. J., Jiang, Y., M atus, D. Q., Plastino, J., & Sherwood, D. R. (2019). Adaptive F-actin polymerization and localized ATP production drive basement membrane invasion in the absence of M M Ps. Developmental Cell, 48, 313-328. e318. Kessenbrock, K., Wang, C.-Y., & Werb, Z. (2015). M atrix metalloproteinases in stem cell regulat ion and cancer. Matrix Biology, 44, 184-190. Latreille, J., Batist, G., Laberge, F., Champagne, P., Croteau, D., Falardeau, P., Levinton, C., Hariton, C., Evans, W. K., & Dupont, E. (2003). Phase I/II trial of the safety and efficacy of AE-941 (Neovastat®) in the treatment of non– small-cell lung cancer. Clinical Lung Cancer, 4, 231-236.
oo
f
Lauer-Fields, J. L., Whitehead, J. K., Li, S., Hammer, R. P., Brew, K., & Fields, G. B. (2008). Selective modulation of matrix metalloproteinase 9 (M M P-9) functions via exosite inhibition. Journal of Biological Chemistry, 283, 20087-20095.
pr
Leighl, N. B., Paz-Ares, L., Douillard, J.-Y., Peschel, C., Arnold, A., Depierre, A., Santoro, A., Betticher, D. C., Gatzemeier,
e-
U., & Jassem, J. (2005). Randomized phase III study of matrix metalloproteinase inhibitor BM S-275291 in combination with paclitaxel and carboplatin in advanced non-small-cell lung cancer: National Cancer Institute of
Pr
Canada-Clinical Trials Group Study BR. 18. Journal of Clinical Oncology, 23, 2831-2839. Levin, M ., Amar, D., & Aharoni, A. (2013). Employing directed evolution for the functional analysis of multi-specific
al
proteins. Bioorganic & Medicinal Chemistry, 21, 3511-3516.
rn
Levin, M ., Udi, Y., Solomonov, I., & Sagi, I. (2017). Next generation matrix metalloproteinase inhibitors—Novel strategies bring new prospects. Biochimica et Biophysica Acta (BBA)-Molecular Cell Research, 1864, 1927-1939.
Jo u
Levitt, N. C., Eskens, F. A., O'Byrne, K. J., Propper, D. J., Denis, L. J., Owen, S. J., Choi, L., ., Foekens, J. A., Wilner, S., ., & Wood, J. M . (2001). Phase I and pharmacological study of the oral matrix metalloproteinase inhibitor, MM I270 (CGS27023A), in patients with advanced solid cancer. Clinical Cancer Research, 7, 1912. Litman, G. W., Rast, J. P., Shamblott, M . J., Haire, R. N., Hulst, M ., Roess, W., Litman, R. T., Hinds-Frey, K. R., Zilch, A., & Amemiya, C. T. (1993). Phylogenetic diversification of immunoglobulin genes and the antibody repertoire. Molecular Biology and Evolution, 10, 60-72. Lu, C., Lee, J. J., Komaki, R., Herbst, R. S., Feng, L., Evans, W. K., Choy, H., Desjardins, P., Esparaz, B. T., & Truong, M . T. (2010). chemoradiotherapy With or Without Ae-941 in Stage iii Non–Small cell lung cancer: A randomized Phase iii trial. Journal of the National Cancer Institute, 102, 859-865. Lu, S., Li, S., & Zhang, J. (2014). Harnessing allostery: a novel approach to drug discovery. Medicinal Research Reviews, 34, 1242-1285. M agnussen, H., Watz, H., Kirsten, A., Wang, M ., Wray, H., Samuelsson, V., M o, J., & Kay, R. (2011). Safety and tolerability
Journal Pre-proof of an oral MM P-9 and-12 inhibitor, AZD1236, in patients with moderate-to-severe COPD: a randomised controlled 6-week trial. Pulmonary Pharmacology & Therapeutics, 24, 563-570. M arshall, D. C., Lyman, S. K., M cCauley, S., Kovalenko, M ., Spangler, R., Liu, C., Lee, M ., O’Sullivan, C., Barry -Hamilton, V., & Ghermazien, H. (2015). Selective allosteric inhibition of MM P9 is efficacious in preclinical models of ulcerative colitis and colorectal cancer. PloS One, 10, e0127063. M artens, E., Leyssen, A., Van Aelst, I., Fiten, P., Piccard, H., Hu, J., Descamps, F. J., Van den Steen, P. E., Proost, P., & Van Damme, J. (2007). A monoclonal antibody inhibits gelatinase B/MM P-9 by selective binding to part of the catalytic domain and not to the fibronectin or zinc binding domains. Biochimica et Biophysica Acta (BBA)-General
oo
f
Subjects, 1770, 178-186.
M askos, K., & Bode, W. (2003). Structural basis of matrix metalloproteinases and tissue inhibit ors of metalloproteinases.
pr
Molecular Biotechnology, 25, 241-266.
Drug Discovery & Development, 7, 513-535.
e-
M atter, H., & Schudok, M . (2004). Recent advances in the design of matrix metalloprotease inhibitors. Current Opinion in
Pr
M iller, K. D., Saphner, T. J., Waterhouse, D. M ., Chen, T.-T., Rush-Taylor, A., Sparano, J. A., Wolff, A. C., Cobleigh, M . A., Galbraith, S., & Sledge, G. W. (2004). A randomized phase II feasibility trial of BM S-275291 in patients with early
al
stage breast cancer. Clinical Cancer Research, 10, 1971-1975.
rn
M inagar, A., Alexander, J. S., Schwendimann, R. N., Kelley, R. E., Gonzalez -Toledo, E., Jimenez, J. J., M auro, L., Jy, W., & Smith, S. J. (2008). Combination therapy with interferon beta-1a and doxycycline in multiple sclerosis: an
Jo u
open-label trial. Archives of Neurology, 65, 199-204. M itternacht, S., & Berezovsky, I. N. (2011). Binding leverage as a molecular basis for allosteric regulation. PLoS Computational Biology, 7, e1002148. M ohan, V., Talmi-Frank, D., Arkadash, V., Papo, N., & Sagi, I. (2016). M atrix metalloproteinase protein inhibitors: highlighting a new beginning for metalloproteinases in medicine. Metalloproteinases in Medicine, 3, 31. M oore, M . J., Hamm, J., Dancey, J., Eisenberg, P., Dagenais, M ., Fields, A., Hagan, K., Greenberg, B., Colwell, B., & Zee, B. (2003). Comparison of gemcitabine versus the matrix metalloproteinase inhibitor BAY 12-9566 in patients with advanced or metastatic adenocarcinoma of the pancreas: a phase III trial of the National Cancer Institute of Canada Clinical Trials Group. Journal of Clinical Oncology, 21, 3296-3302. M urphy, G., Houbrechts, A., Cockett, M. I., Williamson, R. A., O'Shea, M ., & Docherty, A. J. (1991). The N-terminal domain of tissue inhibitor of metalloproteinases retains metalloproteinase inhibitory activity. Biochemistry, 30, 8097-8102.
Journal Pre-proof Nagase, H., & Brew, K. (2002). Engineering of tissue inhibitor of metalloproteinases mutants as potential therapeutics. Arthritis Research & Therapy, 4, S51. Nam, D. H., Rodriguez, C., Remacle, A. G., Strongin, A. Y., & Ge, X. (2016). Active-site MMP-selective antibody inhibitors discovered from convex paratope synthetic libraries. Proceedings of the National Academy of Sciences, 113, 14970-14975. Nemunaitis, J., Poole, C., Primrose, J., Rosemurgy, A., M alfetano, J., Brown, P., Berrington, A., Cornish, A., Lynch, K., & Rasmussen, H. (1998). Combined analysis of studies of the effects of the matrix metalloproteinase inhibitor marimastat on serum tumor markers in advanced cancer: selection of a biologically active and tolerable dose for
oo
f
longer-term studies. Clinical Cancer Research, 4, 1101-1109.
Overall, C. M . (2002). M olecular determinants of metalloproteinase substrate specificity. Molecular Biotechnology, 22,
pr
51-86.
e-
Overall, C. M ., & Butler, G. S. (2007). Protease yoga: extreme flexibility of a matrix metalloproteinase. Structure, 15, 1159-1161.
British Journal of Cancer, 94, 941.
Pr
Overall, C. M ., & Kleifeld, O. (2006a). Towards third generation matrix metalloproteinase inhibitors for cancer therapy.
al
Overall, C. M ., & Kleifeld, O. (2006b). Validating matrix metalloproteinases as drug targets and anti-targets for cancer
rn
therapy. Nature Reviews Cancer, 6, 227.
Overall, C. M ., & López-Otín, C. (2002). Strategies for MMP inhibition in cancer: innovations for the post -trial era. Nature
Jo u
Reviews Cancer, 2, 657.
Paemen, L., M artens, E., M asure, S., & Opdenakker, G. (1995). M onoclonal antibodies specific for natural human neutrophil gelatinase B used for affinity purification, quantitation by two ‐ site ELISA and inhibition of enzymatic activity. European Journal of Biochemistry, 234, 759-765. Park, H. I., Jin, Y., Hurst, D. R., M onroe, C. A., Lee, S., Schwartz, M . A., & Sang, Q.-X. A. (2003). The intermediate S1′ pocket of the endometase/matrilysin-2 active site revealed by enzyme inhibition kinetic studies, protein sequence analyses, and homology modeling. Journal of Biological Chemistry, 278, 51646-51653. Peng, S. X., Strojnowski, M . J., Hu, J. K., Smith, B. J., Eichhold, T. H., Wehmeyer, K. R., Pikul, S., & Almstead, N. G. (1999). Gas chromatographic–mass spectrometric analysis of hydroxylamine for monitoring the metabolic hydrolysis of metalloprotease inhibitors in rat and human liver microsomes. Journal of Chromatography B: Biomedical Sciences and Applications, 724, 181-187. Peress, N., Perillo, E., & Zucker, S. (1995). Localization of tissue inhibitor of matrix metalloproteinases in Alzheimer's
Journal Pre-proof disease and normal brain. Journal of Neuropathology & Experimental Neurology, 54, 16-22. Pirard, B. (2007). Insight into the structural determinants for selective inhibition of matrix metalloproteinases. Drug Discovery Today, 12, 640-646. Pochetti, G., M ontanari, R., Gege, C., Chevrier, C., Taveras, A. G., & M azza, F. (2009). Extra binding region induced by non-zinc chelating inhibitors into the S1′ subsite of matrix metalloproteinase 8 (MM P -8). Journal of Medicinal Chemistry, 52, 1040-1049. Preshaw, P. M., Hefti, A. F., Novak, M. J., M ichalowicz, B. S., Pihlstrom, B. L., Schoor, R., Trummel, C. L., Dean, J., Van Dyke, T. E., & Walker, C. B. (2004). Subantimicrobial dose doxycycline enhances the efficacy of scaling and root
oo
f
planing in chronic periodontitis: a multicenter trial. Journal of Periodontology, 75, 1068-1076. Purcell, B. P., Lobb, D., Charati, M . B., Dorsey, S. M ., Wade, R. J., Zellars, K. N., Doviak, H., Pettaway, S., Logdon, C. B.,
pr
& Shuman, J. A. (2014). Injectable and bioresponsive hydrogels for on-demand matrix metalloproteinase
e-
inhibition. Nature Materials, 13, 653.
Rao, B. G. (2005). Recent developments in the design of specific matrix metalloproteinase inhibitors aided by structural and
Pr
computational studies. Current Pharmaceutical Design, 11, 295-322. Reich, R., Thompson, E. W., Iwamoto, Y., M artin, G. R., Deason, J. R., Fuller, G. C., & M iskin, R. (1988). Effects of
al
inhibitors of plasminogen activator, serine proteinases, and collagenase IV on the invasion of basement membranes
rn
by metastatic cells. Cancer Research, 48, 3307-3312. Remacle, A. G., Golubkov, V. S., Shiryaev, S. A., Dahl, R., Stebbins, J. L., Chernov, A. V., Cheltsov, A. V., Pellecchia, M ., &
Jo u
Strongin, A. Y. (2012). Novel M T1-MMP small-molecule inhibitors based on insights into hemopexin domain function in tumor growth. Cancer Research, 72, 2339-2349. Rigas, J. R., Denham, C. A., Rinaldi, D., M oore, T., Smith, J. W., Winston, R. D., & Humphrey, R. (2003). O -107 adjuvant targeted therapy in unresectable lung cancer: the results of two randomized placebo-controlled trials of bay 12-9566, a matrix metalloproteinase inhibitor (M M PI). Lung Cancer, 41, S34. Rizvi, N. A., Humphrey, J. S., Ness, E. A., Johnson, M . D., Gupta, E., Williams, K., Daly, D. J., Sonnichsen, D., Conway, D., & M arshall, J. (2004). A phase I study of oral BM S-275291, a novel nonhydroxamate sheddase-sparing matrix metalloproteinase inhibitor, in patients with advanced or metastatic cancer. Clinical Cancer Research, 10, 1963-1970. Rouanet-M ehouas, C., Czarny, B., Beau, F., Cassar-Lajeunesse, E., Stura, E. A., Dive, V., & Devel, L. (2016). Zinc– M etalloproteinase Inhibitors: Evaluation of the Complex Role Played by the Zinc-Binding Group on Potency and Selectivity. Journal of Medicinal Chemistry, 60, 403-414.
Journal Pre-proof Rudek, M . A., Figg, W. D., Dyer, V., Dahut, W., Turner, M . L., Steinberg, S. M ., Liewehr, D. J., Kohler, D. R., Pluda, J. M ., & Reed, E. (2001). Phase I clinical trial of oral COL-3, a matrix metalloproteinase inhibitor, in patients with refractory metastatic cancer. Journal of Clinical Oncology, 19, 584-592. Sandborn, W., Bhandari, B., Fogel, R., Onken, J., Yen, E., Zhao, X., Jiang, Z., Ge, D., Xin, Y., & Ye, Z. (2016). Randomised clinical trial: a phase 1, dose‐ ranging study of the anti‐ matrix metalloproteinase‐ 9 monoclonal antibody GS‐ 5745 versus placebo for ulcerative colitis. Alimentary Pharmacology & Therapeutics, 44, 157-169. Sandborn, W. J., Bhandari, B. R., Randall, C., Younes, Z. H., Romanczyk, T., Xin, Y., Wendt, E., Chai, H., M ckevitt, M ., & Zhao, S. (2018). Andecaliximab [Anti-matrix M etalloproteinase-9] Induction Therapy for Ulcerative Colitis: A
oo
f
Randomised, Double-Blind, Placebo-Controlled, Phase 2/3 Study in Patients With M oderate to Severe Disease. Journal of Crohn’s and Colitis, 12, 1021-1029.
pr
Scannevin, R. H., Alexander, R., Haarlander, T. M ., Burke, S. L., Singer, M ., Huo, C., Zhang, Y.-M ., M aguire, D., Spurlino,
e-
J., & Deckman, I. (2017). Discovery of a highly selective chemical inhibitor of matrix metalloproteinase-9 (M M P-9) that allosterically inhibits zymogen activation. Journal of Biological Chemistry, 292, 17963-17974.
Pr
Schmalfeldt, B., Prechtel, D., Härting, K., Späthe, K., Rutke, S., Konik, E., Fridman, R., Berger, U., Schmitt, M ., & Kuhn, W. (2001). Increased expression of matrix metalloproteinases (MMP)-2, MMP-9, and the urokinase-type plasminogen
al
activator is associated with progression from benign to advanced ovarian cancer. Clinical Cancer Research, 7,
rn
2396-2404.
Schreiber, S., Siegel, C. A., Friedenberg, K. A., Younes, Z. H., Seidler, U., Bhandari, B. R., Wang, K., Wendt, E., M ckevitt,
Jo u
M ., & Zhao, S. (2018). A phase 2, Randomized, Placebo-controlled Study Evaluating M atrix M etalloproteinase-9 Inhibitor, Andecaliximab, in Patients with M oderately to Severely Active Crohn's Disease. Journal of Crohn’s and Colitis, 12. 1014-1020
Sela-Passwell, N., Rosenblum, G., Shoham, T., & Sagi, I. (2010). Structural and functional bases for allosteric control of MMP activities: can it pave the path for selective inhibition? Biochimica et Biophysica Acta (BBA)-Molecular Cell Research, 1803, 29-38. Shah, M . A., Starodub, A., Sharma, S., Berlin, J., Patel, M ., Wainberg, Z. A., Chaves, J., Gordon, M ., Windsor, K., & Brachmann, C. B. (2018). Andecaliximab/GS-5745 alone and combined with mFOLFOX6 in advanced gastric and gastroesophageal junction adenocarcinoma: results from a phase I study. Clinical Cancer Research, 24, 3829-3837. Sharabi, O., Shirian, J., Grossman, M ., Lebendiker, M ., Sagi, I., & Shifman, J. (2014). Affinity -and specificity-enhancing mutations are frequent in multispecific interactions between TIM P2 and M M Ps. PloS One, 9, e93712. Shiryaev, S., Remacle, A., Golubkov, V., Ingvarsen, S., Porse, A., Behrendt, N., Cieplak, P., & Strongin, A. (2013). A
Journal Pre-proof monoclonal antibody interferes with TIM P-2 binding and incapacitates the MMP-2-activating function of multifunctional, pro-tumorigenic M M P-14/MT1–MMP. Oncogenesis, 2, e80. Sircar, A., Sanni, K. A., Shi, J., & Gray, J. J. (2011). Analysis and modeling of the variable region of camelid single-domain antibodies. The Journal of Immunology, 186, 6357-6367. Skiles, J.W., Gonnella, N. C., & Jeng, A. Y. (2001). The design, structure, and therapeutic application of matrix metalloproteinase inhibitors. Current Medicinal Chemistry, 8, 425-474. Skiles, J. W., Gonnella, N. C., & Jeng, A. Y. (2004). The design, structure, and clinical update of small molecular weight matrix metalloproteinase inhibitors. Current Medicinal Chemistry, 11, 2911-2977.
oo
f
Sparano, J. A., Patricia, B., Patricia, S., Gradishar, W. J., Ingle, J. N., Stanley, Z., & Davidson, N. E. (2004). Randomized phase III trial of marimastat versus placebo in patients with metastatic breast cancer who have responding or stable
pr
disease after first-line chemotherapy: Eastern Cooperative Oncology Group trial E2196. Journal of Clinical
e-
Oncology, 22, 4683-4690.
Tallant, C., M arrero, A., & Gomis-Rüth, F. X. (2010). M atrix metalloproteinases: fold and function of their catalytic domains.
Pr
Biochimica et Biophysica Acta (BBA)-Molecular Cell Research, 1803, 20-28. Tam, E. M . , M oore, T. R. , Butler, G. S. , & Overall, C. M . . (2004). Characterization of the distinct collagen binding,
al
helicase and cleavage mechanisms of matrix metalloproteinase 2 and 14 ( gelatinase a and mt1-mmp): the
rn
differential roles of the mmp hemopexin c domains and the mmp -2 fibronectin type ii modules in collagen triple helicase activities. Journal of Biological Chemistry, 279, 43336-43344.
Jo u
Terp, G. E., Cruciani, G., Christensen, I. T., & Jørgensen, F. S. (2002). Structural differences of matrix metalloproteinases with potential implications for inhibitor selectivity examined by the GRID/CPCA approach. Journal of Medicinal Chemistry, 45, 2675-2684.
Udi, Y., Fragai, M ., Grossman, M ., M itternacht, S., Arad-Yellin, R., Calderone, V., M elikian, M ., Toccafondi, M ., Berezovsky, I. N., & Luchinat, C. (2013). Unraveling hidden regulatory sites in structurally homologous metalloproteases. Journal of Molecular Biology, 425, 2330-2346. Udi, Y., Grossman, M ., Solomonov, I., Dym, O., Rozenberg, H., M oreno, V., Cuniasse, P., Dive, V., Arroyo, A. G., & Sagi, I. (2015). Inhibition mechanism of membrane metalloprotease by an exosite-swiveling conformational antibody. Structure, 23, 104-115. Van M arle, S., Van Vliet, A., Sollie, F., Kambayashi, Y., & Yamada-Sawada, T. (2005). Safety, tolerability and pharmacokinetics of oral S-3304, a novel matrix metalloproteinase inhibitor, in single and multiple dose escalation studies in healthy volunteers. International Journal of Clinical Pharmacology & Therapeutics, 43. 282-293.
Journal Pre-proof Vandenbroucke, R. E., & Libert, C. (2014). Is there new hope for therapeutic matrix metalloproteinase inhibition? Nature Reviews Drug Discovery, 13, 904. Vanlaere, I., & Libert, C. (2009). M atrix metalloproteinases as drug targets in infections caused by gram-negative bacteria and in septic shock. Clinical Microbiology Reviews, 22, 224-239. Visse, R., & Nagase, H. (2003). M atrix metalloproteinases and tissue inhibitors of metalloproteinases: structure, function, and biochemistry. Circulation Research, 92, 827-839. Wang, L., Li, X., Zhang, S., Lu, W., Liao, S., Liu, X., Shan, L., Shen, X., Jiang, H., & Zhang, W. (2012). Natural products as a gold mine for selective matrix metalloproteinases inhibitors. Bioorganic & Medicinal Chemistry, 20, 4164-4171.
oo
f
Wei, S., Chen, Y., Chung, L., Nagase, H., & Brew, K. (2003). Protein engineering of TIM P-1 inhibitory domain: In search of selective M M P inhibitors. Journal of Biological Chemistry, 278, 9831-9834.
pr
Whittaker, M ., Floyd, C. D., Brown, P., & Gearing, A. J. (1999). Design and therapeutic application of matrix
e-
metalloproteinase inhibitors. Chemical Reviews, 99, 2735-2776.
Winer, A., Adams, S., & M ignatti, P. (2018). M atrix metalloproteinase inhibitors in cancer therapy: turning past failures into
Pr
future successes. Molecular Cancer Therapeutics, 17, 1147-1155. Wojtowicz-Praga, S., Low, J., M arshall, J., Ness, E., Dickson, R., Barter, J., Sale, M ., M cCann, P., M oore, J., & Cole, A.
al
(1996). Phase I trial of a novel matrix metalloproteinase inhibitor batimastat (BB-94) in patients with advanced
rn
cancer. Investigational New Drugs, 14, 193-202.
Woolley, D. E., Roberts, D. R., & Evanson, J. M . (1975). Inhibition of human collagenase activity by a small molecular
Jo u
weight serum protein. Biochemical and Biophysical Research Communications, 66, 747-754. Wynn, R. (1999). Latest FDA approvals for dentistry. General Dentistry, 47, 19. Xu, X., Chen, Z., Wang, Y., Bonewald, L., & Steffensen, B. (2007). Inhibition of MMP-2 gelatinolysis by targeting exodomain–substrate interactions. Biochemical Journal, 406, 147-155. Young, D.A., Barter, M .J., & Wilkinson, D.J. (2019) Recent advances in understanding the regulation of metalloproteinases. F1000Research, 8,195 Yao, Q., Kou, L., Tu, Y., & Zhu, L. (2018). M MP-responsive ‘smart’drug delivery and tumor targeting. Trends in Pharmacological Sciences, 39, 766-781. Zucker, S., Cao, J., & Chen, W.-T. (2000). Critical appraisal of the use of matrix metalloproteinase inhibitors in cancer treatment. Oncogene, 19, 6642. Zucker, S., Lysik, R. M., Zarrabi, M . H., Stetler-Stevenson, W., Liotta, L. A., Birkedal-Hansen, H., M ann, W., & Furie, M . (1992). Type IV collagenase/gelatinase (MMP-2) is not increased in plasma of patients with cancer. Cancer
Journal Pre-proof
Jo u
rn
al
Pr
e-
pr
oo
f
Epidemiology and Prevention Biomarkers, 1, 475-479.