ArticlePDF Available

Recent Advances in the Development and Application of Radiolabeled Kinase Inhibitors for PET Imaging

Authors:

Abstract and Figures

Over the last 20 years, intensive investigation and multiple clinical successes targeting protein kinases, mostly for cancer treatment, have identified small molecule kinase inhibitors as a prominent therapeutic class. In the course of those investigations, radiolabeled kinase inhibitors for positron emission tomography (PET) imaging have been synthesized and evaluated as diagnostic imaging probes for cancer characterization. Given that inhibitor coverage of the kinome is continuously expanding, in vivo PET imaging will likely find increasing applications for therapy monitoring and receptor density studies both in- and outside of oncological conditions. Early investigated radiolabeled inhibitors, which are mostly based on clinically approved tyrosine kinase inhibitor (TKI) isotopologues, have now entered clinical trials. Novel radioligands for cancer and PET neuroimaging originating from novel but relevant target kinases are currently being explored in preclinical studies. This article reviews the literature involving radiotracer design, radiochemistry approaches, biological tracer evaluation and nuclear imaging results of radiolabeled kinase inhibitors for PET reported between 2010 and mid-2015. Aspects regarding the usefulness of pursuing selective vs. promiscuous inhibitor scaffolds and the inherent challenges associated with intracellular enzyme imaging will be discussed.
Content may be subject to copyright.
Review
Recent Advances in the Development and
Application of Radiolabeled Kinase Inhibitors for
PET Imaging
Vadim Bernard-Gauthier *,† , Justin J. Bailey , Sheldon Berke and Ralf Schirrmacher *
Received: 22 October 2015 ; Accepted: 1 December 2015 ; Published: 9 December 2015
Academic Editor: James W. Leahy
Division of Oncological Imaging, Department of Oncology, University of Alberta, 11560 University Ave.,
Edmonton, AB T6G 1Z2, Canada; jjbailey@ualberta.ca (J.J.B.); sberke@ualberta.ca (S.B.)
*Correspondence: bernardg@ualberta.ca (V.B.-G.); schirrma@ualberta.ca (R.S.); Tel.: +1-780-248-1829 (R.S.)
These authors contributed equally to this work.
Abstract: Over the last 20 years, intensive investigation and multiple clinical successes targeting
protein kinases, mostly for cancer treatment, have identified small molecule kinase inhibitors
as a prominent therapeutic class. In the course of those investigations, radiolabeled kinase
inhibitors for positron emission tomography (PET) imaging have been synthesized and evaluated
as diagnostic imaging probes for cancer characterization. Given that inhibitor coverage of the
kinome is continuously expanding, in vivo PET imaging will likely find increasing applications for
therapy monitoring and receptor density studies both in- and outside of oncological conditions.
Early investigated radiolabeled inhibitors, which are mostly based on clinically approved tyrosine
kinase inhibitor (TKI) isotopologues, have now entered clinical trials. Novel radioligands for
cancer and PET neuroimaging originating from novel but relevant target kinases are currently being
explored in preclinical studies. This article reviews the literature involving radiotracer design,
radiochemistry approaches, biological tracer evaluation and nuclear imaging results of radiolabeled
kinase inhibitors for PET reported between 2010 and mid-2015. Aspects regarding the usefulness of
pursuing selective vs. promiscuous inhibitor scaffolds and the inherent challenges associated with
intracellular enzyme imaging will be discussed.
Keywords: positron emission tomography; tyrosine kinase inhibitors; protein kinases; nuclear
imaging; cancer imaging; neuroimaging; fluorine-18; carbon-11
1. Introduction
Through mediation of crucial signal transduction pathways, the catalytic activity of kinases lies
at the cornerstone of a myriad of cellular functions and regulates various cellular processes including
differentiation, apoptosis, proliferation and metabolism. The last quarter century has witnessed
fundamental discoveries and prominent advances in protein and lipid kinase function and inhibition.
The FDA approval of the Bcr-Abl inhibitor imatinib in 2001 and subsequent clinical successes
in the treatment of chronic myeloid leukemia (CML) marked the emergence and rapid growth
of small molecule kinase inhibitors as a novel therapeutic instrument for cancer treatment [1,2].
As of July 2015, a total of 29 small molecule kinase inhibitors, mostly aimed at tyrosine kinases
for the management of neoplastic diseases, have been approved [3,4]. It is remarkable that 20
of those approvals were completed from 2011 onward. Undoubtedly, kinase inhibitors currently
constitute one of the most keenly investigated therapeutic classes. Yet, approved kinase inhibitors
converge around very narrow pharmacophores and targets, often derived from previously successful
inhibitors, and represent only a minute fraction of the structural diversity globally encountered
Molecules 2015,20, 22000–22027; doi:10.3390/molecules201219816 www.mdpi.com/journal/molecules
Molecules 2015,20, 22000–22027
within kinase inhibitors characterized preclinically [1,5]. To date, approved inhibitors have largely
left untouched many potential promising applications in areas including the central nervous system
(CNS), inflammatory and cardiovascular diseases. Current kinome coverage data from publicly
available kinase inhibitors indicate that about half of the 518 human kinases have been targeted,
with inhibitors well distributed among the seven main protein kinase groups (AGC, CAMK, CK1,
CMGC, STE, TK and TKL) [6,7]. Despite this great progress, the human kinome remains mostly
unmapped terrain, providing ample emerging possibilities for drug development in the field of
kinase inhibitors [8].
Regardless of sequence differences, protein kinases share common tertiary structure components
which include a highly conserved ATP-binding site located at the interface of N- and C-lobes.
This binding cleft constitutes the primary anchorage site of most inhibitors (Figure 1a–d) [9,10].
Within the hinge region, a “gatekeeper” residue regulates access to an adjacent hydrophobic pocket.
Additional structural features include a glycine-rich loop and a flexible activation loop within the
C-lobe initiated by a conserved Asp-Phe-Gly (DFG) motif which dictates the accessibility of the
catalytic site. Inhibitors either interact reversibly or irreversibly with kinases. Irreversible inhibitors
which bear a Michael acceptor fragment covalently react with a cysteine residue in the environment
of the ATP-binding site (Figure 1b) [11,12]. Otherwise, most inhibitors exert reversible binding
and are organized in distinct groups according to their binding mode and DFG orientation. Type-I
inhibitors engage in ATP-competitive interaction typically binding at the hinge region, with the Asp
residue from the DFG motif directed towards the ATP binding site (Figure 1a; DFG-in inhibitors).
Type-II inhibitors bind the inactive kinase conformation in which the DFG motif is rotated and allows
interaction with an additional allosteric pocket while maintaining contact with the hinge residues
(Figure 1c; DFG-out inhibitors).
Molecules 2015, 20, page–page
2
applications in areas including the central nervous system (CNS), inflammatory and cardiovascular
diseases. Current kinome coverage data from publicly available kinase inhibitors indicate that about
half of the 518 human kinases have been targeted, with inhibitors well distributed among the seven
main protein kinase groups (AGC, CAMK, CK1, CMGC, STE, TK and TKL) [6,7]. Despite this great
progress, the human kinome remains mostly unmapped terrain, providing ample emerging possibilities
for drug development in the field of kinase inhibitors [8].
Regardless of sequence differences, protein kinases share common tertiary structure components
which include a highly conserved ATP-binding site located at the interface of N- and C-lobes. This
binding cleft constitutes the primary anchorage site of most inhibitors (Figure 1a–d) [9,10]. Within the
hinge region, a “gatekeeper” residue regulates access to an adjacent hydrophobic pocket. Additional
structural features include a glycine-rich loop and a flexible activation loop within the C-lobe initiated
by a conserved Asp-Phe-Gly (DFG) motif which dictates the accessibility of the catalytic site.
Inhibitors either interact reversibly or irreversibly with kinases. Irreversible inhibitors which bear a
Michael acceptor fragment covalently react with a cysteine residue in the environment of the
ATP-binding site (Figure 1b) [11,12]. Otherwise, most inhibitors exert reversible binding and are
organized in distinct groups according to their binding mode and DFG orientation. Type-I inhibitors
engage in ATP-competitive interaction typically binding at the hinge region, with the Asp residue
from the DFG motif directed towards the ATP binding site (Figure 1a; DFG-in inhibitors). Type-II
inhibitors bind the inactive kinase conformation in which the DFG motif is rotated and allows
interaction with an additional allosteric pocket while maintaining contact with the hinge residues
(Figure 1c; DFG-out inhibitors).
Figure 1. Representative examples of the primary types of small molecule kinase inhibitors including
a DFG-in irreversible inhibitor. Kinases are shown in grey, the activation loops in burgundy, the
glycine-rich loops in blue and the DFG motif in red (sticks). The carbon atoms of the inhibitors and
the ATP are respectively represented in cyan and green (ac); The chemical structures and schematic
binding modes of the inhibitors are depicted. For inhibitors that have been radiolabeled for P ET imaging,
the position of the radionuclide (carbon-11 and fluorine-18) is indicated with a black arrow (eg); (a,e)
Type I inhibitor; co-crystal structure of erlotinib bound to epidermal growth factor receptor (EGFR)
(DFG-in) (PDB ID: 1M17); (b,f) Type I inhibitor (irreversible); co-crystal structure of Afatinib bound to
EGFR T790M (DFG-in) (PDB ID: 4G5P); (c,g) Type II inhibitor; co-crystal structure of imatinib bound to
Bcr-Abl (DFG-out) (PDB ID: 1IEP); (d,h) Type III inhibitor; co-crystal structure of Tak-733 bound to
MEK1 (DFG-in) (PDB ID: 3PP1).
Figure 1. Representative examples of the primary types of small molecule kinase inhibitors including
a DFG-in irreversible inhibitor. Kinases are shown in grey, the activation loops in burgundy, the
glycine-rich loops in blue and the DFG motif in red (sticks). The carbon atoms of the inhibitors and
the ATP are respectively represented in cyan and green (ac); The chemical structures and schematic
binding modes of the inhibitors are depicted. For inhibitors that have been radiolabeled for PET
imaging, the position of the radionuclide (carbon-11 and fluorine-18) is indicated with a black arrow
(eg); (a,e) Type I inhibitor; co-crystal structure of erlotinib bound to epidermal growth factor receptor
(EGFR) (DFG-in) (PDB ID: 1M17); (b,f) Type I inhibitor (irreversible); co-crystal structure of Afatinib
bound to EGFR T790M (DFG-in) (PDB ID: 4G5P); (c,g) Type II inhibitor; co-crystal structure of imatinib
bound to Bcr-Abl (DFG-out) (PDB ID: 1IEP); (d,h) Type III inhibitor; co-crystal structure of Tak-733
bound to MEK1 (DFG-in) (PDB ID: 3PP1).
22001
Molecules 2015,20, 22000–22027
Type III inhibitors interact with the allosteric pocket without overlapping with the hinge region
(Figure 1d). In addition, more rarely encountered type IV inhibitors exploit binding sites distal from
the ATP site [13,14].
Despite the relatively high response rates encountered within this drug class, only a minority of
patients will ultimately benefit from kinase targeted therapy, due to target expression/engagement,
mutation status and pharmacokinetic parameters [15]. Moreover, the initial response rates are
hampered by drug resistance occurring over the course of treatment [16]. The identification of
treatment-responsive patient subsets for a given inhibitor, and the validation of the most suitable
lead candidates from emerging generations of inhibitors early in clinical settings constitute key
challenges associated with the advancement and clinical use of kinase inhibitors. In this context, the
implementation of nuclear imaging protocols as part of the clinical assessment of treatment response
and the drug development process has emerged as a promising approach [1719]. Positron emission
tomography (PET) imaging is among the most successful in vivo imaging technologies currently in
use. PET imaging relies on the coincidental detection of γrays ensuing from annihilation events
triggered by positron emitters incorporated into biologically relevant molecules. Small molecule
radiotracers for PET are typically labeled with 18F (t1/2 = 109.8 min; 97% β+;
Emax (β+) = 0.64 MeV)
or
11C (t1/2 = 20.3 min; 99.8% β+; Emax (β+) = 0.96 MeV) and generate high resolution three-dimensional
images showing the dynamic distribution of the molecular probe, thus allowing for the localization
and quantification of specific biological targets non-invasively. PET imaging has found broad
applications in oncology with the use of 18F-labeled 2-deoxyglucose ([18F]FDG) for the assessment
of tumor metabolic activity. Once injected, [18F]FDG is actively transported in metabolically active
cell, including tumor cells, and is trapped as [18F]FDG-6-phosphate, allowing tissue visualization.
In recent years, [18F]FDG PET and [18F]FDG PET/computed tomography (CT) have been utilized
in treatment monitoring imaging for patients undergoing kinase inhibitor therapy [17,2022]. This
approach though provides an indirect assessment for kinase inhibitor treatment response. A
more targeted approach relies on the PET evaluation of radiolabeled small molecule inhibitors.
Radiolabeled isotopologues of approved kinase inhibitors have the potential to differentiate
responders from non-responders undergoing treatment similarly to [18F]FDG. However, in this
context, and in contrast to [18F]FDG, the PET signal of interest ensues directly from the interaction of
the radiolabeled inhibitor with one or more kinase targets. In the absence of acquired resistance which
may compromise the binding of kinase inhibitors, in vivo PET imaging of radiolabeled approved
drugs or drug candidates can in theory provide expedient data regarding spatiotemporal protein
kinase expression. In addition, data relating to whole body distribution and metabolism of the
radiolabeled compound may be obtained [18,19]. Considering the magnitude of the kinase inhibitor
field and the number of inhibitor development programs, it is clear that the use of radiolabeled
inhibitors in early drug development stages can also assist in target identification and validation,
while defining compound prioritization and target engagement based on robust in vivo data. From a
more fundamental perspective, radiolabeled inhibitors taken from scaffolds under development may
offer opportunities in neurology and neuro-oncology to measure brain penetration and to quantify
kinase density in brain tissue under normal and pathological conditions.
As for radiotracers in general, stringent criteria have to be met in the design of radiolabeled
protein kinase inhibitors [2325]. These requirements are valid for the translation of both approved
inhibitors and compounds selected from preclinical screenings. First, the intended molecule should
display high affinity (typically low nanomolar to picomolar) for its target(s). This is of particular
importance for ATP-competitive kinase inhibitors targeting kinases with low KM, ATP values (vide
infra) due to the high cellular ATP concentrations (1–5 mM) [26]. Also, it is essential that the
compounds be efficiently radiolabeled in high specific activity (SA) and useful radiochemical yields
(RCYs) following a time-efficient procedure suitably matching the short half-life of the radionuclide
used (18F, 11C, or 64Cu). Kinase inhibitors are advantageous in this regard as they often contain readily
available positions for labeling with 11C and 18F. For example, of the 29 approved inhibitors to date, 15
22002
Molecules 2015,20, 22000–22027
bear at least one readily accessible O-CH3or N-CH3fragment for radiolabeling through conventional
radiomethylation methods with [11C]CH3I or [11C]CH3OTf (see examples in Figure 1e–f). Three of
these compounds also contain fluorine atoms in activated ortho- or para- aryl positions. The majority
of the approved inhibitors also bear additional potential, albeit more challenging, labeling sites such
as asymmetrical ureas, non-activated fluoroaryl moieties and tolyl groups. Compounds identified in
preclinical research can also be selected based on the availability of potential radiolabeling positions.
In cases where no labeling position is evident, rational radiotracer design may be straightforward
due to available crystallographic and structure-activity relationship (SAR) data—both of which are
extensive in this field. It is also possible that a readily accessible radiolabeling position may not
be suitable due to metabolic susceptibility. In such cases, 18F-defluorination will lead to bone
uptake while heteroatom demethylation, which is one of the primary CYP450 metabolic pathways,
may liberate reactive 11C-labeled side products, reducing signal-to-noise ratios and confounding
the target-specific PET signal. It is therefore important to include the choice of position for
labeling in a larger perspective which includes such in vivo considerations. Another important
element to consider is selectivity. A large fraction of approved kinase inhibitors are not selective
compounds [27,28]. While multitargeted inhibitors may still translate into useful PET radiotracers
in terms of biodistribution, metabolism and tumor imaging studies, their application outside of a
“drug validation” paradigm may be more limited. Finally, when considering brain imaging with
kinase inhibitors, multiple physicochemical properties including molecular weight, lipophilicity,
polar surface area and hydrogen bond donors become increasingly important due to the restrictive
blood brain barrier (BBB) [29,30]. In previous years, excellent reviews have covered the topics of
radiolabeled kinase inhibitors for PET imaging [25,31]. Within such a rapidly growing field, the
present review summarizes the work accomplished within the last 5 years and provides an overview
of the current stage of the field.
2. Synthesis and Evaluation of Radiolabeled Small Molecule Kinase Inhibitors for PET Imaging
2.1. Radiolabeled Tyrosine Kinase Inhibitors
The ErbB tyrosine kinase family is composed of analogous receptors which include EGFR
(Erb1B), HER2/neu (human epidermal growth factor receptor 2, ErbB2), HER3 (human epidermal
growth factor receptor 3, ErbB3) and HER4 (human epidermal growth factor receptor 4,
ErbB4). It is arguably the most investigated kinase group both in terms of signaling pathway
mapping and drug development. Multiple inhibitors targeting ErbB receptors, which are often
overexpressed or mutated in human cancer, have progressed into clinical research. Such
compounds tend to bear a 4-anilinoquinazoline core as a preferred hinge-binding motif. Expectedly,
4-anilinoquinazoline-bearing ErbB inhibitors were also the first and most explored radiolabeled
inhibitor class for PET imaging applications [3252]. Current applications in this class have converged
around 11C- and 18F-isotopologues of clinically approved inhibitors which in some cases has helped
fast-tracking the translation towards human imaging applications.
Gefitinib (Iressar, AstraZeneca) is a selective single-digit nanomolar EGFR inhibitor approved
as a third line of treatment in patients with non-small cell lung cancer (NSCLC). Positive clinical
response to gefitinib treatment is primarily dependent on two EGFR-activating mutations leading
to ligand-independent activation (the L858R mutation and exon 19 deletions). These mutations
significantly reduce ATP affinity for EGFR while simultaneously increasing inhibitor affinity which
elicits a favorable initial treatment response [53].
Subsequent mutation events (e.g., T790M) partially restore the affinity of ATP for EGFR
and impair type-I EGFR inhibitor treatment such as gefitinib or erlotinib (Tarcevar, OSI
Pharmaceuticals), inevitably leading to resistance. Of importance, only 5%–20% of NSCLC patients
carry EGFR-activating mutations [54]. Therefore, the pharmacokinetic evaluation of gefitinib and
tumoral EGFR status imaging has been the initial driver in the development of radiolabeled versions
22003
Molecules 2015,20, 22000–22027
of gefitinib [5559]. [11C]gefitinib ([11C]2, Figure 2a), obtained via [11C]CH3I or [11C]CH3OTf
methylation of precursor 1was evaluated in fibrosarcoma (NFSa)-bearing mice [59]. The murine
origin of the NFSa cell line limits the relevance of the study in the context of human cancer.
Moreover, although the tracer was shown to be stable and accumulated specifically in the tumor
model (Figure 2b), in vitro characterization of the NFSa cells failed to detect EGFR which suggests
that in vivo specific tumor binding in this case may not be related to EGFR expression. Those
results are in line with previous reports involving [18F]gefitinib. In this case [18F]gefitinib can be
obtained following a 3-step procedure starting from the trimethylammonium triflate precursor 3
(Figure 2c). An automated reliable procedure for routine production was recently reported which
provides [18F]gefitinib in 17.2% ˘3.3% RCY (decay corrected, d.c., n= 22) and >99% radiochemical
purity following a 2.5 h procedure [60]. Despite the lengthiness of this approach compared to the
11C-labeling, the longer half-life of 18F was justifiably considered advantageous. Unfortunately,
in vivo, despite suitable in vitro profile and in vivo stability, [18F]gefitinib failed to display uptake
in various xenograft models derived from EGFR-expressing human cell lines including treatment
responsive models (H3255 and H1975 cell lines) [56]. Both high non-specific binding due to
the lipophilicity of the radiotracer, and efflux susceptibility from the ATP-binding cassette (ABC)
transporter ABCB1 (P-glycoprotein) and ABCG2 (breast cancer resistance protein) have been put
forward to explain the observed results [25,56]. These transporters are highly expressed at the BBB, in
excretory organs, and in many tumors and they constitute the primary efflux proteins responsible for
the reduction of intracellular xenobiotic levels. Expression of ABCG2 is a well-known drug resistance
mechanism in gefitinib treatment [61]. In light of the positive results obtained with [11C]erlotinib
which has similar lipophilicity, discussed below, it appears likely that the lack of EGFR-specific signal
in microdosing PET experiments with [11C]/[18F]gefitinib ensues mainly from efflux mechanisms.
In fact, whereas both gefitinib and erlotinib are known dual ABCB1/ABCG2 substrates, gefitinib
efflux susceptibility is significantly more pronounced compared to erlotinib, as demonstrated by
comparative MDCKII permeability experiments [62].
Molecules 2015, 20, page–page
5
may not be related to EGFR expression. Those results are in line with previous reports involving
[18F]gefitinib. In this case [18F]gefitinib can be obtained following a 3-step procedure starting from the
trimethylammonium triflate precursor 3 (Figure 2c). An automated reliable procedure for routine
production was recently reported which provides [18F]gefitinib in 17.2% ± 3.3% RCY (decay corrected,
d.c., n = 22) and >99% radiochemical purity following a 2.5 h procedure [60]. Despite the lengthiness
of this approach compared to the 11C-labeling, the longer half-life of 18F was justifiably considered
advantageous. Unfortunately, in vivo, despite suitable in vitro profile and in vivo stability, [18F]gefitinib
failed to display uptake in various xenograft models derived from EGFR-expressing human cell lines
including treatment responsive models (H3255 and H1975 cell lines) [56]. Both high non-specific
binding due to the lipophilicity of the radiotracer, and efflux susceptibility from the ATP-binding
cassette (ABC) transporter ABCB1 (P-glycoprotein) and ABCG2 (breast cancer resistance protein) have
been put forward to explain the observed results [25,56]. These transporters are highly expressed at the
BBB, in excretory organs, and in many tumors and they constitute the primary efflux proteins
responsible for the reduction of intracellular xenobiotic levels. Expression of ABCG2 is a well-known
drug resistance mechanism in gefitinib treatment [61]. In light of the positive results obtained with
[11C]erlotinib which has similar lipophilicity, discussed below, it appears likely that the lack of
EGFR-specific signal in microdosing PET experiments with [11C]/[18F]gefitinib ensues mainly from
efflux mechanisms. In fact, whereas both gefitinib and erlotinib are known dual ABCB1/ABCG2
substrates, gefitinib efflux susceptibility is significantly more pronounced compared to erlotinib, as
demonstrated by comparative MDCKII permeability experiments [62].
Figure 2. Radiolabeled gefitinib. (a) Alternative conditions for the radiosynthesis of [11C]gefitinib;
(b) In vivo PET evaluation (coronal images) of [11C]gefitinib in tumor-bearing (NFSa) mice at baseline
(a) and with increasing non-radioactive gefitinib (bd) (adapted with kind permission from Springer
Science + Business Media: [59], Springer International Publishing© AG); (c) Radiosynthesis of [18F]gefitinib.
(SUV: standardized uptake value).
Taking advantage of these observations, two distinct studies, first using [11C]gefitinib [63] and
more recently using [18F]gefitinib [64], have explored the ABCB1/ABCG2-mediated brain penetration
of radiolabeled gefitinib. In the most recent study, the brain penetration of [18F]gefitinib was shown
to be limited by both transporters in a synergistic manner using Abcb1a/1b/, Abcg2/ and Abcb1a/1b;
Abcg2/ mice (Figure 3a,b). Pretreatment with the dual ABCB1/ABCG2 inhibitor elacridar (10 mg/kg)
led to enhanced brain penetration of the radiotracer in wild-type mice. With the limited availability
of 18F-labeled PET radioligands targeting both ABCB1/ABCG2 and the promising clinical applications
for such probes, this study suggests the repurposing of [18F]gefitinib as an imaging agent for the
assessment of ABCB1/ABCG2 activity in the context of CNS diseases. [18F]Gefitinib may also be
applicable for the quantification of drug-drug interactions (DDIs) at the BBB. As it appears likely that
upcoming studies will explore these avenues, work towards the validation of this potential application
will be positively influenced by the fact that gefitinib is already well established in the clinic.
Figure 2. Radiolabeled gefitinib. (a) Alternative conditions for the radiosynthesis of [11 C]gefitinib;
(b)In vivo PET evaluation (coronal images) of [11C]gefitinib in tumor-bearing (NFSa) mice at baseline
(a) and with increasing non-radioactive gefitinib (bd) (adapted with kind permission from Springer
Science + Business Media: [59], Springer International Publishing©AG); (c) Radiosynthesis of
[18F]gefitinib. (SUV: standardized uptake value).
Taking advantage of these observations, two distinct studies, first using [11C]gefitinib [63] and
more recently using [18F]gefitinib [64], have explored the ABCB1/ABCG2-mediated brain penetration
of radiolabeled gefitinib. In the most recent study, the brain penetration of [18 F]gefitinib was
shown to be limited by both transporters in a synergistic manner using Abcb1a/1b´{´,Abcg2´{´
and Abcb1a/1b;Abcg2´{´ mice (Figure 3a,b). Pretreatment with the dual ABCB1/ABCG2 inhibitor
22004
Molecules 2015,20, 22000–22027
elacridar (10 mg/kg) led to enhanced brain penetration of the radiotracer in wild-type mice. With
the limited availability of 18F-labeled PET radioligands targeting both ABCB1/ABCG2 and the
promising clinical applications for such probes, this study suggests the repurposing of [18F]gefitinib
as an imaging agent for the assessment of ABCB1/ABCG2 activity in the context of CNS diseases.
[18F]Gefitinib may also be applicable for the quantification of drug-drug interactions (DDIs) at the
BBB. As it appears likely that upcoming studies will explore these avenues, work towards the
validation of this potential application will be positively influenced by the fact that gefitinib is already
well established in the clinic.
Molecules 2015, 20, page–page
6
Figure 3. From gefitinib to AZD3759. (a) Representative PET images of [18F]gefitinib in wild-type and
Abcb1a/1b;Abcg2/ mice; (b) Comparative quantitative results of brain uptake in wild-type, Abcb1a/1b/,
Abcg2/ and Abcb1a/1b;Abcg2/ mice (*** p < 0.0001, summed 30–60 min post i.v. [18F]gefitinib injection)
(Images in (a,b) adapted from [64] with permission from Elsevier); (c) Chemical structure of [11C]AZD3759.
Most recently, AstraZeneca reported on the preclinical development and validation of AZD3759,
a brain penetrating orally active EGFR inhibitor currently undergoing a phase I clinical trial in patients
with mutation positive NSCLC [62,65]. This lead, derived from gefitinib, was rationally developed in
order to mitigate the efflux transport at the BBB and offer a new line of treatment for the significant
portion of NSCLC patients who will ultimately develop CNS metastases. A radiolabeled version of
this inhibitor, [11C]AZD3759 (Figure 3c), was developed as part of a brain tissue engagement proof-of-
principle PET microdosing study in cynomologus monkeys. The radiotracer showed excellent brain
penetration in line with the favorable MDCKII permeability results. This preliminary study provides
an unambiguous illustration of the potential of radiolabeled kinase inhibitors in drug development.
It will be interesting to see if [11C]AZD3759 can be applied to the imaging of brain metastases of EGFR
inhibitor-responsive NSCLC cases or if it will be synthesized as an 18F-labeled tracer in the near
future. The identification of [11C]AZD3759 also provides an alternative to ABCB1/ABCG2-substrate
radiolabeled EGFR inhibitors for peripheral tumor imaging and could be used to validate the
hypothesis that ABCB1/ABCG2 interaction has been the key determinant in the failure of [11C]gefitinib
and [18F]gefitinib for peripheral tumor imaging in preclinical models.
Erlotinib (Tarceva®, OSI Pharmaceutical) is a type-I EGFR inhibitor analogous to gefitinib, yet
more potent and less selective, which is efficient in similar mutation-positive NSCLC patients (EGFR
Kd = 0.67 nM) [27,66]. As mentioned above, erlotinib has been radiolabeled with carbon-11 and
evaluated favorably in an erlotinib-sensitive preclinical model using study designs comparable to those
used for the radiolabeled gefitinib cases (Figure 4a) [67–69]. Rapid clinical translation of [11C]erlotinib
provided promising preliminary results in two distinct studies with small NSCLC patient cohorts
(10 and 13 patients) [70,71]. In a first study, [11C]erlotinib was shown to accumulate in malignant lesions
and lymph node metastases of undefined mutation status. Interestingly, variation in [11C]erlotinib
accumulation in the primary tumor and metastatic sites within the same subject was observed,
suggesting that a single biopsy may provide incomplete data for EGFR profiling, supporting the
development of a [11C]erlotinib-based imaging procedure for EGFR status assessment (Figure 4b,c) [70].
In a following proof-of-concept study, the correlative relationship between EGFR mutation status, in
this case the 19 exon deletions, and [11C]erlotinib tumor uptake was demonstrated [71]. Interestingly,
[11C]erlotinib was also shown to accumulate in an erlotinib-responsive brain metastatic lesion from a
patient harboring an erlotinib-sensitized exon 19 deletion within the primary lung tumor (Figure 4d) [72].
The potential of [11C]erlotinib to image brain metastases in this context may be driven by the fact that
those patients may already present a compromised BBB. Collectively, these three studies supported
further investigation towards PET-driven personalized therapy based on the EGFR status in larger
clinical studies. However, Traxl et al., recently presented data evidencing potential problems with the
use of [11C]erlotinib for this purpose [73]. [11C]Erlotinib brain distribution was shown to be substantially
superior in Abcb1a/1b;Abcg2/ mice than previously described with [18F]gefitinib. Elacridar pretreatment
(10 mg/kg) in wild type mice restored brain uptake levels comparable to Abcb1a/1b;Abcg2/ mice, an
effect which was only partial using the same dosing with [18F]gefitinib. This difference may be related
in part to the fact the [18F]gefitinib study [64] used a therapeutically relevant [18F]gefitinib dose (1 mg/kg)
whereas the [11C]erlotinib/elacridar data was derived from [11C]erlotinib microdosing experiments.
Figure 3. From gefitinib to AZD3759. (a) Representative PET images of [18F]gefitinib in wild-type
and Abcb1a/1b;Abcg2´{´ mice; (b) Comparative quantitative results of brain uptake in wild-type,
Abcb1a/1b´{´,Abcg2´{´ and Abcb1a/1b;Abcg2´{´ mice (*** p< 0.0001, summed 30–60 min post
i.v. [18F]gefitinib injection) (Images in (a,b) adapted from [64] with permission from Elsevier);
(c) Chemical structure of [11C]AZD3759.
Most recently, AstraZeneca reported on the preclinical development and validation of AZD3759,
a brain penetrating orally active EGFR inhibitor currently undergoing a phase I clinical trial in
patients with mutation positive NSCLC [62,65]. This lead, derived from gefitinib, was rationally
developed in order to mitigate the efflux transport at the BBB and offer a new line of treatment
for the significant portion of NSCLC patients who will ultimately develop CNS metastases. A
radiolabeled version of this inhibitor, [11C]AZD3759 (Figure 3c), was developed as part of a
brain tissue engagement proof-of-principle PET microdosing study in cynomologus monkeys. The
radiotracer showed excellent brain penetration in line with the favorable MDCKII permeability
results. This preliminary study provides an unambiguous illustration of the potential of radiolabeled
kinase inhibitors in drug development. It will be interesting to see if [11C]AZD3759 can be applied
to the imaging of brain metastases of EGFR inhibitor-responsive NSCLC cases or if it will be
synthesized as an 18F-labeled tracer in the near future. The identification of [11C]AZD3759 also
provides an alternative to ABCB1/ABCG2-substrate radiolabeled EGFR inhibitors for peripheral
tumor imaging and could be used to validate the hypothesis that ABCB1/ABCG2 interaction has been
the key determinant in the failure of [11C]gefitinib and [18F]gefitinib for peripheral tumor imaging in
preclinical models.
Erlotinib (Tarcevar, OSI Pharmaceutical) is a type-I EGFR inhibitor analogous to gefitinib, yet
more potent and less selective, which is efficient in similar mutation-positive NSCLC patients (EGFR
Kd= 0.67 nM) [27,66]. As mentioned above, erlotinib has been radiolabeled with carbon-11 and
evaluated favorably in an erlotinib-sensitive preclinical model using study designs comparable to
those used for the radiolabeled gefitinib cases (Figure 4a) [6769]. Rapid clinical translation of
[11C]erlotinib provided promising preliminary results in two distinct studies with small NSCLC
patient cohorts (10 and 13 patients) [70,71]. In a first study, [11C]erlotinib was shown to accumulate
in malignant lesions and lymph node metastases of undefined mutation status. Interestingly,
variation in [11C]erlotinib accumulation in the primary tumor and metastatic sites within the same
subject was observed, suggesting that a single biopsy may provide incomplete data for EGFR
profiling, supporting the development of a [11C]erlotinib-based imaging procedure for EGFR status
assessment (Figure 4b,c) [70]. In a following proof-of-concept study, the correlative relationship
22005
Molecules 2015,20, 22000–22027
between EGFR mutation status, in this case the 19 exon deletions, and [11C]erlotinib tumor
uptake was demonstrated [71]. Interestingly, [11C]erlotinib was also shown to accumulate in an
erlotinib-responsive brain metastatic lesion from a patient harboring an erlotinib-sensitized exon
19 deletion within the primary lung tumor (Figure 4d) [72]. The potential of [11C]erlotinib to
image brain metastases in this context may be driven by the fact that those patients may already
present a compromised BBB. Collectively, these three studies supported further investigation towards
PET-driven personalized therapy based on the EGFR status in larger clinical studies. However,
Traxl et al., recently presented data evidencing potential problems with the use of [11C]erlotinib
for this purpose [73]. [11C]Erlotinib brain distribution was shown to be substantially superior
in Abcb1a/1b;Abcg2´{´ mice than previously described with [18F]gefitinib. Elacridar pretreatment
(10 mg/kg) in wild type mice restored brain uptake levels comparable to Abcb1a/1b;Abcg2´{´ mice, an
effect which was only partial using the same dosing with [18F]gefitinib. This difference may be related
in part to the fact the [18F]gefitinib study [64] used a therapeutically relevant [18F]gefitinib dose
(1 mg/kg) whereas the [11C]erlotinib/elacridar data was derived from [11C]erlotinib microdosing
experiments. An important observation from the work by Traxl and colleagues, which may
have important clinical significance for the use of [11C]erlotinib, is the fact that the distribution
of [11C]erlotinib in peripheral, and probably tumor tissues, was affected by ABCB1/ABCG2. It
follows that erlotinib displays non-linear pharmacokinetics due to efflux transporters and that it
may be challenging to derive information regarding erlotinib distribution at therapeutic doses from
[11C]erlotinib disposition in PET studies. This should be a factor to address in upcoming broader
validations of [11C]erlotinib in NSCLC patients.
Molecules 2015, 20, page–page
7
An important observation from the work by Traxl and colleagues, which may have important clinical
significance for the use of [11C]erlotinib, is the fact that the distribution of [11C]erlotinib in peripheral,
and probably tumor tissues, was affected by ABCB1/ABCG2. It follows that erlotinib displays
non-linear pharmacokinetics due to efflux transporters and that it may be challenging to derive
information regarding erlotinib distribution at therapeutic doses from [11C]erlotinib disposition in
PET studies. This should be a factor to address in upcoming broader validations of [11C]erlotinib in
NSCLC patients.
Figure 4. Selected in vivo results with [11C]erlotinib. (a) Chemical structure of [11C]erlotinib and
comparative PET imaging studies; (b) Bone metastasis (c) and tumor/lymph node metastasis
accumulation of [11C]erlotinib in non-small-cell lung carcinoma (with CT and [18F]FDG PET) (Figures
in (b,c) adapted by permission from Macmillan Publishers Ltd on behalf of Cancer Research©, UK, [70]
2011); (d) Magnetic resonance imaging-PET image from [11C]erlotinib accumulation in brain metastatic
lesions in a patient diagnosed with a non-small-cell lung carcinoma. (CT: computed tomography)
(Adapted by permission from [71], from the publisher Wolters Kluwer).
The comparative in vivo evaluation of [11C]erlotinib with the irreversible inhibitor [18F]afatinib
has been recently described in mutation sensitized and wild-type EGFR-expressing tumor bearing
mice [74]. Although irreversible radiolabeled EGFR inhibitors have been previously described, in vivo
data and preclinical validation in relevant tumor-bearing models have been limited so far [43–51].
Afatinib (Gilotrif®, Boehringer Ingelheim) is a type I EGFR, ErbB2/HER2, and ErbB4/HER4 inhibitor,
which displays low nanomolar affinity for EGFRL858R/T790M in contrast to first generation EGFR inhibitors.
Afatinib covalently and irreversibly reacts with hinge proximal cysteine residues (Figure 5a) [75].
Afatinib also displays off-ErbB selectivity comparable to first generation inhibitors such as gefitinib [75].
The radiosynthesis of [18F]afatinib employed a strategy reminiscent of the approach devised to obtain
[18F]gefitinib (Figure 5a) with the difference that the 3-chloro-4-[18F]fluoroaniline ([18F]5, Figure 5a)
intermediate was used in a BOP-mediated condensation reaction with 4-quinazolinone precursor 10
instead of a chloro precursor due to the presence of the Michael acceptor side chain [76]. Despite the
rather long radiosynthesis, [18F]afatinib was obtained in sufficient RCYs for in vivo evaluation (17.0%
± 2.5% RCY d.c.). Initial evaluation in A549 (wild-type EGFR) and HCC827 (exon 19 deletion EGFR)
tumors showed moderate tumor uptake in both models (about or under 1%ID/g from 0–120 min p.i.).
Unexpectedly, no difference in tumor uptake between these two cell lines was observed. As afatinib
is a known P-gp substrate [77], and immunohistochemical staining confirmed P-gp expression in both
cell lines, the possible influence of efflux transporters to explain those results were addressed with a
comparative evaluation of [18F]afatinib with [11C]erlotinib under baseline and tariquidar pretreament [74].
Overall, both tracers displayed similar tumor uptake kinetics and tumor-to-background ratios in
three tumor models (including A549 and HCC827) at baseline. The only divergence from those results
following tariquidar administration was observed in the HCC827 model with [18F]afatinib which
Figure 4. Selected in vivo results with [11C]erlotinib. (a) Chemical structure of [11C]erlotinib
and comparative PET imaging studies; (b) Bone metastasis (c) and tumor/lymph node metastasis
accumulation of [11C]erlotinib in non-small-cell lung carcinoma (with CT and [18F]FDG PET) (Figures
in (b,c) adapted by permission from Macmillan Publishers Ltd on behalf of Cancer Research©,
UK, [70] 2011); (d) Magnetic resonance imaging-PET image from [11C]erlotinib accumulation in brain
metastatic lesions in a patient diagnosed with a non-small-cell lung carcinoma. (CT: computed
tomography) (Adapted by permission from [71], from the publisher Wolters Kluwer).
The comparative in vivo evaluation of [11C]erlotinib with the irreversible inhibitor [18F]afatinib
has been recently described in mutation sensitized and wild-type EGFR-expressing tumor bearing
mice [74]. Although irreversible radiolabeled EGFR inhibitors have been previously described,
in vivo data and preclinical validation in relevant tumor-bearing models have been limited so
far [4351]. Afatinib (Gilotrifr, Boehringer Ingelheim) is a type I EGFR, ErbB2/HER2, and
ErbB4/HER4 inhibitor, which displays low nanomolar affinity for EGFRL858R/T790M in contrast to
first generation EGFR inhibitors. Afatinib covalently and irreversibly reacts with hinge proximal
22006
Molecules 2015,20, 22000–22027
cysteine residues (Figure 5a) [75]. Afatinib also displays off-ErbB selectivity comparable to first
generation inhibitors such as gefitinib [75]. The radiosynthesis of [18 F]afatinib employed a strategy
reminiscent of the approach devised to obtain [18F]gefitinib (Figure 5a) with the difference that
the 3-chloro-4-[18F]fluoroaniline ([18F]5, Figure 5a) intermediate was used in a BOP-mediated
condensation reaction with 4-quinazolinone precursor 10 instead of a chloro precursor due to the
presence of the Michael acceptor side chain [76]. Despite the rather long radiosynthesis, [18F]afatinib
was obtained in sufficient RCYs for in vivo evaluation (17.0% ˘2.5% RCY d.c.). Initial evaluation
in A549 (wild-type EGFR) and HCC827 (exon 19 deletion EGFR) tumors showed moderate tumor
uptake in both models (about or under 1%ID/g from 0–120 min p.i.). Unexpectedly, no difference in
tumor uptake between these two cell lines was observed. As afatinib is a known P-gp substrate [77],
and immunohistochemical staining confirmed P-gp expression in both cell lines, the possible
influence of efflux transporters to explain those results were addressed with a comparative evaluation
of [18F]afatinib with [11C]erlotinib under baseline and tariquidar pretreament [74]. Overall, both
tracers displayed similar tumor uptake kinetics and tumor-to-background ratios in three tumor
models (including A549 and HCC827) at baseline. The only divergence from those results following
tariquidar administration was observed in the HCC827 model with [18F]afatinib which demonstrated
significantly higher absolute tumor uptake (1.9% ˘0.1%ID/g vs. 1.2% ˘0.2%ID/g). Concomitant
higher background (from contralateral tissue) however led to unaltered tumor-to-background ratios.
Selective irreversible inhibitors may be advantageous compared to reversible ATP-competitive
inhibitors in terms of residence time [78,79]. It may be of interest to further explore the potential
of this binding mode for in vivo PET imaging (see Figure 7).
Molecules 2015, 20, page–page
8
demonstrated significantly higher absolute tumor uptake (1.9% ± 0.1%ID/g vs. 1.2% ± 0.2%ID/g).
Concomitant higher background (from contralateral tissue) however led to unaltered tumor-to-
background ratios. Selective irreversible inhibitors may be advantageous compared to reversible
ATP-competitive inhibitors in terms of residence time [78,79]. It may be of interest to further explore
the potential of this binding mode for in vivo PET imaging (see Figure 7).
Figure 5. (a) Radiosynthesis of [
18
F]afatinib and chemical structures of the irreversible covalent adducts
upon interaction at the ATP binding sites of the ErBb protein family; (b) Comparative coronal PET
images of [
18
F]afatinib in lung cancer-bearing mice with different EGFR mutation status. Red arrows
indicate reference tissue. (This research was originally published in EJNMMI Research, [74]).
PD153035 is an experimental EGFR inhibitor, prototypical of the 4-anilinoquinazoline scaffold [80].
Along with staurospaurine analogues [81], [
11
C]PD153035 (Figure 6a) is one of the first reported
radiolabeled protein kinase inhibitors [34,35,82–84]. PET [
11
C]PD153035 has been used in biodistribution
and radiation dosimetry studies in humans [85] and PET/CT [
11
C]PD153035 was validated as a
noninvasive survival predictor in advanced chemotherapy-refractory NSCLC in a 21 patient cohort
undergoing erlotinib treatment (Figure 6b) [86]. A detailed metabolic investigation in rodents was
also reported [87]. More recently, preliminary data detailing the potential of [
11
C]PD153035 for
EGFR-expressing glioma imaging has been reported [88]. In that study, [
11
C]PD153035 tracer uptake
correlated to EGFR expression irrespective of the mutation status. Although this tracer is also likely
to be amongst the most investigated radiolabeled kinase inhibitors, more recent research in EGFR
imaging has shifted towards approved derivatives. Its further use as a radiotracer for lung cancer or
malignant gliomas may also be impeded by the same intrinsic limitations as mentioned above for
other 4-anilinoquinazoline-based EGFR radiolabeled inhibitors which were not optimized for efflux
transporter liabilities.
Lapatinib (Tykerb
®
, GlaxoSmithKline) is a highly selective dual EGFR, ErbB2/HER2 inhibitor
approved in combination therapy for the treatment of HER2-overexpressing advanced metastatic
breast cancer. Different to other ErbB inhibitors described so far, lapatinib is a type-II inhibitor which
interacts with the inactivated DFG-out kinase conformation of EGFR and HER2 [3]. In recent years,
two radiolabeled versions of lapatinib have been described. The radiosynthesis of [
18
F]lapatinib was
first reported by Basuli and colleagues and employed a manual multi-step strategy in which the
3-[
18
F]fluorobenzyl bromide intermediate was reacted with the Boc protected intermediate 14
(Figure 6d) [89]. This study did not include a biological evaluation of [
18
F]lapatinib. While the half-life of
fluorine-18 is more suitable than carbon-11, a recent clinical study opted to develop a carbon-11
lapatinib isotopologue ([
11
C]lapatinib, Figure 6d) [90]. Since this clinical trial was directed at the
exploration of brain and CNS metastasis penetration of lapatinib in breast cancer patients with
secondary brain tumors, the carbon-11 option may have been preferred in order to exclude potential
Figure 5. (a) Radiosynthesis of [18F]afatinib and chemical structures of the irreversible covalent
adducts upon interaction at the ATP binding sites of the ErBb protein family; (b) Comparative coronal
PET images of [18F]afatinib in lung cancer-bearing mice with different EGFR mutation status. Red
arrows indicate reference tissue. (This research was originally published in EJNMMI Research, [74]).
PD153035 is an experimental EGFR inhibitor, prototypical of the 4-anilinoquinazoline
scaffold [80]. Along with staurospaurine analogues [81], [11C]PD153035 (Figure 6a) is one of the
first reported radiolabeled protein kinase inhibitors [34,35,8284]. PET [11C]PD153035 has been
used in biodistribution and radiation dosimetry studies in humans [85] and PET/CT [11C]PD153035
was validated as a noninvasive survival predictor in advanced chemotherapy-refractory NSCLC
in a 21 patient cohort undergoing erlotinib treatment (Figure 6b) [86]. A detailed metabolic
investigation in rodents was also reported [87]. More recently, preliminary data detailing the
potential of [11C]PD153035 for EGFR-expressing glioma imaging has been reported [88]. In that
study, [11C]PD153035 tracer uptake correlated to EGFR expression irrespective of the mutation status.
Although this tracer is also likely to be amongst the most investigated radiolabeled kinase inhibitors,
22007
Molecules 2015,20, 22000–22027
more recent research in EGFR imaging has shifted towards approved derivatives. Its further use
as a radiotracer for lung cancer or malignant gliomas may also be impeded by the same intrinsic
limitations as mentioned above for other 4-anilinoquinazoline-based EGFR radiolabeled inhibitors
which were not optimized for efflux transporter liabilities.
Lapatinib (Tykerbr, GlaxoSmithKline) is a highly selective dual EGFR, ErbB2/HER2 inhibitor
approved in combination therapy for the treatment of HER2-overexpressing advanced metastatic
breast cancer. Different to other ErbB inhibitors described so far, lapatinib is a type-II inhibitor
which interacts with the inactivated DFG-out kinase conformation of EGFR and HER2 [3]. In
recent years, two radiolabeled versions of lapatinib have been described. The radiosynthesis
of [18F]lapatinib was first reported by Basuli and colleagues and employed a manual multi-step
strategy in which the 3-[18F]fluorobenzyl bromide intermediate was reacted with the Boc protected
intermediate 14 (Figure 6d) [89]. This study did not include a biological evaluation of [18F]lapatinib.
While the half-life of fluorine-18 is more suitable than carbon-11, a recent clinical study opted to
develop a carbon-11 lapatinib isotopologue ([11C]lapatinib, Figure 6d) [90]. Since this clinical trial
was directed at the exploration of brain and CNS metastasis penetration of lapatinib in breast
cancer patients with secondary brain tumors, the carbon-11 option may have been preferred in
order to exclude potential confounding factors due to defluorination issues leading to cranial
18F´deposition. [11C]Lapatinib was synthesized following a two-pot four-step procedure making
use of [11C]-3-fluorobenzyl iodide (via Grignard reaction with [11C]CO2) and precursor 15 (Figure 6d).
Although the implementation of a Grignard reaction using trace amounts of radiolabeled [11C]CO2
remains a challenging task, the authors successfully synthesized [11C]lapatinib in useful RCYs and
satisfactory specific activity following an automated process. The study showed that [11C]lapatinib
is stable in vivo and differentially accumulates in normal brain tissue vs. brain metastasis enabling
tumor visualization (Figure 6c). With lapatinib being a known ABCB1/ABCG2 substrate [91], it was
hypothesized that therapeutic doses could partially saturate these efflux transporters and enhance
brain uptake. However, [11C]lapatinib PET following lapatinib administration at therapeutic doses
did not result in enhanced brain penetration suggesting that prophylactic treatment with lapatinib
to prevent brain metastasis formation is likely to prove unsuccessful. It is interesting to note, as
did the authors of this study, that the earlier completion of this study would have provided valuable
information to a larger concomitant study attempting to determine the possible prophylactic potential
of lapatinib in HER2-positive metastatic breast cancer patients [92]. This illustrates furthermore the
potential of radiolabeled kinase inhibitors in the drug development process.
Molecules 2015, 20, page–page
9
confounding factors due to defluorination issues leading to cranial 18F deposition. [11C]Lapatinib
was synthesized following a two-pot four-step procedure making use of [11C]-3-fluorobenzyl iodide
(via Grignard reaction with [11C]CO2) and precursor 15 (Figure 6d). Although the implementation of
a Grignard reaction using trace amounts of radiolabeled [11C]CO2 remains a challenging task, the
authors successfully synthesized [11C]lapatinib in useful RCYs and satisfactory specific activity
following an automated process. The study showed that [11C]lapatinib is stable in vivo and differentially
accumulates in normal brain tissue vs. brain metastasis enabling tumor visualization (Figure 6c). With
lapatinib being a known ABCB1/ABCG2 substrate [91], it was hypothesized that therapeutic doses
could partially saturate these efflux transporters and enhance brain uptake. However, [11C]lapatinib
PET following lapatinib administration at therapeutic doses did not result in enhanced brain
penetration suggesting that prophylactic treat ment with lapatinib to prevent brain metastasis formation
is likely to prove unsuccessful. It is interesting to note, as did the authors of this study, that the earlier
completion of this study would have provided valuable information to a larger concomitant study
attempting to determine the possible prophylactic potential of lapatinib in HER2-positive metastatic
breast cancer patients [92]. This illustrates furthermore the potential of radiolabeled kinase inhibitors
in the drug development process.
Figure 6. (a) Chemical structure of [11C]PD153035; (b) [11C]PD153035 accumulation in a adenocarcinoma
(A,C) and squamous cell carcinoma (B,D) with corresponding PET and CT images (originally published
in [86] © the Society of Nuclear Medicine and Molecular Imaging, Inc.); (c) [11C]Lapatinib accumulation
in a brain metastasis from a Her-2-positive breast cancer patient with corresponding MRI images
(originally published in [90] © 2015 Springer); (d) Radiosynthesis of [11C]lapatinib and [18F]lapatinib.
Within the last five years, multiple additional experimental type-I ErbB inhibitors exclusively
based on the 4-anilinoquinazoline scaffold were synthesized and in some instances evaluated in vivo
in preclinical settings [49,93–101]. In addition to the data presented for [18F]afatinib, work towards
the development of cysteine reactive irreversible radiotracers have included compounds [18F]16, [18F]17
and [18F]PEG6-IPQA [93–96] (Figure 7). Of these, [18F]16 and [18F]PEG6-IPQA were investigated in vivo
and were shown to preferentially accumulate in high EGFR-expressing A431 tumor xenografts.
Together with the data gleaned from the [18F]afatinib preliminary evaluation, these studies failed so
far to indicate a distinctive advantage of irreversible over reversible ErbB inhibitors for PET imaging
in preclinical models. Novel exploratory ErbB reversible inhibitors include a carbon-11 isotopologue of
the clinical candidate AZD8931, [11C]AZD8931 ([18F]17, Figure 7), but no imaging data accompanied the
report relating this radiosynthesis [98]. As can be expected, despite in some instances similar pre-clinical
data with these experimental inhibitors as compared to approved radiolabeled compounds, these
efforts have still remained at preclinical levels years after their original publication. Overall,
4-anilinoquinazoline-based inhibitors remain the only ErbB radiotracer scaffold explored to date.
In this context, the radiolabeling and validation of emerging non-4-anilinoquinazoline ErbB inhibitor
scaffolds may prove advantageous [102–105].
Figure 6. (a) Chemical structure of [11C]PD153035; (b) [11C]PD153035 accumulation in a
adenocarcinoma (A,C) and squamous cell carcinoma (B,D) with corresponding PET and CT images
(originally published in [86] © the Society of Nuclear Medicine and Molecular Imaging, Inc.);
(c) [11C]Lapatinib accumulation in a brain metastasis from a Her-2-positive breast cancer patient
with corresponding MRI images (originally published in [90] © 2015 Springer); (d) Radiosynthesis
of [11C]lapatinib and [18F]lapatinib.
22008
Molecules 2015,20, 22000–22027
Within the last five years, multiple additional experimental type-I ErbB inhibitors exclusively
based on the 4-anilinoquinazoline scaffold were synthesized and in some instances evaluated
in vivo in preclinical settings [49,93101]. In addition to the data presented for [18F]afatinib, work
towards the development of cysteine reactive irreversible radiotracers have included compounds
[18F]16, [18F]17 and [18F]PEG6-IPQA [9396] (Figure 7). Of these, [18F]16 and [18F]PEG6-IPQA were
investigated in vivo and were shown to preferentially accumulate in high EGFR-expressing A431
tumor xenografts. Together with the data gleaned from the [18F]afatinib preliminary evaluation,
these studies failed so far to indicate a distinctive advantage of irreversible over reversible ErbB
inhibitors for PET imaging in preclinical models. Novel exploratory ErbB reversible inhibitors include
a carbon-11 isotopologue of the clinical candidate AZD8931, [11C]AZD8931 ([18 F]17, Figure 7), but
no imaging data accompanied the report relating this radiosynthesis [98]. As can be expected,
despite in some instances similar pre-clinical data with these experimental inhibitors as compared
to approved radiolabeled compounds, these efforts have still remained at preclinical levels years
after their original publication. Overall, 4-anilinoquinazoline-based inhibitors remain the only ErbB
radiotracer scaffold explored to date. In this context, the radiolabeling and validation of emerging
non-4-anilinoquinazoline ErbB inhibitor scaffolds may prove advantageous [102105].
Molecules 2015, 20, page–page
10
Figure 7. Chemical structures of recently identified and evaluated ErbB-1/ErbB-2/ErbB-3
kinase inhibitors.
Inhibitors of the vascular endothelial growth factor receptors (VEGFR1, VEGFR2, VEGFR3),
frequently overexpressed in human cancers, have been extensively explored and proven clinically
beneficial for the blocking of critical tumoral angiogenic and growth pathways. Approved inhibitors
having VEGFRs as their primary kinase targets include: sorafenib (Nexavar®, Bayer), sunitinib (Sutent®,
Pfizer), pazopanib (Votrient®, GlaxoSmithKline), axitinib (Inlyta®, Pfizer), regorafenib (Stivarga®,
Bayer), nintedanib (Ofev®, Boehringer Ingelheim), lenvatinib (Lenvima®, Eisai Inc.) and vandetanib
(Caprelsa®, AstraZeneca) [3]. Inhibitors of this class tend to display various levels of promiscuity,
disrupting several non-VEGFR kinases in similar nanomolar potencies, including the platelet-derived
growth factor receptor (PDFGR), c-Kit, Fms-like tyrosine kinase 3 (Flt-3), RET, TIE-2 and Raf kinases
[27,28,106–110]. Despite the fact that the translation of these polypharmacological inhibitors into PET
radiotracers may still generate useful results for tumor imaging (where many such kinases may be
overexpressed in parallel), these probes are likely to be of limited use as tools to elucidate VEGFR
expression selectively.
Derivatives or isotopologues of vandetanib [111–113], sorafenib [114,115], sunitinib [116] and
nintedanib [117] as well as preclinical leads [118–120] have been radiolabeled for PET imaging. The
initial radiosynthesis and evaluation of the vandetanib analogue R-[11C]PAQ in the genetically
modified metastatic MMTV-PyMT mice model showed only marginal tumor uptake (Figure 8a) [112].
Straightforward [11C]CH3OTf methylation of suitable des-methyl precursors yielded the O-[11C]20,
N-[11C]20, O-[11C]21 and N-[11C]21 radiotracers, but none were evaluated in vivo [113]. A major
preclinical contribution by Li and colleagues [111] recently demonstrated the promising potential of
a 64Cu-labeled vandetanib-based dimer probe ([64Cu]24) using the U-87MG tumor xenograft model
(Figure 8b; t1/2 64Cu = 12.7 h) [111]. Although previous EGFR inhibitors have been labeled with 99mTc
for SPECT [121] (but not evaluated in vivo), this is the first instance of both a 64Cu-labeled kinase
inhibitor and of the utilization of a multimerization strategy in this context. Receptor binding assays
demonstrated a 100-fold affinity improvement for the dimeric probe [64Cu]24 vs. a monomeric analogous
derivative (44.7 nM for [64Cu]23 vs. 0.45 nM for [64Cu]24 in U-87MG cells). Remarkably, a pronounced
difference in tumor uptake was observed between [64Cu]23 and [64Cu]24. While [64Cu]23 displayed
very low tumor uptake (0.46% ± 0.06%ID/g, 24 h p.i.), [64Cu]24 showed rapid and lasting specific
accumulation into the U-87MG tumors (3.84% ± 0.05%ID/g, 24 h p.i.), (Figure 8c,d). Tumor-to-muscle
ratio was optimal at ~5 h post injection (p.i.) (~40) and decreased thereafter, but remained >30 until
the last time point imaged (24 h p.i.). Such ratios are significantly superior to what is typically
observed with small molecule kinase inhibitors labeled with carbon-11 or fluorine-18 within the
allowable scanning time frame permitted by those radioisotopes (typically 60–90 min p.i.). No in vivo
stability data were provided. With the recent demonstration that free 64Cu accumulates in U-87MG
tumor tissue in vivo (among other tumor cells lines) [122], this data would prove valuable. The success
of [64Cu]24 compared to [64Cu]23 was attributed to the favorable synergistic effect of the multivalent
inhibitor and the possible prolonged circulation time and slower tumor washout. This unique work
raises important questions regarding the value of using longer-lived isotopes to image kinases with
small molecule-type constructs. It appears that this type of multivalency platform deserves careful
examination in the PET kinase inhibitor field as a whole.
Figure 7. Chemical structures of recently identified and evaluated ErbB-1/ErbB-2/ErbB-3
kinase inhibitors.
Inhibitors of the vascular endothelial growth factor receptors (VEGFR1, VEGFR2, VEGFR3),
frequently overexpressed in human cancers, have been extensively explored and proven clinically
beneficial for the blocking of critical tumoral angiogenic and growth pathways. Approved inhibitors
having VEGFRs as their primary kinase targets include: sorafenib (Nexavarr, Bayer), sunitinib
(Sutentr, Pfizer), pazopanib (Votrientr, GlaxoSmithKline), axitinib (Inlytar, Pfizer), regorafenib
(Stivargar, Bayer), nintedanib (Ofevr, Boehringer Ingelheim), lenvatinib (Lenvimar, Eisai Inc.) and
vandetanib (Caprelsar, AstraZeneca) [3]. Inhibitors of this class tend to display various levels of
promiscuity, disrupting several non-VEGFR kinases in similar nanomolar potencies, including the
platelet-derived growth factor receptor (PDFGR), c-Kit, Fms-like tyrosine kinase 3 (Flt-3), RET, TIE-2
and Raf kinases [27,28,106110]. Despite the fact that the translation of these polypharmacological
inhibitors into PET radiotracers may still generate useful results for tumor imaging (where many
such kinases may be overexpressed in parallel), these probes are likely to be of limited use as tools to
elucidate VEGFR expression selectively.
Derivatives or isotopologues of vandetanib [111113], sorafenib [114,115], sunitinib [116] and
nintedanib [117] as well as preclinical leads [118120] have been radiolabeled for PET imaging.
The initial radiosynthesis and evaluation of the vandetanib analogue R-[11C]PAQ in the genetically
modified metastatic MMTV-PyMT mice model showed only marginal tumor uptake (Figure 8a) [112].
Straightforward [11C]CH3OTf methylation of suitable des-methyl precursors yielded the O-[11C]20,
N-[11C]20,O-[11C]21 and N-[11C]21 radiotracers, but none were evaluated in vivo [113]. A major
preclinical contribution by Li and colleagues [111] recently demonstrated the promising potential of
a64Cu-labeled vandetanib-based dimer probe ([64Cu]24) using the U-87MG tumor xenograft model
(Figure 8b; t1/2 64Cu = 12.7 h) [111]. Although previous EGFR inhibitors have been labeled with
22009
Molecules 2015,20, 22000–22027
99mTc for SPECT [121] (but not evaluated in vivo), this is the first instance of both a 64Cu-labeled
kinase inhibitor and of the utilization of a multimerization strategy in this context. Receptor binding
assays demonstrated a 100-fold affinity improvement for the dimeric probe [64Cu]24 vs. a monomeric
analogous derivative (44.7 nM for [64Cu]23 vs. 0.45 nM for [64Cu]24 in U-87MG cells). Remarkably,
a pronounced difference in tumor uptake was observed between [64Cu]23 and [64Cu]24. While
[64Cu]23 displayed very low tumor uptake (0.46% ˘0.06%ID/g, 24 h p.i.), [64Cu]24 showed rapid and
lasting specific accumulation into the U-87MG tumors (3.84% ˘0.05%ID/g, 24 h p.i.), (Figure 8c,d).
Tumor-to-muscle ratio was optimal at ~5 h post injection (p.i.) (~40) and decreased thereafter, but
remained >30 until the last time point imaged (24 h p.i.). Such ratios are significantly superior to
what is typically observed with small molecule kinase inhibitors labeled with carbon-11 or fluorine-18
within the allowable scanning time frame permitted by those radioisotopes (typically 60–90 min p.i.).
No in vivo stability data were provided. With the recent demonstration that free 64Cu accumulates in
U-87MG tumor tissue in vivo (among other tumor cells lines) [122], this data would prove valuable.
The success of [64Cu]24 compared to [64Cu]23 was attributed to the favorable synergistic effect of
the multivalent inhibitor and the possible prolonged circulation time and slower tumor washout.
This unique work raises important questions regarding the value of using longer-lived isotopes to
image kinases with small molecule-type constructs. It appears that this type of multivalency platform
deserves careful examination in the PET kinase inhibitor field as a whole.
Molecules 2015, 20, page–page
11
Figure 8. VEGFRs targeted radiotracers. (a) Chemical structures of radiolabeled inhibitors primarily
targeting VEGFRs: [11C]vandetanib, [11C]chloro-vandetanib and R-[11C]PAQ; (b) Chemical structures
of vandetanib-derivatized 64Cu-labeled monomeric ([64Cu]23) and dimeric ([64Cu]24) radiolabeled
inhibitors; (c) Coronal PET/CT images of inhibitors [64Cu]23 and [64Cu]24 in U-87 tumor bearing mice
and (d) corresponding quantitative analysis. ** p < 0.0001 (this research was originally published in [111]
© the Society of Nuclear Medicine and Molecular Imaging, Inc.).
Sorafenib is a multikinase type-II inhibitor which was initially labeled at the carbonyl position using
[11C]phosgene (obtained from [11C]CO2 [11C]CH4 [11C]CCl4 [11C]COCl2) by Asakawa et al. [114].
The tracer [carbonyl-11C]sorafenib was obtained in RCYs of 8%–11% (from [11C]CO2) in a 40 min
synthesis using precursor 25 (Figure 8a). No tumor imaging results were presented but injection in
Abcb1a/1b;Abcg2/ mice confirmed that sorafenib brain accumulation is limited by both transporters [123].
A second study by Poot and colleagues delineated an alternative route towards [carbonyl-11C]sorafenib
making use a [11C]carbon monoxide rhodium-mediated carbonylation reaction [115]. This approach
delivered [carbonyl-11C]sorafenib in higher RCY (27%) than the [11C]phosgene synthesis but was
abandoned in favor of a more straightforward and reliable [11C]CH3I methylation synthesis as both
radiotracers were shown to be similarly stable in vivo ([methyl-11C]sorafenib, Figure 9b). This alternative
tracer, [methyl-11C]sorafenib, was obtained in 60% RCY (d.c.) and evaluated in three xenografts from
cell lines characterized for Raf-1 expression. Of those, only the renal cancer cell line RXF393 xenografts
showed modest tumor uptake over the background signal (2.52% ± 0.33%ID/g, 7.5 min p.i.). The limited
tumor uptake and tumor-to-background ratios with this probe may be related to its inherent
promiscuity and the influence of efflux transporters. Neither data regarding the expression in the
tested xenograft models of other high affinity kinase targets of sorafenib (which include VEGFR) nor
blocking experiment results were provided. Other radiolabeled multitargeted inhibitors included
[11C]ATV-1, the sunitinib derivative [11C]29 and the 1H-pyrrole-2,5-dione inhibitor [11C]20. Interestingly,
[11C]ATV-1 was used in a rat model of myocardial infarction (MI) as a putative preclinical tool to
image angiogenesis processes during tissue repair following MI. [11C]ATV-1 showed superior uptake
in infarct region after MI induction in correlation with Tie-2, PDGFRα and VEGFR-2 expression as
validated by immunohistochemistry.
Figure 8. VEGFRs targeted radiotracers. (a) Chemical structures of radiolabeled inhibitors primarily
targeting VEGFRs: [11 C]vandetanib, [11C]chloro-vandetanib and R-[11C]PAQ; (b) Chemical structures
of vandetanib-derivatized 64Cu-labeled monomeric ([64Cu]23) and dimeric ([64Cu]24) radiolabeled
inhibitors; (c) Coronal PET/CT images of inhibitors [64Cu]23 and [64Cu]24 in U-87 tumor bearing
mice and (d) corresponding quantitative analysis. ** p< 0.0001 (this research was originally published
in [111] © the Society of Nuclear Medicine and Molecular Imaging, Inc.).
Sorafenib is a multikinase type-II inhibitor which was initially labeled at the carbonyl
position using [11C]phosgene (obtained from [11C]CO2Ñ[11C]CH4Ñ[11C]CCl4Ñ[11C]COCl2) by
Asakawa et al. [114]. The tracer [carbonyl-11C]sorafenib was obtained in RCYs of 8%–11% (from
[11C]CO2) in a 40 min synthesis using precursor 25 (Figure 8a). No tumor imaging results were
presented but injection in Abcb1a/1b;Abcg2´{´ mice confirmed that sorafenib brain accumulation is
limited by both transporters [123]. A second study by Poot and colleagues delineated an alternative
route towards [carbonyl-11C]sorafenib making use a [11C]carbon monoxide rhodium-mediated
carbonylation reaction [115]. This approach delivered [carbonyl-11C]sorafenib in higher RCY (27%)
than the [11C]phosgene synthesis but was abandoned in favor of a more straightforward and
reliable [11C]CH3I methylation synthesis as both radiotracers were shown to be similarly stable
in vivo ([methyl-11C]sorafenib, Figure 9b). This alternative tracer, [methyl-11C]sorafenib, was obtained
22010
Molecules 2015,20, 22000–22027
in 60% RCY (d.c.) and evaluated in three xenografts from cell lines characterized for Raf-1
expression. Of those, only the renal cancer cell line RXF393 xenografts showed modest tumor
uptake over the background signal (2.52% ˘0.33%ID/g, 7.5 min p.i.). The limited tumor uptake
and tumor-to-background ratios with this probe may be related to its inherent promiscuity and the
influence of efflux transporters. Neither data regarding the expression in the tested xenograft models
of other high affinity kinase targets of sorafenib (which include VEGFR) nor blocking experiment
results were provided. Other radiolabeled multitargeted inhibitors included [11C]ATV-1, the sunitinib
derivative [11C]29 and the 1H-pyrrole-2,5-dione inhibitor [11C]20. Interestingly, [11C]ATV-1 was used
in a rat model of myocardial infarction (MI) as a putative preclinical tool to image angiogenesis
processes during tissue repair following MI. [11C]ATV-1 showed superior uptake in infarct region
after MI induction in correlation with Tie-2, PDGFRαand VEGFR-2 expression as validated
by immunohistochemistry.
Molecules 2015, 20, page–page
12
Figure 9. Other radiolabeled multikinase inhibitors targeting VEGFRs. (a) Radiosynthesis of [carbonyl-
11C]sorafenib; (b) Chemical structure of [methyl-11C]sorafenib; (c) [18F]FDG (A) and [methyl-11C]sorafenib
coronal PET images of RXF393 tumor bearing mice (reprinted from [115] with permission from Elsevier);
(d) Chemical structures of additional recently radiosynthesized multikinase inhibitors targeting VEGFRs.
Imatinib was the first FDA approved kinase inhibitor and initially approved for the treatment of
CML via Bcr-Abl inhibiton. Imatinib also strongly inhibits c-Kit and PDGFR but shows a fairly good
selectivity outside those targets. The DFG-out binding mode of imatinib was a serendipitous discovery
that served as a foundation in the development of inhibitors targeting inactive kinase conformations [1].
The initial [11C]imatinib PET study delineated biodistribution and pharmacokinetic data including
low brain uptake compatible with active transporter efflux in baboons, but was not followed by
advanced preclinical experiments using tumor-bearing mice (Figure 10a) [124]. Instead, [18F]SKI696,
a fluorinated imatinib surrogate [125], and [124I]SKI230 [126], were both developed and shown to
accumulate moderately in Bcr-Abl overexpressing K562 cell xenografts (human immortalized
myelogenous leukemia cells, Figure 10b–d) [127]. [18F]SKI696 was synthesized by reacting 1-bromo-
2-[18F]fluoroethane with precursor 32. Despite detectable tumor uptake (1.2% ± 0.4%ID/g, 60 min p.i.),
unfavorable tumor-to-background ratio and high radioactivity uptake in the abdominal cavity were
observed (Figure 10d). More recently, preclinical imaging of [18F]SKI696 (also identified as [18F]STI-575)
was revisited by Peng and colleagues [128]. This study confirmed the previous K562 xenografts
findings and showed similarly low uptake profile in c-Kit expressing U87WT tumor-bearing mice.
In the absence of a clear imaging rationale for Bcr-Abl targeting in CML, and a potentially limited
applicability to c-Kit-positive gastrointestinal stromal tumor (GIST), for which imatinib is also
approved for therapy, none of these radioligands were clinically validated.
Figure 10. Bcr-Abl targeted radiotracers. (a) Chemical structure of [11C]imatinib; (b) Radiosynthesis
of [18F]SKI696; (c) Chemical structure of [124I]SKI230; (d) Transverse and coronal [18F]SKI696 PET images
of K562 tumor bearing mice at 1 h p.i. (A,B) and 2 h p.i. (C,D) (this research was originally published
in [127] © the Society of Nuclear Medicine and Molecular Imaging, Inc.).
Figure 9. Other radiolabeled multikinase inhibitors targeting VEGFRs. (a) Radiosynthesis of
[carbonyl-11C]sorafenib; (b) Chemical structure of [methyl-11C]sorafenib; (c) [18F]FDG (A) and
[methyl-11C]sorafenib coronal PET images of RXF393 tumor bearing mice (reprinted from [115]
with permission from Elsevier); (d) Chemical structures of additional recently radiosynthesized
multikinase inhibitors targeting VEGFRs.
Imatinib was the first FDA approved kinase inhibitor and initially approved for the treatment
of CML via Bcr-Abl inhibiton. Imatinib also strongly inhibits c-Kit and PDGFR but shows a
fairly good selectivity outside those targets. The DFG-out binding mode of imatinib was a
serendipitous discovery that served as a foundation in the development of inhibitors targeting
inactive kinase conformations [1]. The initial [11C]imatinib PET study delineated biodistribution
and pharmacokinetic data including low brain uptake compatible with active transporter efflux
in baboons, but was not followed by advanced preclinical experiments using tumor-bearing mice
(Figure 10a) [124]. Instead, [18F]SKI696, a fluorinated imatinib surrogate [125], and [124I]SKI230 [126],
were both developed and shown to accumulate moderately in Bcr-Abl overexpressing K562 cell
xenografts (human immortalized myelogenous leukemia cells, Figure 10b–d) [127]. [18F]SKI696 was
synthesized by reacting 1-bromo-2-[18F]fluoroethane with precursor 32. Despite detectable tumor
uptake (1.2% ˘0.4%ID/g, 60 min p.i.), unfavorable tumor-to-background ratio and high radioactivity
uptake in the abdominal cavity were observed (Figure 10d). More recently, preclinical imaging of
[18F]SKI696 (also identified as [18F]STI-575) was revisited by Peng and colleagues [128]. This study
confirmed the previous K562 xenografts findings and showed similarly low uptake profile in c-Kit
expressing U87WT tumor-bearing mice. In the absence of a clear imaging rationale for Bcr-Abl
targeting in CML, and a potentially limited applicability to c-Kit-positive gastrointestinal stromal
tumor (GIST), for which imatinib is also approved for therapy, none of these radioligands were
clinically validated.
22011
Molecules 2015,20, 22000–22027
Molecules 2015, 20, page–page
12
Figure 9. Other radiolabeled multikinase inhibitors targeting VEGFRs. (a) Radiosynthesis of [carbonyl-
11C]sorafenib; (b) Chemical structure of [methyl-11C]sorafenib; (c) [18F]FDG (A) and [methyl-11C]sorafenib
coronal PET images of RXF393 tumor bearing mice (reprinted from [115] with permission from Elsevier);
(d) Chemical structures of additional recently radiosynthesized multikinase inhibitors targeting VEGFRs.
Imatinib was the first FDA approved kinase inhibitor and initially approved for the treatment of
CML via Bcr-Abl inhibiton. Imatinib also strongly inhibits c-Kit and PDGFR but shows a fairly good
selectivity outside those targets. The DFG-out binding mode of imatinib was a serendipitous discovery
that served as a foundation in the development of inhibitors targeting inactive kinase conformations [1].
The initial [11C]imatinib PET study delineated biodistribution and pharmacokinetic data including
low brain uptake compatible with active transporter efflux in baboons, but was not followed by
advanced preclinical experiments using tumor-bearing mice (Figure 10a) [124]. Instead, [18F]SKI696,
a fluorinated imatinib surrogate [125], and [124I]SKI230 [126], were both developed and shown to
accumulate moderately in Bcr-Abl overexpressing K562 cell xenografts (human immortalized
myelogenous leukemia cells, Figure 10b–d) [127]. [18F]SKI696 was synthesized by reacting 1-bromo-
2-[18F]fluoroethane with precursor 32. Despite detectable tumor uptake (1.2% ± 0.4%ID/g, 60 min p.i.),
unfavorable tumor-to-background ratio and high radioactivity uptake in the abdominal cavity were
observed (Figure 10d). More recently, preclinical imaging of [18F]SKI696 (also identified as [18F]STI-575)
was revisited by Peng and colleagues [128]. This study confirmed the previous K562 xenografts
findings and showed similarly low uptake profile in c-Kit expressing U87WT tumor-bearing mice.
In the absence of a clear imaging rationale for Bcr-Abl targeting in CML, and a potentially limited
applicability to c-Kit-positive gastrointestinal stromal tumor (GIST), for which imatinib is also
approved for therapy, none of these radioligands were clinically validated.
Figure 10. Bcr-Abl targeted radiotracers. (a) Chemical structure of [11C]imatinib; (b) Radiosynthesis
of [18F]SKI696; (c) Chemical structure of [124I]SKI230; (d) Transverse and coronal [18F]SKI696 PET images
of K562 tumor bearing mice at 1 h p.i. (A,B) and 2 h p.i. (C,D) (this research was originally published
in [127] © the Society of Nuclear Medicine and Molecular Imaging, Inc.).
Figure 10. Bcr-Abl targeted radiotracers. (a) Chemical structure of [11C]imatinib; (b) Radiosynthesis
of [18F]SKI696; (c) Chemical structure of [124I]SKI230; (d) Transverse and coronal [18F]SKI696 PET
images of K562 tumor bearing mice at 1 h p.i. (A,B) and 2 h p.i. (C,D) (this research was originally
published in [127] © the Society of Nuclear Medicine and Molecular Imaging, Inc.).
Type-I Abl inhibitors such as dasatinib (Sprycelr, Bristol-Myers Squibb) have also been
approved for clinical applications. Contrary to imatinib, dasatinib is a highly promiscuous
inhibitor, especially within the tyrosine kinase class [27,28]. [18F]SKI249380, a potent 18F-fluorodeoxy
dasatinib derivative, was initially validated in K562 tumor xenografts and in a dosimetry study
(Figure 11a) [129,130]. Tumor uptakes were similar to [18F]SKI696 using the same preclinical
paradigm (~1%ID/g). However, [18F]SKI249380 is currently under investigation in a clinical trial
for potential diagnostic imaging in a wide range of solid tumors [131]. Although the promiscuity
of [18F]SKI249380, which likely mimics that of dasatinib, may lead to a wider applicability of the
tracer, it may likely be challenging in this context to extrapolate reliable data regarding specific kinase
biomarkers. However, such an encompassing study design may help to experimentally identify select
cases where a multi-targeted probe like [18F]SKI249380 can be applicable. The results of this trial will
provide a first insight into the clinical potential of radiolabeled promiscuous kinase inhibitors for
diagnostic PET imaging.
Molecules 2015, 20, page–page
13
Type-I Abl inhibitors such as dasatinib (Sprycel®, Bristol-Myers Squibb) have also been approved for
clinical applications. Contrary to imatinib, dasatinib is a highly promiscuous inhibitor, especially
within the tyrosine kinase class [27,28]. [18F]SKI249380, a potent 18F-fluorodeoxy dasatinib derivative,
was initially validated in K562 tumor xenografts and in a dosimetry study (Figure 11a) [129,130].
Tumor uptakes were similar to [18F]SKI696 using the same preclinical paradigm (~1%ID/g). However,
[18F]SKI249380 is currently under investigation in a clinical trial for potential diagnostic imaging in a
wide range of solid tumors [131]. Although the promiscuity of [18F]SKI249380, which likely mimics that
of dasatinib, may lead to a wider applicability of the tracer, it may likely be challenging in this context
to extrapolate reliable data regarding specific kinase biomarkers. However, such an encompassing
study design may help to experimentally identify select cases where a multi-targeted probe like
[18F]SKI249380 can be applicable. The results of this trial will provide a first insight into the clinical
potential of radiolabeled promiscuous kinase inhibitors for diagnostic PET imaging.
Figure 11. (a) Chemical structure of dasatinib and [18F]SKI249380 (deoxy-[18F]fluoro-dasatinib); (b)
Axial PET images representing the brain uptake for different nanoformulations of [18F]SKI249380 in a
PDGFR-driven mouse model of high grade glioma. * p = 0.002, ** p = 0.014, *** p = 0.016. (reprinted
from [132], with permission from Elsevier).
This 18F-dasatinib derivative has also been shown to favorably image CNS tumors in a PDGFB
driven model of high-grade glioma using a nanocarrier-encapsulated formulation platform (4.9% ±
0.9%ID/g, 60 min p.i. with micelle encapsulation, and 3.5% ± 0.6%ID/g, 60 min p.i. with liposome
encapsulation Figure 11b) [132]. A liposome nanoparticle encapsulation strategy was also demonstrated
to be favorable by Medina and colleagues [100] in the imaging of A431 xenografts with the EGFR
inhibitor [124I]SKI 243. Those results constitute some of the highest tumor uptakes observed in
preclinical models so far with radiolabled kinase inhibitors. Hence, liposomal or micellar radiotracer
delivery may be a favorable avenue for radiolabled kinase inhibitors imaging [133].
Other efforts for the in vivo imaging of tyrosine kinases have been directed at the tropomyosin
receptor kinases family (TrkA, TrkB and TrkC) (Figure 12) [134,135], the mesenchymal-epithelial
transition receptor (MET) [136] and Ephrin type-B receptor 4 (EphB4) (Figure 13) [137].
Figure 12. Trk radiolabeled inhibitors. (a) Chemical structures of [11C]GW441756 and [18F]38; (b,c) In vitro
validation of [11C]GW441756 using rat brain and TrkB-expressing human neuroblastoma cryosections
(reprinted with permission from [134] © 2015 American Chemical Society); (d) PET images of
[11C]GW441756 in the rat brain; (e) Chemical structure of [18F]40.
Figure 11. (a) Chemical structure of dasatinib and [18F]SKI249380 (deoxy-[18F]fluoro-dasatinib);
(b) Axial PET images representing the brain uptake for different nanoformulations of [18F]SKI249380
in a PDGFR-driven mouse model of high grade glioma. * p= 0.002, ** p= 0.014, *** p= 0.016. (reprinted
from [132], with permission from Elsevier).
This 18F-dasatinib derivative has also been shown to favorably image CNS tumors in a
PDGFB driven model of high-grade glioma using a nanocarrier-encapsulated formulation platform
(4.9% ˘0.9%ID/g, 60 min p.i. with micelle encapsulation, and 3.5% ˘0.6%ID/g, 60 min p.i. with
liposome encapsulation Figure 11b) [132]. A liposome nanoparticle encapsulation strategy was also
demonstrated to be favorable by Medina and colleagues [100] in the imaging of A431 xenografts
with the EGFR inhibitor [124I]SKI 243. Those results constitute some of the highest tumor uptakes
22012
Molecules 2015,20, 22000–22027
observed in preclinical models so far with radiolabled kinase inhibitors. Hence, liposomal or micellar
radiotracer delivery may be a favorable avenue for radiolabled kinase inhibitors imaging [133].
Other efforts for the in vivo imaging of tyrosine kinases have been directed at the tropomyosin
receptor kinases family (TrkA, TrkB and TrkC) (Figure 12) [134,135], the mesenchymal-epithelial
transition receptor (MET) [136] and Ephrin type-B receptor 4 (EphB4) (Figure 13) [137].
Molecules 2015, 20, page–page
13
Type-I Abl inhibitors such as dasatinib (Sprycel®, Bristol-Myers Squibb) have also been approved for
clinical applications. Contrary to imatinib, dasatinib is a highly promiscuous inhibitor, especially
within the tyrosine kinase class [27,28]. [18F]SKI249380, a potent 18F-fluorodeoxy dasatinib derivative,
was initially validated in K562 tumor xenografts and in a dosimetry study (Figure 11a) [129,130].
Tumor uptakes were similar to [18F]SKI696 using the same preclinical paradigm (~1%ID/g). However,
[18F]SKI249380 is currently under investigation in a clinical trial for potential diagnostic imaging in a
wide range of solid tumors [131]. Although the promiscuity of [18F]SKI249380, which likely mimics that
of dasatinib, may lead to a wider applicability of the tracer, it may likely be challenging in this context
to extrapolate reliable data regarding specific kinase biomarkers. However, such an encompassing
study design may help to experimentally identify select cases where a multi-targeted probe like
[18F]SKI249380 can be applicable. The results of this trial will provide a first insight into the clinical
potential of radiolabeled promiscuous kinase inhibitors for diagnostic PET imaging.
Figure 11. (a) Chemical structure of dasatinib and [18F]SKI249380 (deoxy-[18F]fluoro-dasatinib); (b)
Axial PET images representing the brain uptake for different nanoformulations of [18F]SKI249380 in a
PDGFR-driven mouse model of high grade glioma. * p = 0.002, ** p = 0.014, *** p = 0.016. (reprinted
from [132], with permission from Elsevier).
This 18F-dasatinib derivative has also been shown to favorably image CNS tumors in a PDGFB
driven model of high-grade glioma using a nanocarrier-encapsulated formulation platform (4.9% ±
0.9%ID/g, 60 min p.i. with micelle encapsulation, and 3.5% ± 0.6%ID/g, 60 min p.i. with liposome
encapsulation Figure 11b) [132]. A liposome nanoparticle encapsulation strategy was also demonstrated
to be favorable by Medina and colleagues [100] in the imaging of A431 xenografts with the EGFR
inhibitor [124I]SKI 243. Those results constitute some of the highest tumor uptakes observed in
preclinical models so far with radiolabled kinase inhibitors. Hence, liposomal or micellar radiotracer
delivery may be a favorable avenue for radiolabled kinase inhibitors imaging [133].
Other efforts for the in vivo imaging of tyrosine kinases have been directed at the tropomyosin
receptor kinases family (TrkA, TrkB and TrkC) (Figure 12) [134,135], the mesenchymal-epithelial
transition receptor (MET) [136] and Ephrin type-B receptor 4 (EphB4) (Figure 13) [137].
Figure 12. Trk radiolabeled inhibitors. (a) Chemical structures of [11C]GW441756 and [18F]38; (b,c) In vitro
validation of [11C]GW441756 using rat brain and TrkB-expressing human neuroblastoma cryosections
(reprinted with permission from [134] © 2015 American Chemical Society); (d) PET images of
[11C]GW441756 in the rat brain; (e) Chemical structure of [18F]40.
Figure 12. Trk radiolabeled inhibitors. (a) Chemical structures of [11C]GW441756 and [18F]38;
(b,c)In vitro validation of [11C]GW441756 using rat brain and TrkB-expressing human neuroblastoma
cryosections (reprinted with permission from [134] © 2015 American Chemical Society); (d) PET
images of [11C]GW441756 in the rat brain; (e) Chemical structure of [18F]40.
Molecules 2015, 20, page–page
14
Figure 13. Chemical structures of other radiolabeled tyrosine kinase inhibitors.
The TrkA/B/C family consists of three structurally analogous tyrosine kinases with pivotal
significance in embryonic development and post-natal maintenance of the mammalian nervous
system as well as in neurodegenerative diseases [138]. Those receptors are associated with aggressive
tumor phenotypes in a set of neurogenic and non-neurogenic neoplasms including neuroblastoma
and pancreatic cancer and constitute an emerging kinase class targeted in clinical trials using kinase
inhibitors [139]. Two recent studies have reported radiolabeled inhibitors targeting TrkA/B/C for dual
application for CNS Trk expression assessment and potentially tumor imaging. A series of derivatives
of the highly potent and selective pan-Trk inhibitor GW441756 were designed and synthesized.
Two radiotracers ensuing from a structure activity relationship study were labeled and evaluated
(Figure 12a) [134]. Both tracers showed highly specific accumulation in rat brain and TrkB-expressing
human neuroblastoma sections in vitro (Figure 12b,c). While [11C]GW441756 displayed a favorable
in vivo distribution in healthy rats with good brain penetration (SUVmax = 2.0, Figure 12d), [18F]38 was
unstable and extensively defluorinated in vivo. No reduction in brain accumulation of [11C]GW441756
under the tested blocking condition was observed. In a distinct study, a fluorinated derivative of the
selective diaminopyrimidine dual CSF-1R/pan-Trk inhibitor GW2580 was identified and radiolabeled
([18F]40, Figure 12e) [135]. Although no in vivo data were provided, [18F]40 was shown to retain the
pronounced selectivity of GW2580 [27,28] and constitutes one of the most selective radiolabeled
kinase inhibitors identified to date (no additional activity at 3 μM on a 442 kinases panel). Both
[11C]GW441756 and [18F]40 are lead tracers for Trk PET imaging.
Single tracers for EphB4 and MET have also been reported for tumor imaging. Compound [18F]41
did not show tumor uptake in EphB4-overexpressing tumors vs. control despite good metabolic
stability. The authors of this study have suggested that the lack of tumor uptake could be due to a
lack of selectivity combined with insufficient potency [136]. The second generation MET inhibitor
SU11274 was also labeled and evaluated in a MET-positive NCI-H1975 xenograft. The tumor uptakes
were only 1.35-fold higher compared to the uptake in a MET-negative NCI-H520 xenograft model and
overall very low (<0.4 SUV). This tracer was not further investigated since its original publication [136].
2.2. Radiolabeled Serine/Threonine Kinase Inhibitors
Serine and threonine kinases include various potential and validated therapeutic targets with
key roles in human cancers (e.g., B-Raf [140], PIM kinases [141], aurora kinases [142], mammalian
target of rapamycin (mTOR) [143]) and in the pathophysiology of neurodegenerative conditions
including Alzheimer’s and Parkinson’s diseases (e.g., glycogen synthase kinase-3 (GSK-3) [144]).
Early work towards radiolabeled serine/threonine kinases were centered mostly on poorly selective,
staurosporine-based protein kinase A, B and C inhibitors [81,145–148]. Currently, GSK-3α/β constitute
the most pursued serine/threonine targets with different research groups developing radiolabeled
inhibitors for CNS imaging. Early work by Vasdev et al. led to the synthesis of [methyl-11C]AR-
A014418 [149]. Despite poor brain penetration of [methyl-11C]AR-A014418, observed both at baseline
and in the presence of a P-gp inhibitor [31,149], a second radiosynthesis, used as a platform to establish
a novel approach to unsymmetrical [carbonyl-11C]ureas using [11C]CO2, was devised and led to the
synthesis of [carbonyl-11C]AR-A014418 (Figure 14a) [150]. This approach relies on the reaction of a
reactive amine with [11C]CO2 followed by dehydration using POCl3 and subsequent addition of a
second amine component [151].
Figure 13. Chemical structures of other radiolabeled tyrosine kinase inhibitors.
The TrkA/B/C family consists of three structurally analogous tyrosine kinases with pivotal
significance in embryonic development and post-natal maintenance of the mammalian nervous
system as well as in neurodegenerative diseases [138]. Those receptors are associated with aggressive
tumor phenotypes in a set of neurogenic and non-neurogenic neoplasms including neuroblastoma
and pancreatic cancer and constitute an emerging kinase class targeted in clinical trials using kinase
inhibitors [139]. Two recent studies have reported radiolabeled inhibitors targeting TrkA/B/C
for dual application for CNS Trk expression assessment and potentially tumor imaging. A series
of derivatives of the highly potent and selective pan-Trk inhibitor GW441756 were designed and
synthesized. Two radiotracers ensuing from a structure activity relationship study were labeled
and evaluated (Figure 12a) [134]. Both tracers showed highly specific accumulation in rat brain
and TrkB-expressing human neuroblastoma sections in vitro (Figure 12b,c). While [11C]GW441756
displayed a favorable in vivo distribution in healthy rats with good brain penetration
(SUVmax = 2.0
,
Figure 12d), [18F]38 was unstable and extensively defluorinated in vivo. No reduction in brain
accumulation of [11C]GW441756 under the tested blocking condition was observed. In a distinct
study, a fluorinated derivative of the selective diaminopyrimidine dual CSF-1R/pan-Trk inhibitor
GW2580 was identified and radiolabeled ([18F]40, Figure 12e) [135]. Although no in vivo data were
provided, [18F]40 was shown to retain the pronounced selectivity of GW2580 [27,28] and constitutes
one of the most selective radiolabeled kinase inhibitors identified to date (no additional activity at
3µM on a 442 kinases panel). Both [11C]GW441756 and [18F]40 are lead tracers for Trk PET imaging.
Single tracers for EphB4 and MET have also been reported for tumor imaging. Compound
[18F]41 did not show tumor uptake in EphB4-overexpressing tumors vs. control despite good
metabolic stability. The authors of this study have suggested that the lack of tumor uptake could
be due to a lack of selectivity combined with insufficient potency [136]. The second generation MET
inhibitor SU11274 was also labeled and evaluated in a MET-positive NCI-H1975 xenograft. The tumor
22013
Molecules 2015,20, 22000–22027
uptakes were only 1.35-fold higher compared to the uptake in a MET-negative NCI-H520 xenograft
model and overall very low (<0.4 SUV). This tracer was not further investigated since its original
publication [136].
2.2. Radiolabeled Serine/Threonine Kinase Inhibitors
Serine and threonine kinases include various potential and validated therapeutic targets with
key roles in human cancers (e.g., B-Raf [140], PIM kinases [141], aurora kinases [142], mammalian
target of rapamycin (mTOR) [143]) and in the pathophysiology of neurodegenerative conditions
including Alzheimer’s and Parkinson’s diseases (e.g., glycogen synthase kinase-3 (GSK-3) [144]).
Early work towards radiolabeled serine/threonine kinases were centered mostly on poorly selective,
staurosporine-based protein kinase A, B and C inhibitors [81,145148]. Currently, GSK-3α/β
constitute the most pursued serine/threonine targets with different research groups developing
radiolabeled inhibitors for CNS imaging. Early work by Vasdev et al. led to the synthesis of
[methyl-11C]AR-A014418 [149]. Despite poor brain penetration of [methyl-11C]AR-A014418, observed
both at baseline and in the presence of a P-gp inhibitor [31,149], a second radiosynthesis, used as a
platform to establish a novel approach to unsymmetrical [carbonyl-11C]ureas using [11C]CO2, was
devised and led to the synthesis of [carbonyl-11C]AR-A014418 (Figure 14a) [150]. This approach
relies on the reaction of a reactive amine with [11C]CO2followed by dehydration using POCl3and
subsequent addition of a second amine component [151].
In a distinct study, Cole et al., reported the synthesis and in vivo evaluation of [11C]PyrATP-1,
a high affinity selective pyrazine-based GSK-3βinhibitor [152]. In spite of apparently favorable
properties for brain penetration, [11C]PyrATP-1 failed to display any appreciable brain uptake in
both rodents and rhesus macaques. In a recent study, Kumata and colleagues have investigated a
series a three novel GSK-3 β11C-labeled inhibitors corresponding to the different oxidation states of
a benzofuran scaffold, specifically [11C]methylsulfanyl ([11C]45), [11C]methylsulfunyl ([11C]46) and
[11C]methylsulfonyl ([11C]47) derivatives [153]. This study found a 2-fold higher brain penetration
for two out of the three tested tracers in a mouse model of cold water stress associated with increased
GSK-3βexpression. The study also provided a preliminary validation of the potential of those tracers
for imaging ([11C]45 and ([11C]47). In addition to those studies, two groups have worked towards the
synthesis and in vivo validation of [11C]SB-216763, the radiolabeled version of a tool GSK-3 maleimide
inhibitor. The synthesis of [11C]SB-216763 was first reported using a two steps approach making use
of a maleic anhydride precursor followed by conversion into the corresponding maleimide once the
11C-methylation was completed on the indole moiety (Figure 14b) [154]. Although reported RCYs
with this approach were 20%–30% (d.c.), no in vivo data were provided as part of this study. Then,
Li et al., provided a detailed radiosynthetic investigation which led to the development of a more
reliable route towards [11C]SB-216763 in light of problems reproducing the previous synthetic
method due to a [3 + 3]-sigmatropic shift observed when starting from precursor 48 as previously
described [155]. Therefore, [11C]SB-216763 was obtained from the protected precursor 50. Although
this method only delivers [11C]SB-216763 in 1% RCY (185–200 MBq non d.c., 150–350 GBq/µmol), this
was sufficient to carry out a detailed preclinical imaging study which revealed that the radiotracer
shows good brain penetration in both rodent and primate (SUVmax rodent = 2.5, 3 min p.i. and
SUVmax rodent = 1.9, 5 min p.i., Figure 14c). No blocking data were provided. Nevertheless,
[11C]SB-216763 is currently the primary lead for GSK-3 PET imaging.
22014
Molecules 2015,20, 22000–22027
Molecules 2015, 20, page–page
15
In a distinct study, Cole et al., reported the synthesis and in vivo evaluation of [11C]PyrATP-1, a
high affinity selective pyrazine-based GSK-3β inhibitor [152]. In spite of apparently favorable properties
for brain penetration, [11C]PyrATP-1 failed to display any appreciable brain uptake in both rodents
and rhesus macaques. In a recent study, Kumata and colleagues have investigated a series a three
novel GSK-3 β11C-labeled inhibitors corresponding to the different oxidation states of a benzofuran
scaffold, specifically [11C]methylsulfanyl ([11C]45), [11C]methylsulfunyl ([11C]46) and [11C]methylsulfonyl
([11C]47) derivatives [153]. This study found a 2-fold higher brain penetration for two out of the three
tested tracers in a mouse model of cold water stress associated with increased GSK-3β expression.
The study also provided a preliminary validation of the potential of those tracers for imaging ([11C]45
and ([11C]47). In addition to those studies, two groups have worked towards the synthesis and in vivo
validation of [11C]SB-216763, the radiolabeled version of a tool GSK-3 maleimide inhibitor. The synthesis
of [11C]SB-216763 was first reported using a two steps approach making use of a maleic anhydride
precursor followed by conversion into the corresponding maleimide once the 11C-methylation was
completed on the indole moiety (Figure 14b) [154]. Although reported RCYs with this approach were
20%–30% (d.c.), no in vivo data were provided as part of this study. Then, Li et al., provided a detailed
radiosynthetic investigation which led to the development of a more reliable route towards [11C]SB-216763
in light of problems reproducing the previous synthetic method due to a [3 + 3]-sigmatropic shift
observed when starting from precursor 48 as previously described [155]. Therefore, [11C]SB-216763
was obtained from the protected precursor 50. Although this method only delivers [11C]SB-216763 in
1% RCY (185–200 MBq non d.c., 150–350 GBq/μmol), this was sufficient to carry out a detailed preclinical
imaging study which revealed that the radiotracer shows good brain penetration in both rodent and
primate (SUVmax rodent = 2.5, 3 min p.i. and SUVmax rodent = 1.9, 5 min p.i., Figure 14c). No blocking data
were provided. Nevertheless, [11C]SB-216763 is currently the primary lead for GSK-3 PET imaging.
Figure 14. (a) Chemical structures of representative novel GSK-3 radiolabeled inhibitors; (b) Two
radiosynthetic approaches for the synthesis of [11C]SB-216763; (c) Rodent (A) and primate (B) brain PET
images of [11C]SB-216763 (reprinted with permission from [155] © 2015 American Chemical Society).
In recent years, a number of other serine/threonine kinase targets have been investigated using
radiolabeled inhibitors and characterized in vivo, including rho-kinases (ROCKs), cyclin-dependent
kinase 4 (CDK4), B-Raf and most recently aurora kinase A (AURKA). The radiosynthesis of the ROCKs
inhibitor [11C]51 (N-[11C]methyl-hydroxyfasudil), a derivative from the potent Rho-kinases fasudil,
was first reported by Valdivia et al. (Figure 15a) [156]. This radiotracer was then used in a preclinical
in vivo model for brain imaging and in a preliminary in vitro study measuring Rho kinase activity in
hypertrophied cardiomyocytes [157,158]. Although [11C]51 displayed limited applicability for brain
imaging due to poor brain penetration [157], it was shown to correlate with Rho-kinase expression
in vitro in hypertrophied cardiomyocytes and may be appropriate in this context [158]. The 124I-labeled
Cdk4 inhibitors [124I]52 and [124I]53 were also synthesized and characterized as potential tumor
imaging agents due to the role of Cdks in cell proliferation in cancer. Unfortunately, negligible tumor
uptake was observed in FaDu tumor xenografts [159]. The clinical validation of other serine/threonine
kinases such as B-raf with driver mutation V600E has motivated the development and subsequent
Figure 14. (a) Chemical structures of representative novel GSK-3 radiolabeled inhibitors;
(b) Two radiosynthetic approaches for the synthesis of [11C]SB-216763; (c) Rodent (A) and primate
(B) brain PET images of [11C]SB-216763 (reprinted with permission from [155] © 2015 American
Chemical Society).
In recent years, a number of other serine/threonine kinase targets have been investigated using
radiolabeled inhibitors and characterized in vivo, including rho-kinases (ROCKs), cyclin-dependent
kinase 4 (CDK4), B-Raf and most recently aurora kinase A (AURKA). The radiosynthesis of the
ROCKs inhibitor [11C]51 (N-[11C]methyl-hydroxyfasudil), a derivative from the potent Rho-kinases
fasudil, was first reported by Valdivia et al. (Figure 15a) [156]. This radiotracer was then used in a
preclinical in vivo model for brain imaging and in a preliminary in vitro study measuring Rho kinase
activity in hypertrophied cardiomyocytes [157,158]. Although [11C]51 displayed limited applicability
for brain imaging due to poor brain penetration [157], it was shown to correlate with Rho-kinase
expression in vitro in hypertrophied cardiomyocytes and may be appropriate in this context [158]. The
124I-labeled Cdk4 inhibitors [124I]52 and [124I]53 were also synthesized and characterized as potential
tumor imaging agents due to the role of Cdks in cell proliferation in cancer. Unfortunately, negligible
tumor uptake was observed in FaDu tumor xenografts [159]. The clinical validation of other
serine/threonine kinases such as B-raf with driver mutation V600E has motivated the development
and subsequent PET biodistribution study in mice of [carbonyl-11C]CEP-32496 ([11C]54) [160], an
isotopologue of an inhibitor currently in clinical development and more recently [11C]vemurafenib
(Figure 15b) [161]. The inhibitor [carbonyl-11 C]CEP-32496 was synthesized using [11C]phosgene and
brain accumulation was shown to be significantly influenced by ABCB1/ABCG2 using wild-type
and Abcb1a/1b;Abcg2´{´ mice (8-fold higher accumulation in Abcb1a/1b;Abcg2´{´ vs. wild-type
mice) (Figure 15c) [160]. Most recently, Goos et al., described the first radiolabeled aurora kinase A
inhibitor using the selective phase 3 inhibitor alisertib (MLN8237) (Figure 15d) [162]. The radiotracer
was characterized in three tumor models (A431, HCT116 and MKN45) and Abcb1a/1b´{´ mice.
[11C]Alisertib accumulated in xenografts according to AURKA expression levels with highest albeit
still modest tumor uptake and tumor-to-background ratios in A431 tumor (1.9% ˘0.2%ID/g at 25 p.i.,
2.3 ˘0.8 ratio at 90 min p.i.). In addition, both cellular assays using [3H]alisertib and PET imaging in
Abcb1a/1´{´ mice with [11C]alisertib indicated that alisertib is a P-gp substrate.
Multiple additional serine/threonine kinase inhibitors have been recently radiolabeled. These
have been aimed at both previously investigated and novel targets including Rho-kinases, Cdk2,
PI3K, mTOR, p38αmitogen-activated protein kinase, PKC and PIM1 (Figure 16) [163171]. However,
these reports have only described radiosynthetic work without in vitro or in vivo validations so far.
22015
Molecules 2015,20, 22000–22027
Molecules 2015, 20, page–page
16
PET biodistribution study in mice of [carbonyl-11C]CEP-32496 ([11C]54) [160], an isotopologue of an
inhibitor currently in clinical development and more recently [11C]vemurafenib (Figure 15b) [161].
The inhibitor [carbonyl-11C]CEP-32496 was synthesized using [11C]phosgene and brain accumulation
was shown to be significantly influenced by ABCB1/ABCG2 using wild-type and Abcb1a/1b;Abcg2/
mice (8-fold higher accumulation in Abcb1a/1b;Abcg2/ vs. wild-type mice) (Figure 15c) [160]. Most
recently, Goos et al., described the first radiolabeled aurora kinase A inhibitor using the selective
phase 3 inhibitor alisertib (MLN8237) (Figure 15d) [162]. The radiotracer was characterized in three
tumor models (A431, HCT116 and MKN45) and Abcb1a/1b/ mice. [11C]Alisertib accumulated in
xenografts according to AURKA expression levels with highest albeit still modest tumor uptake and
tumor-to-background ratios in A431 tumor (1.9% ± 0.2%ID/g at 25 p.i., 2.3 ± 0.8 ratio at 90 min p.i.).
In addition, both cellular assays using [3H]alisertib and PET imaging in Abcb1a/1/ mice with [11C]alisertib
indicated that alisertib is a P-gp substrate.
Figure 15. (a) Chemical structures of recently identified and evaluated radiolabeked serine/threonine
kinase inhibitors; (b) Chemical structure of [carbonyl-11C]CEP-32496; (c) Brain PET images of [carbonyl-
11C]CEP-32496 in wild type (top panel) and P-gp/BCRP knockout mice (bottom panel) (reprinted from [160]
© 2014, with permission from Elsevier); (d) Chemical structure of [11C]alisertib.
Multiple additional serine/threonine kinase inhibitors have been recently radiolabeled. These
have been aimed at both previously investigated and novel targets including Rho-kinases, Cdk2, PI3K,
mTOR, p38α mitogen-activated protein kinase, PKC and PIM1 (Figure 16) [163–171]. However, these
reports have only described radiosynthetic work without in vitro or in vivo validations so far.
Figure 16. Chemical structures of recently synthesized serine/threonine kinase inhibitors.
3. Concluding Remarks
Significant progress and diversification has been achieved in recent years in the field of radiolabeled
kinase inhibitors for PET imaging. Recent research has illustrated advances towards prospective
imaging applications including: (1) the clinical validation of target expression as a potential predictor
of patients’ response to kinase inhibitor treatment using isotopologues of approved inhibitors; (2) the
preclinical validation of tumor or tissue engagement during drug development using advanced
radiolabeled leads and (3) the visualization and quantification of kinases for tumor, CNS and cardiac
imaging with tool inhibitors exclusively developed as radiotracers, outside of a drug development
Figure 15. (a) Chemical structures of recently identified and evaluated radiolabeked serine/threonine
kinase inhibitors; (b) Chemical structure of [carbonyl-11C]CEP-32496; (c) Brain PET images of
[carbonyl-11C]CEP-32496 in wild type (top panel) and P-gp/BCRP knockout mice (bottom panel)
(reprinted from [160] © 2014, with permission from Elsevier); (d) Chemical structure of [11C]alisertib.
Molecules 2015, 20, page–page
16
PET biodistribution study in mice of [carbonyl-11C]CEP-32496 ([11C]54) [160], an isotopologue of an
inhibitor currently in clinical development and more recently [11C]vemurafenib (Figure 15b) [161].
The inhibitor [carbonyl-11C]CEP-32496 was synthesized using [11C]phosgene and brain accumulation
was shown to be significantly influenced by ABCB1/ABCG2 using wild-type and Abcb1a/1b;Abcg2/
mice (8-fold higher accumulation in Abcb1a/1b;Abcg2/ vs. wild-type mice) (Figure 15c) [160]. Most
recently, Goos et al., described the first radiolabeled aurora kinase A inhibitor using the selective
phase 3 inhibitor alisertib (MLN8237) (Figure 15d) [162]. The radiotracer was characterized in three
tumor models (A431, HCT116 and MKN45) and Abcb1a/1b/ mice. [11C]Alisertib accumulated in
xenografts according to AURKA expression levels with highest albeit still modest tumor uptake and
tumor-to-background ratios in A431 tumor (1.9% ± 0.2%ID/g at 25 p.i., 2.3 ± 0.8 ratio at 90 min p.i.).
In addition, both cellular assays using [3H]alisertib and PET imaging in Abcb1a/1/ mice with [11C]alisertib
indicated that alisertib is a P-gp substrate.
Figure 15. (a) Chemical structures of recently identified and evaluated radiolabeked serine/threonine
kinase inhibitors; (b) Chemical structure of [carbonyl-11C]CEP-32496; (c) Brain PET images of [carbonyl-
11C]CEP-32496 in wild type (top panel) and P-gp/BCRP knockout mice (bottom panel) (reprinted from [160]
© 2014, with permission from Elsevier); (d) Chemical structure of [11C]alisertib.
Multiple additional serine/threonine kinase inhibitors have been recently radiolabeled. These
have been aimed at both previously investigated and novel targets including Rho-kinases, Cdk2, PI3K,
mTOR, p38α mitogen-activated protein kinase, PKC and PIM1 (Figure 16) [163–171]. However, these
reports have only described radiosynthetic work without in vitro or in vivo validations so far.
Figure 16. Chemical structures of recently synthesized serine/threonine kinase inhibitors.
3. Concluding Remarks
Significant progress and diversification has been achieved in recent years in the field of radiolabeled
kinase inhibitors for PET imaging. Recent research has illustrated advances towards prospective
imaging applications including: (1) the clinical validation of target expression as a potential predictor
of patients’ response to kinase inhibitor treatment using isotopologues of approved inhibitors; (2) the
preclinical validation of tumor or tissue engagement during drug development using advanced
radiolabeled leads and (3) the visualization and quantification of kinases for tumor, CNS and cardiac
imaging with tool inhibitors exclusively developed as radiotracers, outside of a drug development
Figure 16. Chemical structures of recently synthesized serine/threonine kinase inhibitors.
3. Concluding Remarks
Significant progress and diversification has been achieved in recent years in the field of
radiolabeled kinase inhibitors for PET imaging. Recent research has illustrated advances towards
prospective imaging applications including: (1) the clinical validation of target expression as
a potential predictor of patients’ response to kinase inhibitor treatment using isotopologues of
approved inhibitors; (2) the preclinical validation of tumor or tissue engagement during drug
development using advanced radiolabeled leads and (3) the visualization and quantification
of kinases for tumor, CNS and cardiac imaging with tool inhibitors exclusively developed as
radiotracers, outside of a drug development rationale. Despite promising clinical proof-of-concept
studies, the successful application for topic (1) remains largely to be established. However, the
usefulness of radiolabeled kinase inhibitors for drug development (2) and for the validation of kinase
expression and density in preclinical settings (3) has become increasingly evident in recent years.
So far, only isotopologues of approved inhibitors or advanced leads have reached a clinical
evaluation stage. While the radiolabeling of approved compounds enjoys the obvious advantages of
a facilitated clinical translation, such a strategy presents some shortcomings as illustrated recently.
The potentially negative influence of non-optimized or suboptimal ABCB1/ABCG2 efflux, often
encountered within approved kinase inhibitors, was illustrated preclinically with the example of the
development of [11C]erlotinib despite initial promising imaging results. As far as EGFR imaging
is concerned, reliable PET data should be attainable with novel generations of inhibitors such as
[11C]AZD3759. Such compounds will likely be investigated and should enable both peripheral
and CNS metastasis imaging, although it may also lead to unwanted brain exposure during
peripheral tumor imaging. To avoid problems relating to saturable ABCB1/ABCG2 transporters,
the inclusion of imaging experiments using Abcb1a/1b;Abcg2´{´ mice in early developmental steps
was shown to be a valuable tool. Selectivity is another parameter which is often suboptimal
22016
Molecules 2015,20, 22000–22027
amongst approved inhibitors when considering PET imaging applications. Although a promiscuous
or multi-targeted inhibitor may accumulate in a predefined tumor model, ultimately, tumor
heterogeneity compounded with the lack of selectivity is likely to lead to tumor uptake data
which will be challenging to interpret. Most kinase inhibitors are approved for patients with
tumors overexpressing specific kinases or harboring specific mutations, even though the kinase
inhibitor treatment may lead to cross interactions with other oncologically relevant or non-relevant
kinases. Ultimately in this context, PET imaging with non-selective inhibitors will potentially limit
any useful information that could be gained regarding expression levels specifically related to a
biomarker on which the treatment decision should be made. Whereas the development of inhibitors
for therapy based on the multi-targeted selectivity principle is established and may offer a safe
compromise between efficacy and potential toxicity, this principle is difficult to reconcile with the
microdosing nature of PET imaging experiments [172]. Furthermore, such polypharmacology may
have unpredictable in vivo outcomes at microdose levels due to the wide range of KM ATP values
found among kinases. Selective kinase inhibitors have been identified for various kinase targets and
could constitute a more favorable starting point in the development of radiolabeled inhibitors [173].
Those compounds exist, but are normally not pursued as therapeutics due to a high risk of inefficacy
following mutation events. The wider availability of comprehensive kinase screening should also
facilitate compound selection in terms of selectivity for radiotracer development.
The question of whether or not type II inhibitors tend to exhibit better selectivity than type I
inhibitors has been subjected to some debate recently [7,10,174]. However, an important observation
to bear in mind, as illustrated by the seminal work of Knight and Shokat [26], is that type II inhibitors
are less susceptible to ATP competition ensuing from the high intracellular ATP concentration
(1–5 mM) compared to type I inhibitors. This is mirrored by KM ATP values, which reflect the affinity
for a given kinase towards ATP, and which have been measured for different kinases in both the
activated and inactivated states. More generally, KM ATP values among the kinome vary widely.
Whereas numerous protein kinases show KM ATP values in the low µM range, other display values
in some instances up to a 1000-fold higher (for example mTOR). Those discrepancies are important
for the development of ATP competitive radiolabeled inhibitors as they imply that certain kinase
targets will be intrinsically more challenging to image than others. Type II inhibitors were also
shown to display longer residence times compared to type I inhibitors [175177]. However, it is
also important to recognize that type II inhibitors tend to engage the allosteric site via a strong
H-bond-based interaction with the DFG motif which often results in additional polar moieties
(e.g., ureas, amides). In theory, this can be detrimental for membrane and BBB penetration at
microdosing within the timeframe of PET experiments. While non-ATP competitive inhibitors would
be advantageous for imaging, such compounds a rare and often do not display sufficient affinity
for imaging applications [178,179]. In brief, those challenges relating to binding modes should be
explored in more detail in coming years as they may have a strong impact on the outcome and the
success of radiolabeled kinase inhibitors.
Finally, it is important to recognize that radiolabeled kinase inhibitors for oncological
applications have attained relatively marginal tumor-to-background ratios compared to other
radiotracer classes in the same context (e.g., metabolic radiotracers, radiolabeled antibodies and
radiolabeled peptides). Further efforts in this area should be encouraged and recent studies
illustrating the efficacy of nanoformulations and multivalency derivatization using the longer lived
isotope copper-64 should be evaluated carefully and explored. The intrinsic lipophilicity of small
molecule kinase inhibitors and high uptakes in excretory organs of such compounds is also likely
to limit the application of this class of imaging agents to visualize metastatic lesions, especially in
the liver. Importantly, the large chemical space covered by kinase inhibitors is at the same time an
enormous opportunity and a possible liability for radiotracer development. The knowledge that only
a minuscule fraction of those compounds will be suitable as radiotracers should motivate careful and
stringent selection criteria, with the inclusion of considerations well beyond the important questions
22017
Molecules 2015,20, 22000–22027
of radiolabel position, metabolic susceptibility and favorable basic physico-chemical properties. As
discussed, these considerations should include the nature of the target in relation with KM ATP, the
nature of the binding mode selected, the selectivity profile and the efflux transporter susceptibility
(such as ABCB1/ABCG2). That being said, those challenges will likely be addressed as part of future
advances. The current rapid expansion of kinase inhibitors as a drug class for cancer treatment and
the emerging opportunities outside oncology should offer ample areas to further validate the utility
of radiolabeled kinase inhibitors in coming years.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Zhang, J.; Yang, P.L.; Gray, N.S. Targeting cancer with small molecules kinase inhibitors. Nat. Rev. Cancer
2009,9, 28–39. [CrossRef] [PubMed]
2. Capdeville, R.; Buchdunger, E.; Zimmermann, J.; Matter, A. Glivec (STI571, imatinib), a rationally
developed, targeted anticancer drug. Nat. Rev. Drug Discov. 2002,1, 493–502. [CrossRef] [PubMed]
3. Wu, P.; Nielson, T.E.; Clausen, M.H. FDA-approved small-molecule kinase inhibitors. Trends Pharmacol. Sci.
2015,36, 422–439. [CrossRef] [PubMed]
4. U.S. Department of Health and Human Services. U.S. Food and Drug Administration. Available
online: http://www.fda.gov/NewsEvents/Newsroom/PressAnnouncements/ucm455862.htm (accessed
on 1 October 2015).
5. Van Linden, O.P.J.; Kooistra, A.J.; Leurs, R.; Esch, I.J.P.; de Graaf, C. KLIFS: A knowledge-based structural
database to navigate kinase-ligand interaction space. J. Med. Chem. 2014,57, 249–277. [CrossRef] [PubMed]
6. Bamborough, P.; Drewry, D.; Harper, G.; Smith, G.K.; Schneider, K. Assessment of chemical coverage of
kinome space and its implications for kinase drug discovery. J. Med. Chem. 2008,51, 7898–7914. [CrossRef]
[PubMed]
7. Hu, Y.; Furtmann, N.; Bajorath, J. Current compound coverage of the kinome. J. Med. Chem. 2015,58, 30–40.
[CrossRef] [PubMed]
8. Cui, J.J. A new challenging and promising era of tyrosine kinase inhibitors. ACS Med. Chem. Lett. 2014,5,
272–274. [CrossRef] [PubMed]
9. Taylor, S.S.; Kornev, A.P. Protein kinases: Evolution of dynamic regulatory proteins. Trends Biochem. Sci.
2011,36, 65–77. [CrossRef] [PubMed]
10. Vijayan, R.S.K.; He, P.; Modi, V.; Duong-Ly, K.C.; Ma, H.; Peterson, J.R.; Dunbrack, R.L., Jr.; Levy, R.M.
Conformational analysis of the DFG-out kinase motif and biochemical profiling of structurally validates
type II inhibitors. J. Med. Chem. 2015,58, 466–479. [CrossRef] [PubMed]
11. Barf, T.; Kaptein, A. Irreversible protein kinase inhibitors: Balancing the benefits and risks. J. Med. Chem.
2012,55, 6243–6262. [CrossRef] [PubMed]
12. Liu, Q.; Sabnis, Y.; Zhao, Z.; Zhang, T.; Buhrlage, S.J.; Jones, L.H.; Gray, N.S. Developing irreversible
inhibitor of the protein kinase cysteinome. Chem. Biol. 2013,20, 146–159. [CrossRef] [PubMed]
13. Cox, K.J.; Shomin, C.D.; Ghosh, I. Tinkering outside the kinase ATP box: Allosteric (type IV) and bivalent
(type V) inhibitors of protein kinases. Future Med. Chem. 2010,3, 29–43. [CrossRef] [PubMed]
14. Fang, Z.; Grutter, C.; Rauh, D. Strategies for the selective regulation of kinases with allosteric modulators:
Exploiting structural features. ACS Chem. Biol. 2013,8, 58–70. [CrossRef] [PubMed]
15. Josephs, D.H.; Fisher, D.S.; Spicer, J.; Flanagan, R.J. Clinical pharmacokinetics of tyrosine kinase inhibitors:
Implication for therapeutic drug monitoring. Ther. Drug Monit. 2013,35, 562–587. [CrossRef] [PubMed]
16. Barouch-Bentov, R.; Sauer, K. Mechanism of drug resistance in kinases. Expert Opin. Investig. Drugs 2011,
20, 153–208. [CrossRef] [PubMed]
17. Zhu, A.; Lee, D.; Shim, H. Metabolic PET imaging in cancer detection and therapy response. Semin. Oncol.
2011,38, 55–69. [CrossRef] [PubMed]
18. Matthews, P.M.; Rabiner, E.A.; Passchier, J.; Gunn, R.N. Positron emission tomography molecular imaging
for drug development. Br. Clin. Pharmacol. 2012,73, 175–186. [CrossRef] [PubMed]
19. Cunha, L.; Szigeti, K.; Mathe, D.; Metello, L.F. The role of molecular imaging in modern drug development.
Drug Discov. Today 2014,19, 936–948. [CrossRef] [PubMed]
22018
Molecules 2015,20, 22000–22027
20. Benz, M.R.; Herrmann, K.; Walter, F.; Garon, E.B.; Reckamp, K.L.; Figlin, R.; Phelps, M.E.; Weber, W.A.;
Czernin, J.; Allen-Auerbach, M.S. (18F)-FDG PET/CT for monitoring treatment responses to the epidermal
growth factor receptor inhibitor erlotinib. J. Nucl. Med. 2011,52, 1689–1689. [CrossRef] [PubMed]
21. Van Gool, M.H.; Aukema, T.S.; Hartemink, K.J.; Valdes Olmos, R.A.; van Tinteren, H.; Klomp, H.M.
FDG-PET/CT response evaluation during EGFR-TKI treatment in patients with NSCLC. World J. Radiol.
2014,6, 392–398. [CrossRef] [PubMed]
22. Caldarella, C.; Muoio, B.; Isgro, M.A.; Porfiri, E.; Treglia, G.; Giovanella, L. The role of
fluorine-18-fluorodeoxyglucose positron emission tomography in evaluating the response to
tyrosine-kinase inhibitors in patients with metastatic primary renal cell carcinoma. Radiol. Oncol.
2014,48, 219–227. [CrossRef] [PubMed]
23. Sharma, R.; Aboagye, E. Development of radiotracers for oncology—The interface with pharmacology.
Br. J. Pharmacol. 2011,163, 1565–1585. [CrossRef] [PubMed]
24. Pike, V.W. PET radiotracers: crossing the blood-brain barrier and surviving metabolism. Trends Pharmacol.
Sci. 2009,30, 431–440. [CrossRef] [PubMed]
25. Slobbe, P.; Poot, A.J.; Windhorst, A.D.; van Dongen, G.A.M.S. PET imaging with small-molecule tyrosine
kinase inhibitors: TKI-PET. Drug Discov. Today 2012,17, 1175–1187. [CrossRef] [PubMed]
26. Knight, Z.A.; Shokat, K.M. Features of selective kinase inhibitors. Chem. Biol. 2005,12, 621–637. [CrossRef]
[PubMed]
27. Davis, M.I.; Hunt, J.P.; Herrgard, S.; Ciceri, P.; Wodicka, L.M.; Pallares, G.; Hocker, M.; Treiber, D.K.;
Zarrinkar, P.P. Comprehensive analysis of kinase inhibitor selectivity. Nat. Biotechnol. 2011,29, 1046–1051.
[CrossRef] [PubMed]
28. Anastassiadis, T.; Deacon, S.D.; Devarajan, K.; Ma, H.; Peterson, J.R. Comprehensive assay of kinase
catalytic activity reveals features of kinase inhibitor selectivity. Nat. Biotechnol. 2011,29, 1039–1045.
[CrossRef] [PubMed]
29. Zhang, L.; Villalobos, A.; Beck, E.M.; Bocan, T.; Chappie, T.A.; Chen, L.; Grimwood, S.; Heck, S.D.;
Helal, C.J.; Hou, X.; et al. Design and selection parameters to accelerate the discovery of novel central
nervous system positron emission tomography (PET) ligands and their application in the development of
a novel phosphodiesterase 2A PET ligand. J. Med. Chem. 2013,56, 4568–4579. [CrossRef] [PubMed]
30. Van de Bittner, G.C.; Ricq, E.L.; Hooker, J.M. A philosophy for CNS radiotracer design. Acc. Chem. Res.
2014,47, 3127–3134. [CrossRef] [PubMed]
31. Hicks, J.W.; VanBrocklin, H.F.; Wilson, A.A.; Houle, S.; Vasdev, N. Radiolabeled small molecule protein
kinase inhibitors for imaging with PET or SPECT. Molecules 2010,15, 8260–8278. [CrossRef] [PubMed]
32. Mulholland, G.K.; Zheng, Q.H.; Winkle, W.L.; Carlson, K.A. Synthesis and biodistribution of new C-11 and
F-18 labeled epidermal growth factor receptor ligands. J. Nucl. Med. 1997,38, 529.
33. DeJesus, O.T.; Murali, D.; Flores, L.G.; Converse, A.K.; Dick, D.W.; Oakes, T.R.; Roberts, A.D.; Nickles, R.J.
Synthesis of [18F]-ZD1839 as a PET imaging agent for epidermal growth factor receptors. J. Label. Compd.
Radiopharm. 2003,46, S1–S9.
34. Johnstrom, P.; Fredriksson, A.; Thorell, J.O.; Stone-Elander, S. Synthesis of [methoxy-11C]PD153035, a
selective EGF receptor tyrosine kinase inhibitor. J. Label. Compd. Radiopharm. 1998,41, 623–629. [CrossRef]
35. Fredriksson, A.; Johnstrom, P.; Thorell, J.O.; von Heijne, G.; Hassan, M.; Eksborg, S.; Kogner, P.;
Borgstrom, P.; Ingvar, M.; Stone-Elander, S. In vivo evaluation of the biodistribution of 11C-labeled
PD153035 in rats without and with neuroblastoma implants. Life Sci. 1999,65, 165–174. [CrossRef]
36. Samén, E.; Thorell, J.-O.; Fredriksson, A.; Stone-Elander, S. The tyrosine kinase inhibitor PD153035:
Implication of labeling position on radiometabolites formed in vitro.Nucl. Med. Biol. 2006,33, 1005–1011.
[CrossRef] [PubMed]
37. Ackermann, U.; Tochon-Danguy, H.J.; Scott, A.M. [11 C]AG1478—A potential PET tracer for the molecular
imaging of epidermal growth factor receptor (EGFR). J. Nucl. Med. 2004,45, 112P.
38. Owino, N.O.; Shao, X.; Snyder, S.E. Preparation of a reversible EGFR inhibitor, [11C]CIRC, as a PET
radiotracer for tumor imaging. J. Label. Compd. Radiopharm. 2005,48, S183. [CrossRef]
39. Snyder, S.E.; Sherman, P.S.; Blair, J.B. 4-(3-chloro-4-[18F]fluorophenylamino)-6,7-dimethoxyquinazoline: A
radiolabeled EGF receptor inhibitor for imaging tumor biochemistry with PET. J. Nucl. Med. 2000,41, 233P.
22019
Molecules 2015,20, 22000–22027
40. Bonasera, T.A.; Ortu, G.; Rozen, Y.; Krais, R.; Freedman, N.M.T.; Chisin, R.; Gazit, A.; Levitzki, A.;
Mishani, E. Potential 18F-labeled biomarkers for epidermal growth factor receptor tyrosine kinase. Nucl.
Med. Biol. 2001,28, 359–374. [CrossRef]
41. Dorff, P.N.; Vasdev, N.; O’Neil, J.P.; Nandanan, E.; Gibbs, A.R.; VanBrocklin, H.F. Synthesis of 21-, 31-, and
41-[18F]fluoroanilinoquinazoline. J. Label. Compd. Radiopharm. 2003,46, S117. [CrossRef]
42. Ben-David, I.; Rozen, Y.; Ortu, G.; Mishani, E. Radiosynthesis of ML03, a novel positron emission
tomography biomarker for targeting epidermal growth factor receptor via the labeling synthon:
[11C]acryloyl chloride. Appl. Radiat. Isot. 2003,58, 209–217. [CrossRef]
43. Vasdev, N.; Dorff, P.N.; Gibbs, A.R.; Nandanan, E.; Reid, L.M.; O’Neill, J.P.; VanBrocklin, H.F. Synthesis of
6-acrylamido-4-(2-[18F]fluoroanilino)quinazoline: A prospective irreversible EGFR binding probe. J. Label.
Compd. Radiopharm. 2005,48, 109–115. [CrossRef]
44. Vasdev, N.; Dorff, P.N.; Gibbs, A.R.; Nandanan, E.; Reid, L.M.; O’Neil, J.P.; VanBrocklin, H.F. Synthesis of
6-acrylamido-4-(2-[18F]fluoro-anilino)quinazoline. J. Nucl. Med. 2004,45, 168P.
45. Waldherr, C.; Satyamurthy, N.; Toyokuni, T.; Wang, S.; Mellinghoff, I.; Tran, C.; Stout, D.;
Halpern, B.; Silverman, D.H.; Barrio, J.R.; Phelps, M.E.; Sawyers, C.L.; Czernin, J. Evaluation of
N-{4-[(31-[18F]fluoroethylphenyl)amino]-6-quinazolinyl} acrylamide ([18F]FEQA), a labeled tyrosine kinase
inhibitor, for imaging epidermal growth factor receptor density. J. Nucl. Med. 2003,44, 372P.
46. Abourbeh, G.; Dissoki, S.; Jacobson, O.; Litchi, A.; Ben Daniel, R.; Laki, D.; Levitzki, A.; Mishani, E.
Evaluation of radiolabeled ML04, a putative irreversible inhibitor of epidermal growth factor receptor, as
a bioprobe for PET imaging of EGFR-overexpressing tumors. Nucl. Med. Biol. 2007,34, 55–70. [CrossRef]
[PubMed]
47. Mishani, E.; Abourbeh, G.; Rozen, Y. Novel carbon-11 labeled 4-dimethylamino-but-2-enoic acid
[4-(phenylamino)-quinazoline-6-yl]-amides: Potential PET bioprobes for molecular imaging of
EGFR-positive tumors. Nucl. Med. Biol. 2004,31, 469–476. [CrossRef] [PubMed]
48. Dissoki, S.; Eshet, R.; Billauer, H.; Mishani, E. Modified PEG-anilinoquinazoline derivatives as potential
EGFR PET agents. J. Label. Compd. Radiopharm. 2009,52, 41–52. [CrossRef]
49. Pantaleo, M.; Mishani, E.; Nanni, C.; Landuzzi, L.; Boschi, S.; Nicoletti, G.; Dissoki, S.; Paterini, P.;
Piccaluga, P.; Lodi, F.; et al. Evaluation of modified Evaluation of modified PEG-anilinoquinazoline
derivatives as potential agents for EGFR imaging in cancer by small animal PET. Mol. Imaging Biol. 2010,
12, 616–625. [CrossRef] [PubMed]
50. Dissoki, S.; Aviv, Y.; Laky, D.; Abourbeh, G.; Levitzki, A.; Mishani, E. The effect of the [18F]- PEG group on
tracer qualification of [4-(phenylamino)-quinazoline-6-yl]-amide moiety—An EGFR putative irreversible
inhibitor. Appl. Radiat. Isot. 2007,65, 1140–1151. [CrossRef] [PubMed]
51. Kobus, D.; Giesen, Y.; Ullrich, R.; Backes, H.; Neumaier, B. A fully automated two-step synthesis of an
18F-labelled tyrosine kinase inhibitor for EGFR kinase activity imaging in tumors. Appl. Radiat. Isot. 2009,
67, 1977–1984. [CrossRef] [PubMed]
52. VanBrocklin, H.F.; Vasdev, N.; Dorff, P.N.; O’Neil, J.P.; Taylor, S.E. Metabolism of
[18F]fluoroanilinoquinazolines by human hepatocytes. J. Label. Compd. Radiopharm. 2003,46, S390.
[CrossRef]
53. Chong, C.R.; Janne, P.A. The quest to overcome resistance to EGFR-targeted therapies in cancer. Nat. Med.
2013,19, 1389–1400. [CrossRef] [PubMed]
54. Gazdar, A.F. Activating and resistance mutations of EGFR in non-small-cell lung cancer: Role in clinical
response to EGFR tyrosine kinase inhibitors. Oncogene 2009,28, S24–S31. [CrossRef] [PubMed]
55. Seimbille, Y.; Phelps, M.E.; Czernin, J.; Silverman, D.H.S. Fluoriue-18 labeling of 6,7-disubstituted
anilinoquinazoline derivatives for positron emission tomography (PET) imaging of tyrosine kinase
receptors: Synthesis of F-18-Iressa and related molecular probes. J. Label. Comp. Radiopharm. 2005,48,
829–843. [CrossRef]
56. Su, H.; Seimbille, Y.; Ferl, G.Z.; Bodenstein, C.; Fueger, B.; Kim, K.J.; Hsu, Y.T.; Dubinett, S.M.; Phelps, M.E.;
Czernin, J.; et al. Evaluation of F-18 Gefitinib as a molecular imaging probe for the assessment of the
epidermal growth factor receptor status in malignant tumors. Eur. J. Nucl. Mol. Imaging 2008,35, 1089–1099.
[CrossRef] [PubMed]
22020
Molecules 2015,20, 22000–22027
57. Wang, J.Q.; Gao, M.Z.; Miller, K.D.; Sledge, G.W.; Zheng, Q.H. Synthesis of [C-11]Iressa as a new potential
PET cancer imaging agent for epidermal growth factor receptor tyrosine kinase. Bioorg. Med. Chem. 2006,
16, 4102–4106. [CrossRef] [PubMed]
58. Holt, D.P.; Ravert, H.T.; Dannals, R.F.; Pomper, M.G. Synthesis of [C-11]Gefitinib for imaging epidermal
growth factor receptor tyrosine kinase with positron emission tomography. J. Label. Compd. Radiopharm.
2006,49, 883–888. [CrossRef]
59. Zhang, M.R.; Kumata, K.; Hatori, A.; Takai, N.; Toyohara, J.; Yamasaki, T.; Yanamoto, K.; Yui, J.;
Kawamura, K.; Koike, S.; Ando, K.; Suzuki, K. [11C]Gefitinib ([11 C]Iressa): Radiosynthesis, in vitro uptake,
and in vivo imaging of intact murine fibrosarcoma. Mol. Imaging Biol. 2010,12, 181–191. [CrossRef]
[PubMed]
60. Lappchen, T.; Vlaming, M.L.; Custers, E.; Lub, J.; Sio, C.F.; DeGroot, J.; Steinbach, O.C. Automated synthesis
of [18F]Gefitinib on a modular system. Appl. Radiat. Isot. 2012,70, 205–209. [CrossRef] [PubMed]
61. Elkind, N.B.; Szentpetery, Z.; Apari, A.; Ozvegy-Laczka, C.; Varady, G.; Ujhelly, O.; Szabó, K.; Homolya, L.;
Váradi, A.; Buday, L.; et al. Multidrug transporter ABCG2 prevents tumor cell death induced by the
epidermal growth factor receptor inhibitor Iressa (ZD1839, gefitinib). Cancer Res. 2005,65, 1770–1777.
[CrossRef] [PubMed]
62. Zeng, Q.; Wang, J.; Cheng, Z.; Chen, K.; Johnstrom, P.; Varnas, K.; Li, D.Y.; Yang, Z.F.; Zhang, X. Discovery
and evaluation of clinical candidate AZD3759, a potent, oral active, central nervous system-penetrant,
epidermal growth factor receptor tyrosine kinase inhibitor (EGFR TKI). J. Med. Chem. 2015. [CrossRef]
63. Kawamura, K.; Yamasaki, T.; Yui, J.; Hatori, A.; Konno, F.; Kumata, K.; Irie, T.; Fukumura, T.; Suzuki, K.;
Kanno, I.; et al. In vivo evaluation of p-glycoprotein and breast cancer resistance protein modulation in the
brain using [11C]Gefitinib. Nucl. Med. Biol. 2009,36, 239–246. [CrossRef] [PubMed]
64. Vlaming, M.L.H.; Lappchen, T.; Jansen, H.T.; Kivits, S.; van Driel, A.; van de Steeg, E.; van der Hoorn, J.W.;
Sio, C.F.; Steinbach, O.C.; DeGroot, J. PET-ET imaging with [18F]-gefitinib to measure Abcb1a/1b (P-gp)
and Abcg2 (Bcrp1) mediated drug-drug interaction at the murine blood-brain barrier. Nucl. Med. Biol.
2015. [CrossRef] [PubMed]
65. ClinicalTrials.gov. Available online: https://clinicaltrials.gov/ct2/show/NCT02228369 (accessed on 3
October 2015).
66. Dowell, J.; Minna, J.D.; Kirkpatrick, P. Erlotinib hydrochloride. Nat. Rev. Drug Discov. 2005,4, 13–14.
[CrossRef] [PubMed]
67. Memon, A.A.; Jakobsen, S.; Dagnaes-Hansen, F.; Sorensen, B.S.; Keiding, S.; Nexo, E. Positron emission
tomography (PET) imaging with C-11 -labeled Erlotinib: A micro-PET study on mice with lung tumor
xenografts. Cancer Res. 2009,69, 873–878. [CrossRef] [PubMed]
68. Petrulli, J.R.; Sullivan, J.M.; Zheng, M.Q.; Bennett, D.C.; Charest, J.; Huang, Y.; Morris, E.D.; Contessa, J.N.
Quantitative analysis of [11C]-Erlotinib PET demonstrates specific binding for activating mutations of the
EGFR kinase domain. Neoplasia 2013,15, 1347–1353. [CrossRef] [PubMed]
69. Abourbeh, G.; Itamar, B.; Salnikov, O.; Beltsov, S.; Mishani, E. Identifying Erlotinib-sensitive non-small cell
lung carcinoma tumors in mice using [11C]Erlotinib PET. EJNMMI Res. 2015,5, 4. [CrossRef] [PubMed]
70. Memon, A.A.; Weber, B.; Winterdahl, M.; Jakobsen, S.; Meldgaard,P.; Madsen, H.H.T.; Keiding, S.; Nexo, E.;
Sorensen, B.S. PET imaging of patients with non-small cell lung cancer employing an egf receptor targeting
drug as tracer. Br. J. Cancer 2011,105, 1850–1855. [CrossRef] [PubMed]
71. Bahce, I.; Smit, E.F.; Lubberink, M.; van der Veldt, A.A.M.; Yaqub, M.; Windhorst, A.D.; Schuit, R.C.;
Thunnissen, E.; Heideman, D.A.M.; Postmus, P.E.; et al. Development of [11 C]erlotinib positron emission
tomography for in vivo evaluation of EGF receptor mutation status. Clin. Cancr Res. 2013,19, 183–193.
[CrossRef] [PubMed]
72. Weber, B.; Winterdahl, M.; Memon, A.; Sorensen, B.S.; Keiding, S.; Sorensen, L.; Nexo, E.; Meldgaard, P.
Erlotinib accumulation in brain metastases from non-small cell lung cancer: Visualization by positron
emission tomography in a patient harboring a mutation in the epidermal growth factor receptor. J. Thorac.
Oncol. 2011,6, 1287–1289. [CrossRef] [PubMed]
73. Traxl, A.; Wanek, T.; Mairinger, S.; Stanek, J.; Filip, T.; Sauberer, M.; Muller, M.; Kuntner, C.; Langer, O.
Breast cancer resistance protein and p-glycoprotein influence in vivo disposition of 11C-erlotinib. J. Nucl.
Med. 2015. [CrossRef] [PubMed]
22021
Molecules 2015,20, 22000–22027
74. Slobbe, P.; Windhorst, A.D.; Stigter-van Walsum, M.; Smit, E.F.; Niessen, H.G.; Solca, F.; Stehle, G.;
van Dongen, G.A.; Poot, A.J. A comparative PET imaging study with the reversible and irreversible EGFR
tyrosine kinase inhibitors [11C]Erlotinib and [18F]Afatinib in lung cancer-bearing mice. EJNMMI Res. 2015,
5, 14. [CrossRef] [PubMed]
75. Li, D.; Ambrogio, L.; Shimamura, T.; Kubo, S.; Takahashi, M.; Chirieac, L.R.; Padera, R.F.; Shapiro, G.I.;
Baum, A.; Himmelsbach, F.; et al. BIBW2992, an irreversible EGFR/HER2 inhibitor highly effective in
preclinical lung cancer models. Oncogene 2008,27, 4702–4711. [CrossRef] [PubMed]
76. Slobbe, P.; Windhorst, A.D.; Stigter-van Walsum, M.; Schuit, R.C.; Smit, E.F.; Niessen, H.G.; Solca, F.;
Stehle, G.; van Dongen, G.A.; Poot, A.J. Development of [18F]afatinib as new TKI-PET tracer for EGFR
positive tumors. Nucl. Med. Biol. 2014,41, 749–757. [CrossRef] [PubMed]
77. European Medicines Agency. Committee for Medicinal Products for Human Use (CHMP) Assessment
Report for Giotrif (afatinib). 16 October 2013. Available online: http://www.ema.europa.eu/docs/en_GB/
document_library/EPAR_-_Product_Information/human/002280/WC500152392.pdf (accessed 20 November 2013)
.
78. Copeland, R.A. The dynamics of drug-target interactions: Drug-target residence time and its impact on
efficacy and safety. Expert Opin. Drug Discov. 2010,5, 305–310. [CrossRef] [PubMed]
79. Johnson, D.S.; Weerapana, E.; Cravatt, B.F. Strategies for discovering and derisking covalent, irreversible
enzyme inhibitors. Future Med. Chem. 2010,2, 949–964. [CrossRef] [PubMed]
80. Bos, M.; Mendelsohn, J.; Kim, Y.M.; Albanell, J.; Fry, D.W.; Baselga, J. PD153035, a tyrosine kinase inhibitor,
prevents epidermal growth factor receptor activation and inhibits growth of cancer cells in a receptor
number-dependent manner. Clin. Cancer Res. 1997,3, 2099–2106. [PubMed]
81. Sasaki, T.; Ishii, S.I.; Senda, M.; Akinaga, S.; Murakata, C. Synthesis of [7β´methoxy 11C]methoxy
staurosporine for imaging protein kinase C localization in the brain. Appl. Radiat. Isot. 1996,47, 67–69. [CrossRef]
82. Wang, H.; Yu, J.M.; Yang, G.; Song, X.; Sun, X.; Zhao, S.; Mu, D. Assessment of
11C-labeled-4-N-(3-bromoanilino)-6,7-dimethoxyquinazoline as a positron emission tomography agent to
monitor epidermal growth factor receptor expression. Cancer Sci. 2007,9, 1413–1416. [CrossRef] [PubMed]
83. Wang, H.; Yu, J.M.; Yang, G.R.; Song, X.R.; Sun, X.R.; Zhao, S.Q.; Wang, X.W.; Zhao, W. Further
characterization of the epidermal growth factor receptor ligand 11C-PD153035. Chin. Med. J. 2007,120,
960–964. [PubMed]
84. Meng, X.; Yu, J.M.; Yang, G.R.; Zhao, S.Q.; Sun, X.D.l. 11C-PD153035 PET/CT for molecular imaging of
EGFR in patients with non-small cell lung cancer (NSCLC) [abstract]. J. Clin. Oncol. 2008, 26.
85. Liu, N.; Li, M.; Li, X.; Meng, X.; Yang, G.; Zhao, S.; Yang, Y.; Ma, L.; Fu, Z.; Yu, J. PET-based biodistribution
and radiation dosimetry of epidermal growth factor receptor-selective tracer 11C-PD153035 in humans.
J. Nucl. Med. 2009,50, 303–308. [CrossRef] [PubMed]
86. Meng, X.; Loo, B.W., Jr.; Ma, L.; Murphy, J.D.; Sun, X.; Yu, J. Molecular imaging with 11 C-PD153035 PET/CT
predicts survival in non-small cell lung cancer treated with EGFR-TKI: A pilot study. J. Nucl. Med. 2011,52,
1573–1579. [CrossRef] [PubMed]
87. Samen, E.; Arnberg, F.; Lu, L.; Olofsson, M.H.; Tegnebratt, T.; Thorell, J.O.; Holmin, S.; Stone-Elander, S.
Metabolism of epidermal growth factor receptor targeting probe [11C]PD153035: Impact on biodistribution
and tumor uptake in rats. J. Nucl. Med. 2013,54, 1804–1811. [CrossRef] [PubMed]
88. Sun, J.; Cai, L.; Zhang, K.; Zhang, A.; Pu, P.; Yang, W.; Gao, S. A pilot study on EGFR-targeted molecular
imaging of PET/CT with tracer 11C-PD153035 in human gliomas. Clin. Nucl. Med. 2014,39, e20–e26.
[CrossRef] [PubMed]
89. Basuli, F.; Wu, H.; Li, C.; Shi, Z.-D.; Sulima, A.; Griffiths, G.L. A first synthesis of 18F-radiolabeled Lapatinib:
A potential tracer for positron emission tomographic imaging of erbb1/erbb2 tyrosine kinase activity.
J. Label. Compd. Radiopharm. 2011,54, 633–636. [CrossRef]
90. Saleem, A.; Searle, G.E.; Kenny, L.M.; Huiban, M.; Kozlowski, K.; Waldman, A.D.; Woodley,L.; Palmieri, C.;
Lowdell, C.; Kaneko, T.; et al. Lapatinib access into normal brain and brain metastases in patients with
HER-2 overexpressing breast cancer. EJNMMI Res. 2015,5, 30. [CrossRef] [PubMed]
91. Polli, J.W.; Humphreys, J.E.; Harmon, K.A.; Castellino, S.; O’Mara, M.J.; Olson, K.L.;
John-Williams, L.S.; Koch, K.M.; Serabjit-Singh, C.J. The role of efflux and uptake transporters in
[N-{3-chloro-4-[(3-fluorobenzyl)oxy]phenyl}-6-[5-({[2-(methylsulfonyl)ethyl]amino}methyl)-2-furyl]-4-
quinazolinamine (GW572016, lapatinib) disposition and drug interactions. Drug Metab. Dispos. 2008,36,
695–701. [CrossRef] [PubMed]
22022
Molecules 2015,20, 22000–22027
92. Pivot, X.; Manikhas, A.; Z’urawski, B.; Chmielowska, E.; Karaszewska, B.; Allerton, R.; Chan, S.; Fabi, A.;
Bidoli, P.; Gori, S.; et al. CEREBEL (EGF111438): A Phase III, Randomized, Open-Label Study of Lapatinib
Plus Capecitabine Versus Trastuzumab Plus Capecitabine in Patients With Human Epidermal Growth
Factor Receptor 2–Positive Metastatic Breast Cancer. J. Clin. Oncol. 2015,33, 1–9. [CrossRef] [PubMed]
93. Pisaneschi, F.; Nguyen, Q.-D.; Shamsaei, E.; Glaser, M.; Robins, E.; Kaliszczak, M.; Smith, G.; Spivey, A.C.;
Aboagye, E.O. Development of a new epidermal growth factor receptor positron emission tomography
imaging agent based on the 3-cyanoquinoline core: Synthesis and biological evaluation. Bioorg. Med. Chem.
2010,18, 6634–6645. [CrossRef] [PubMed]
94. Pisaneschi, F.; Slade, R.L.; Iddon, L.; George, G.P.; Nguyen, Q.D.; Spivey, A.C.; Aboagye, E.O. Synthesis of a
new fluorine-18 glycosylated “click” cyanoquinoline for the imaging of epidermal growth factor receptor.
J. Label. Comp. Radiopharm. 2014,57, 92–96. [CrossRef] [PubMed]
95. Pal, A.; Balatoni, J.; Mukhopadhyay, U.; Ogawa, K.; Gonzalez-Lepera, C.; Shavrin, A.; Volgin, A.; Tong, W.;
Alauddin, M.; Gelovani, J. Radiosynthesis and initial in vitro evaluation of [18F]FPEG6-IPQA—A novel PET
radiotracer for imaging EGFR expression-activity in lung carcinomas. Mol. Imaging Biol. 2011,13, 853–861.
[CrossRef] [PubMed]
96. Yeh, H.H.; Ogawa, K.; Balatoni, J.; Mukhapadhyay, U.; Pal, A.; Gonzalez-Lepera, C.; Shavrin, A.;
Soghomonyan, S.; Flores, L., 2nd; Young, D.; et al. Molecular imaging of active mutant l858R EGF receptor
(EGFR) kinase-expressing nonsmall cell lung carcinomas using PET/ct. Proc. Natl. Acad. Sci. USA 2011,
108, 1603–1608. [CrossRef] [PubMed]
97. Yeh, S.H.; Lin, C.F.; Kong, F.L.; Wang, H.E.; Hsieh, Y.J.; Gelovani, J.G.; Liu, R.S. Molecular
imaging of nonsmall cell lung carcinomas expressing active mutant EGFR kinase using PET with
[124I]-morpholino-ipqa. Biomed. Res. Int. 2013,2013. [CrossRef] [PubMed]
98. Wang, M.; Gao, M.; Zheng, Q.H. The first radiosynthesis of [11C]AZD8931 as a new potential PET agent
for imaging of EGFR, HER2 and HER3 signaling. Bioorg. Med. Chem. Lett. 2014,24, 4455–4459. [CrossRef]
[PubMed]
99. Vasdev, N.; Dorff, P.N.; O’Neil, J.P.; Chin, F.T.; Hanrahan, S.; VanBrocklin, H.F. Metabolic stability of
6,7-dialkoxy-4-(2-, 3- and 4-[18F]fluoroanilino)quinazolines, potential EGFR imaging probes. Bioorg. Med.
Chem. 2011,19, 2959–2965. [CrossRef] [PubMed]
100. Medina, O.P.; Pillarsetty, N.; Glekas, A.; Punzalan, B.; Longo, V.; Gonen, M.; Zanzonico, P.; Smith-Jones, P.;
Larson, S.M. Optimizing tumor targeting of the lipophilic EGFR-binding radiotracer SKI 243 using a
liposomal nanoparticle delivery system. J. Control. Release 2011,149, 292–298. [CrossRef] [PubMed]
101. Neto, C.; Fernandes, C.; Oliveira, M.C.; Gano, L.; Mendes, F.; Kniess, T.; Santos, I. Radiohalogenated
4-anilinoquinazoline-based EGFR-TK inhibitors as potential cancer imaging agents. Nucl. Med. Biol. 2012,
39, 247–260. [CrossRef] [PubMed]
102. Chang, S.; Zhang, L.; Xu, S.; Luo, J.; Lu, X.; Zhang, Z.; Xu, T.; Liu, Y.; Tu, Z.; Xu, Y.; et al. Design,
synthesis, and biological evaluation of novel conformationally constrained inhibitors targeting epidermal
growth factor receptor threonine790Ñmethionine790mutant. J. Med. Chem. 2012,55, 2711–2723. [CrossRef]
[PubMed]
103. Han, C.; Huang, Z.; Zheng, C.; Wan, L.; Zhang, L.; Peng, S.; Ding, K.; Ji, H.; Tian, J.; Zhang, Y. Novel hybrids
of (phenylsulfonyl)furoxan and anilinopyrimidine as potent and selective epidermal growth factor receptor
inhibitors for intervention of non-small-cell lung cancer. J. Med. Chem. 2013,56, 4738–4748. [CrossRef]
[PubMed]
104. Heald, R.A.; Chan, B.K.; Bryan, C.; Eigenbrot, C.; Yu, C.; Burdick, D.; Hanan, E.J.; Chan, E.; Schaefer, G.;
La, H.; et al. Noncovalent Mutant Selective Epidermal Growth Factor Receptor Inhibitors: A Lead
Optimization Case History. J. Med. Chem. 2015. in press. [CrossRef] [PubMed]
105. Zhou, W.; Ercan, D.; Chen, L.; Yun, C.-H.; Li, D.; Capelletti, M.; Cortot, A.B.; Chirieac, L.; Iacob, R.E.;
Padera, R.; et al. Novel mutant-selective EGFR kinase inhibitors against EGFR T790M. Nature 2009,462,
1070–1074. [CrossRef] [PubMed]
106. Kumar, R.; Crouthamel, M.-C.; Rominger, D.H.; Gontarek, R.R.; Tummino1, P.J.; Levin, R.A.; King, A.G.
Myelosuppression and kinase selectivity of multikinase angiogenesis inhibitors. Br. J. Cancer 2009,101,
1717–1723. [CrossRef] [PubMed]
22023
Molecules 2015,20, 22000–22027
107. Hu-Lowe, D.D.; Zou, H.Y.; Grazzini, M.L.; Hallin, M.E.; Wickman, G.R.; Amundson, K.; Chen, J.H.;
Rewolinski, D.A.; Yamazaki, S.; Wu, E.Y.; et al. Nonclinical Antiangiogenesis and Antitumor Activities
of Axitinib (AG-013736), an Oral, Potent, and Selective Inhibitor of Vascular Endothelial Growth Factor
Receptor Tyrosine Kinases 1, 2, 3. Clin. Cancer Res. 2008,14, 7272–7283. [CrossRef] [PubMed]
108. Boss, D.S.; Glen, H.; Beijnen, J.H.; Keesen, M.; Morrison, R.; Tait, B.; Copalu, W.; Mazur, A.; Wanders, J.;
O’Brien, J.P.; Schellens, J.H.M.; Evans, T.R.J. A phase I study of E7080, a multitargeted tyrosine kinase
inhibitor, in patients with advanced solid tumours. Br. J. Cancer 2012,106, 1598–1604. [CrossRef] [PubMed]
109. Herbst, R.S.; Heymach, J.V.; O’Reilly, M.S.; Onn, A.; Ryan, A.J. Vandetanib (ZD6474): An orally
available receptor tyrosine kinase inhibitor that selectively targets pathways critical for tumor growth and
angiogenesis. Expert Opin. Investig. Drugs 2007,16, 239–249. [CrossRef] [PubMed]
110. Peters, J.-U. Polypharmacology—Foe or friend. J. Med. Chem. 2013,56, 8955–8971. [CrossRef] [PubMed]
111. Li, F.; Jiang, S.; Zu, Y.; Lee, D.Y.; Li, Z. A tyrosine kinase inhibitor-based high-affinity PET
radiopharmaceutical targets vascular endothelial growth factor receptor. J. Nucl. Med. 2014,55, 1525–1531.
[CrossRef] [PubMed]
112. Samen, E.; Lu, L.; Mulder, J.; Thorell, J.O.; Damberg, P.; Tegnebratt, T.; Holmgren, L.; Rundqvist, H.;
Stone-Elander, S. Visualization of angiogenesis during cancer development in the polyoma middle t breast
cancer model: Molecular imaging with (R)-[11C]PAQ. EJNMMI Res. 2014,4, 17. [CrossRef] [PubMed]
113. Gao, M.Z.; Lola, C.M.; Wang, M.; Miller, K.D.; Sledge, G.W.; Zheng, Q.H. Radiosynthesis of C-11 Vandetanib
and C-11 chloro-Vandetanib as new potential PET agents for imaging of VEGFR in cancer. Bioorg. Med.
Chem. Lett. 2011,21, 3222–3226. [CrossRef] [PubMed]
114. Asakawa, C.; Ogawa, M.; Kumata, K.; Fujinaga, M.; Kato, K.; Yamasaki, T.; Yui, J.; Kawamura, K.;
Hatori, A.; Fukumura, T.; et al. [11C]Sorafenib: Radiosynthesis and preliminary PET study of brain uptake
in p-gp/bcrp knockout mice. Bioorg. Med. Chem. Lett. 2011,21, 2220–2223. [CrossRef] [PubMed]
115. Poot, A.J.; van der Wildt, B.; Stigter-van Walsum, M.; Rongen, M.; Schuit, R.C.; Hendrikse, N.H.;
Eriksson, J.; van Dongen, G.A.; Windhorst, A.D. [11C]Sorafenib: Radiosynthesis and preclinical evaluation
in tumor-bearing mice of a new tki-PET tracer. Nucl. Med. Biol. 2013,40, 488–497. [CrossRef] [PubMed]
116. Caballero, J.; Muñoz, C.; Alzate-Morales, J.H.; Cunha, S.; Gano, L.; Bergmann, R.; Steinbach, J.; Kniess, T.
Synthesis, in silico, in vitro, and in vivo investigation of 5-[¹¹C]methoxy-substituted sunitinib, a tyrosine
kinase inhibitor of VEGFR-2. Eur. J. Med. Chem. 2012,58, 272–280. [CrossRef] [PubMed]
117. Slobbe, P.; Windhorst, A.D.; Haumann, R.; Schuit, R.; van Dongen, G.A.; Poot, A.J. Development of an
LC-MS method to analyze the primary metabolite of [11C]nintedanib in vivo.J. Label. Compd. Radiopharm.
2015, S30. [CrossRef]
118. Ilovich, O.; Billauer, H.; Dotan, S.; Mishani, E. Labeled 3-aryl-4-indolylmaleimide derivatives and their
potential as angiogenic PET biomarkers. Bioorg. Med. Chem. 2010,18, 612–620. [CrossRef] [PubMed]
119. Ilovich, O.; Åberg, O.; Långström, B.; Mishani, E. Rhodium-mediated [11C]carbonylation: A library of
n-phenyl-n1-{4-(4-quinolyloxy)-phenyl}-[11c]-urea derivatives as potential pet angiogenic probes. J. Label.
Compd. Radiopharm. 2009,52, 151–157. [CrossRef]
120. Dissoki, S.; Abourbeh, G.; Salnikov, O.; Mishani, E.; Jacobson, O. PET molecular imaging of angiogenesis
with a multiple tyrosine kinase receptor-targeted agent in a rat model of myocardial infarction. Mol.
Imaging Biol. 2015,17, 222–230. [CrossRef] [PubMed]
121. Fernandes, C.; Santos, I.C.; Santos, I.; Pietzsch, H.-J.; Kunstler, J.-U.; Kraus, W.; Rey, A.; Margaritis, N.;
Bourkoula, A.; Chiotellis, A.; et al. Rhenium and technetium complexes bearing quinazoline derivatives:
Progress towards a 99mTc biomarker for EGFR-TK imaging. Dalton Trans. 2008, 3215–3225. [CrossRef]
[PubMed]
122. Jørgensen, J.T.; Persson, M.; Madsen, J.; Kjær, A. High tumor uptake of 64Cu: implications for molecular
imaging of tumor characteristics with copper-based PET tracers. Nucl. Med. Biol. 2013,40, 345–350.
[CrossRef] [PubMed]
123. Lagas, J.S.; van Waterschoot, R.A.; Sparidans, R.W.; Wagenaar, E.; Beijnen, J.H.; Schinkel, A.H. Breast cancer
resistance protein and P-glycoprotein limit sorafenib brain accumulation. Mol. Cancer Ther. 2010,9, 319–326.
[CrossRef] [PubMed]
124. Kil, K.-E.; Ding, Y.-S.; Lin, K.-S.; Alexoff, D.; Kim, S.W.; Shea, C.; Xu, Y.; Muench, L.; Fowler, J.S. Synthesis
and positron emission tomography studies of carbon-11-labeled Imatinib (Gleevec). Nucl. Med. Biol. 2007,
34, 153–163. [CrossRef] [PubMed]
22024
Molecules 2015,20, 22000–22027
125. Peng, Z.; Bornman, W.; Pal, A.; Ghosh, P.; Lim, S.T.; Gelovani, J.; Maxwell, D.; Alauddin, M.M. STI571
analogs: 18F-STI571as potential agents for PET imaging of c-kit expression at a kinase level. In Proceedings
of the 233rd ACS National Meeting, Chicago, IL, USA, 25 March 2007.
126. Doubrovin, M.; Kochetkova, T.; Santos, E.; Veach, D.R.; Smith-Jones, P.; Pillarsetty, N.; Balatoni, J.;
Bornmann, W.; Gelovani, J.; Larson, S.M. (124)i-iodopyridopyrimidinone for PET of abl kinase-expressing
tumors in vivo.J. Nucl. Med. 2010,51, 121–129. [CrossRef] [PubMed]
127. Glekas, A.P.; Pillarsetty, N.K.; Punzalan, B.; Khan,N.; Smith-Jones, P.; Larson, S.M. In vivo imaging of bcr-abl
overexpressing tumors with a radiolabeled Imatinib analog as an imaging surrogate for Imatinib. J. Nucl.
Med. 2011,52, 1301–1307. [CrossRef] [PubMed]
128. Peng, Z.; Maxwell, D.S.; Sun, D.; Bhanu Prasad, B.A.; Pal, A.; Wang, S.; Balatoni, J.; Ghosh, P.; Lim, S.T.;
Volgin, A.; et al. Imatinib analogs as potential agents for PET imaging of bcr-abl and c-kit expression at a
kinase level. Bioorg. Med. Chem. 2014,22, 623–632. [CrossRef] [PubMed]
129. Veach, D.R.; Namavari, M.; Pillarsetty, N.; Santos, E.B.; Beresten-Kochetkov, T.; Lambek, C.; Punzalan, B.J.;
Antczak, C.; Smith-Jones, P.M.; Djaballah, H.; et al. Synthesis and biological evaluation of a fluorine-18
derivative of Dasatinib. J. Med. Chem. 2007,50, 5853–5857. [CrossRef] [PubMed]
130. Dunphy, M.P.; Zanzonico, P.; Veach, D.; Somwar, R.; Pillarsetty, N.; Lewis, J.; Larson, S. Dosimetry of
18F-labeled tyrosine kinase inhibitor ski-249380, a Dasatinib-tracer for PET imaging. Mol. Imaging Biol.
2012,14, 25–31. [CrossRef] [PubMed]
131. ClinicalTrials.gov. Available online: https://clinicaltrials.gov/ct2/show/NCT01916135 (accessed on 17
October 15).
132. Benezra, M.; Hambardzumyan, D.; Penate-Medina, O.; Veach, D.R.; Pillarsetty, N.; Smith-Jones, P.;
Phillips, E.; Ozawa, T.; Zanzonico, P.B.; Longo, V.; et al. Fluorine-labeled Dasatinib nanoformulations
as targeted molecular imaging probes in a pdgfb-driven murine glioblastoma model. Neoplasia 2012,14,
1132–1143. [CrossRef] [PubMed]
133. Chakravarty, R.; Hong, H.; Cai, W. Positron emission tomography image-guided drug delivery: Current
status and future perspectives. Mol. Pharm. 2014,11, 3777–3797. [CrossRef] [PubMed]
134. Bernard-Gauthier, V.; Aliaga, A.; Aliaga, A.; Boudjemeline, M.; Hopewell, R.; Kostikov, A.; Rosa-Neto, P.;
Thiel, A.; Schirrmacher, R. Syntheses and evaluation of carbon-11 and fluorine-18-radiolabeled
pan-tropomyosin receptor kinase (Trk) inhibitors : Exploration of the 4-aza-2-oxindole scaffold as Trk PET
imaging agents. ACS Chem. Neurosci. 2015,6, 260–276. [CrossRef] [PubMed]
135. Bernard-Gauthier, V.; Schirrmacher, R. 5-(4-((4-[18F]fluorobenzyl)oxy)-3-methoxybenzyl)pyrimidine-2,
4-diamine: A selective dual inhibitor for potential PET imaging of Trk/CSF-1R. Bioorg. Med. Chem. Lett.
2014,24, 4784–4790. [CrossRef] [PubMed]
136. Wu, C.; Tang, Z.; Fan, W.; Zhu, W.; Wang, C.; Somoza, E.; Owino, N.; Li, R.; Ma, P.C.; Wang, Y. In vivo
positron emission tomography (PET) imaging of mesenchymal-epithelial transition (MET) receptor. J. Med.
Chem. 2010,53, 139–146. [CrossRef] [PubMed]
137. Mamat, C.; Mosch, B.; Neuber, C.; Köckerling, M.; Bergmann, R.; Pietzsch, J. Fluorine-18 radiolabeling
and radiopharmacological characterization of a benzodioxolylpyrimidine-based radiotracer targeting the
receptor tyrosine kinase EphB4. ChemMedChem 2012,7, 1991–2003. [CrossRef] [PubMed]
138. Huang, E.J.; Reichardt, L.F. Trk receptors: Roles in neuronal signal transduction. Annu. Rev. Neurosci. 2003,
72, 609–642.
139. Vaishnavi, A.; Le, A.T.; Doebele, R.C. TRKing down an old oncogene in a new era of targeted therapy.
Cancer Discov. 2015,5, 25–34. [CrossRef] [PubMed]
140. Dienstmann, R.; Tabernero, J. BRAF as a target for cancer therapy. Anticancer Agents Med. Chem. 2011,11,
285–295. [CrossRef] [PubMed]
141. Blanco-Aparicio, C.; Carnero, A. Pim kinases in cancer: Diagnostic, prognostic and treatment opportunities.
Biochem. Pharmacol. 2013,85, 629–643. [CrossRef] [PubMed]
142. Gavriilidis, P.; Giakoustidis, A.; Giakoustidis, D. Aurora Kinases and Potential Medical Applications of
Aurora Kinase Inhibitors: A Review. J. Clin. Med. Res. 2015,7, 742–751. [CrossRef] [PubMed]
143. Ciuffreda, L.; di Sanza, C.; Incani, U.C.; Milella, M. The mTOR pathway: A new target in cancer therapy.
Curr. Cancer Drug Targets 2010,10, 484–495. [CrossRef] [PubMed]
144. Arfeen, M.; Bharatam, P.V. Design of glycogen synthase kinase-3 inhibitors: An overview on recent
advancements. Curr. Pharm. Des. 2013,19, 4755–4775. [CrossRef] [PubMed]
22025
Molecules 2015,20, 22000–22027
145. Vasdev, N.; LaRonde, F.J.; Woodgett, J.R.; Garcia, A.; Rubie, E.A.; Meyer, J.H.; Houle, S.; Wilson, A.A.
Rationally designed PKA inhibitors for positron emission tomography: Synthesis and cerebral
biodistribution of N-(2-(4-bromocinnamylamino)ethyl)-N-[11C]methyl-isoquinoline-5-sulfonamide. Bioorg.
Med. Chem. 2008,16, 5277–5284. [CrossRef] [PubMed]
146. Wang, M.; Gao, M.; Miller, K.D.; Sledge, G.W.; Hutchins, G.D.; Zheng, Q.-H. The first design and synthesis
of [11C]MKC-1 ([11C]Ro 31–7453), a new potential PET cancer imaging agent. Nucl. Med. Biol. 2010,37,
763–775. [CrossRef] [PubMed]
147. Takahashi, K.; Kudo, K.; Okada, M.; Yanamoto, K.; Hatori, A.; Irie, T.; Suzuki, K.; Miura, S. Synthesis and
biodistribution of [11C]methyl-bisindolylmareimide III, an inhibitor of protein kinase C. J. Label. Compd.
Radiopharm. 2007,50, S398.
148. Cai, L.; Ozaki, H.; Fujita, M.; Hong, J.S.; Bukhari, M.; Innis, R.B.; Pike, V.W. [11C]GO6976 as a potential
radioligand for imaging protein kinase C with PET. J. Label. Compd. Radiopharm. 2009,52, S337.
149. Vasdev, N.; Garcia, A.; Stableford, W.T.; Young, A.B.; Meyer, J.H.; Houle, S.; Wilson, A.A. Synthesis
and ex vivo evaluation of carbon-11 labelled N-(4-methoxybenzyl)-]N1-(5-nitro-1,3-thiazol-2-yl)urea
([C-11]AR-A014418): A radiolabelled glycogen synthase kinase-3 beta specific inhibitor for PET studies.
Bioorg. Med. Chem. 2005,15, 5270–5273. [CrossRef] [PubMed]
150. Hicks, J.W.; Wilson, A.A.; Rubie, E.A.; Woodgett, J.R.; Houle, S.; Vasdev, N. Towards the preparation of
radiolabeled 1-aryl-3-benzyl ureas: Radiosynthesis of [11C-carbonyl] AR-A014418 by [11 C]CO2fixation.
Bioorg. Med. Chem. Lett. 2012,22, 2099–2101. [CrossRef] [PubMed]
151. Rotstein, B.H.; Liang, S.H.; Holland, J.P.; Collier, T.L.; Hooker, J.M.; Wilson, A.A.; Vasdev, N. 11CO2 fixation:
A renaissance in PET radiochemistry. Chem. Commun. 2013,49, 5621–5629. [CrossRef] [PubMed]
152. Cole, E.L.; Shao, X.; Sherman, P.; Quesada, C.; Fawaz, M.V.; Desmond, T.J.; Scott, P.J. Synthesis and
evaluation of [11C]pyrATP-1, a novel radiotracer for PET imaging of glycogen synthase kinase-3beta
(GSK-3β). Nucl. Med. Biol. 2014,41, 507–512. [CrossRef] [PubMed]
153. Kumata, K.; Yui, J.; Xie, L.; Zhang, Y.; Nengaki, N.; Fujinaga, M.; Yamasaki, T.; Shimoda, Y.; Zhang, M.R.
Radiosynthesis and preliminary PET evaluation of glycogen synthase kinase 3beta (GSK-3β) inhibitors
containing [11C]methylsulfanyl, [11C]methylsulfinyl or [11C]methylsulfonyl groups. Bioorg. Med. Chem.
Lett. 2015,25, 3230–3233. [CrossRef] [PubMed]
154. Wang, M.; Gao, M.; Miller, K.D.; Sledge, G.W.; Hutchins, G.D.; Zheng, Q.H. The first synthesis of
[11C]SB-216763, a new potential PET agent for imaging of glycogen synthase kinase-3 (GSK-3). Bioorg.
Med. Chem. Lett. 2011,21, 245–249. [CrossRef] [PubMed]
155. Li, L.; Shao, X.; Cole, E.L.; Ohnmacht, S.A.; Ferrari, V.; Hong, Y.T.; Williamson, D.J.; Fryer, T.D.;
Quesada, C.A.; Sherman, P.; Riss, P.J.; et al. Synthesis and initial in vivo studies with [11C]SB-216763: The
first radiolabeled brain penetrative inhibitor of GSK-3. ACS Med. Chem. Lett. 2015,6, 548–552. [CrossRef]
[PubMed]
156. Valdivia,A.C.; Mason, S.; Collins, J.; Buckley, K.R.; Coletta, P.; Beanlands, R.S.; Dasilva, J.N. Radiosynthesis
of N-[11C]-methyl-hydroxyfasudil as a new potential PET radiotracer for rho-kinases (ROCKs). Appl.
Radiat. Isot. 2010,68, 325–328. [CrossRef] [PubMed]
157. Taniguchi, J.; Seki, C.; Takuwa, H.; Kawaguchi, H.; Ikoma, Y.; Fujinaga, M.; Kanno, I.; Zhang, M.R.;
Kuwabara, S.; Ito, H. Evaluation of rho-kinase activity in mice brain using N-[11C]methyl-hydroxyfasudil
with positron emission tomography. Mol. Imaging Biol. 2014,16, 395–402. [CrossRef] [PubMed]
158. Moreau, S.; DaSilva, J.N.; Valdivia, A.; Fernando, P. N-[11C]-Methyl-hydroxyfasudil is a potential biomarker
of cardiac hypertrophy. Nucl. Med. Biol. 2015,42, 192–197. [CrossRef] [PubMed]
159. Koehler, L.; Graf, F.; Bergmann, R.; Steinbach, J.; Pietzsch, J.; Wuest, F. Radiosynthesis and
radiopharmacological evaluation of cyclin-dependent kinase 4 (CDK4) inhibitors. Eur. J. Med. Chem. 2010,
45, 727–737. [CrossRef] [PubMed]
160. Shimoda, Y.; Yui, J.; Fujinaga, M.; Xie, L.; Kumata, K.; Ogawa, M.; Yamasaki, T.; Hatori, A.; Kawamura, K.;
Zhang, M.R. [11C-carbonyl]CEP-32496: Radiosynthesis, biodistribution and PET study of brain uptake in
p-gp/bcrp knockout mice. Bioorg. Med. Chem. Lett. 2014,24, 3574–3577. [CrossRef] [PubMed]
161. Slobbe, P.; Windhorst, A.D.; van Dongen, G.A.; Poot, A.J. Synthesis of [11C]vemurafenib via a unique
[11C]CO carbonylative Stille coupling to image V600E mutated B-Raf in cancer. J. Label. Compd. Radiopharm.
2015,58, S281. [CrossRef]
22026
Molecules 2015,20, 22000–22027
162. Goos, J.A.; Verbeek, J.; Geldof, A.A.; Hiemstra, A.C.; van de Wiel, M.A.; Adamzek, K.A.; Delis-Van
Diemen, P.M.; Stroud, S.G.; Bradley, D.P.; Meijer, G.A.; et al. Molecular imaging of aurora kinase A
(AURKA) expression: Synthesis and preclinical evaluation of radiolabeled alisertib (MLN8237). Nucl. Med.
Biol. 2015. [CrossRef] [PubMed]
163. Suzuki, M.; Takashima-Hirano, M.; Koyama, H.; Yamaoka, T.; Sumi, K.; Nagata, H.; Hidaka, H.; Doi, H.
Efficient synthesis of [11C]H-1152, a PET probe specific for Rho-kinases, highly potential targets in
diagnostic medicine and drug development. Tetrahedron 2012,68, 2336–2341. [CrossRef]
164. Svensson, F.; Kniess, T.; Gergmann, R.; Pietzsch, J.; Wuest, F. Synthesis of an 18F-labeled cyclin-dependant
kinase-2 inhibitor. J. Label. Compd. Radiopharm. 2011,54, 769–774. [CrossRef]
165. Wang, M.; Gao, M.; Miller, K.D.; Zheng, Q.H. Synthesis of 2,6-difluoro-N-(3-[11C]methoxy-1H-
pyrazolo[3,4-b]
pyridine-5-yl)-3-(propylsulfonamidio)benzamide as a new potential PET agent for imaging of B-RafV6˝˝E
in cancers. Bioorg. Med. Chem. Lett. 2013,23, 1017–1021. [CrossRef] [PubMed]
166. Wang, M.; Gao, M.; Miller, K.D.; Sledge, G.W.; Zheng, Q.H. [11C]GSK2126458 and [18F]GSK2126458, the
first radiosynthesis of new potential PET agents for imaging of PI3K and mTOR in cancers. Bioorg. Med.
Chem. Lett. 2012,22, 1569–1574. [CrossRef] [PubMed]
167. Majo, V.J.; Simpson, N.R.; Prabhakaran, J.; Mann, J.J.; Kumar, J.S. Radiosynthesis of [18F]ATPFU: A potential
PET ligand for mTOR. J. Label. Comp. Radiopharm. 2014,57, 705–709. [CrossRef] [PubMed]
168. Wang, M.; Gao, M.; Zheng, Q.H. Synthesis of carbon-11-labeled 4-(phenylamino)-pyrrolo[2,1-f]triazine
derivatives as new potential PET tracers for imaging of p38αmitogen-activated protein kinase. Bioorg.
Med. Chem. Lett. 2014,24, 3700–3705. [CrossRef] [PubMed]
169. Wang, M.; Xu, L.; Gao, M.; Miller, K.D.; Sledge, G.W.; Zheng, Q.H. [11C]enzastaurin, the first design and
radiosynthesis of a new potential PET agent for imaging of protein kinase C. Bioorg. Med. Chem. Lett. 2011,
21, 1649–1653. [CrossRef] [PubMed]
170. Gao, M.; Wang, M.; Miller, K.D.; Zheng, Q.H. Synthesis of (Z)-2-((1h-indazol-3-yl)methylene)-
6-[11C]methoxy-
7-(piperazin-1-ylmethyl)benzofuran-3(2H)-one as a new potential PET probe for imaging of the enzyme
PIM1. Bioorg. Med. Chem. Lett. 2013,23, 4342–4346. [CrossRef] [PubMed]
171. Wang, M.; Tzintzun, R.; Gao, M.; Xu, Z.; Zheng, Q.H. Synthesis of [11 C]CX-6258 as a new PET tracer for
imaging of pim kinases in cancer. Bioorg. Med. Chem. Lett. 2015,25, 3831–3835. [CrossRef] [PubMed]
172. Morph, R. Selectively Nonselective Kinase Inhibition: Striking the Right Balance. J. Med. Chem. 2010,53,
1413–1437. [CrossRef] [PubMed]
173. Uitdehaag, J.C.; Verkaar, F.; Alwan, H.; de Man, J.; Buijsman, R.C.; Zaman, G.J. A guide to picking the most
selective kinase inhibitor tool compounds for pharmacological validation of drug targets. Br. J. Pharmacol.
2012,166, 858–876. [CrossRef] [PubMed]
174. Zhao, Z.; Wu, H.; Wang, L.; Liu, Y.; Knapp, S.; Liu, Q.; Gray, N.S. Exploration of type II binding mode:
A privileged approach for kinase inhibitor focused drug discovery? ACS Chem. Biol. 2014,9, 1230–1241.
[CrossRef] [PubMed]
175. Stachel, S.J.; Sanders, J.M.; Henze, D.A.; Rudd, M.T.; Su, H.P.; Li, Y.; Nanda, K.K.; Egbertson, M.S.;
Manley, P.J.; Jones, K.L.; et al. Maximizing diversity from a kinase screen: Identification of novel and
selective pan-Trk inhibitors for chronic pain. J. Med. Chem. 2014,57, 5800–5816. [CrossRef] [PubMed]
176. Regan, J.; Pargellis, C.A.; Cirillo, P.F.; Gilmore, T.; Hickey, E.R.; Peet, G.W.; Proto, A.; Swinamer, A.; Moss, N.
The Kinetics of Binding to p38 MAP Kinase by Analogues of BIRB 796. Bioorg. Med. Chem. Lett. 2003,13,
3101–3104. [CrossRef]
177. Pargellis, C.; Tong, L.; Churchill, L.; Cirillo, P.F.; Gilmore, T.; Graham, A.G.; Grob, P.M.; Hickey, E.R.;
Moss, N.; Pav, S.; et al. Inhibition of p38 MAP Kinase by Utilizing a Novel Allosteric Binding Site. Nat.
Struct. Biol. 2002,9, 268–272. [CrossRef] [PubMed]
178. Gavrin, L.K.; Saiah, E. Approaches to discover non-ATP site kinase inhibitors. Med. Chem. Commun. 2013,
4, 41–51. [CrossRef]
179. Breen, M.E.; Soellner, M.B. Small molecule substrate phosphorylation site inhibitors of protein kinases:
Approaches and challenges. ACS Chem. Biol. 2014,10, 175–189. [CrossRef] [PubMed]
© 2015 by the authors; licensee MDPI, Basel, Switzerland. This article is an open
access article distributed under the terms and conditions of the Creative Commons by
Attribution (CC-BY) license (http://creativecommons.org/licenses/by/4.0/).
22027
... Lapatinib has been approved by the FDA for patients with ERBB2-positive breast carcinoma (31,(34)(35)(36). It is a competitive inhibitor of EGFR and ERBB2, previously reported to reduce the proliferation, inhibit multiplication and increase the apoptosis of HNSCC cells and other tumor xenografts expressing EGFR and ERBB2 (13,37). ...
... By shutting out the introduction of the phosphate group to the receptor and the subsequent activation of these routes, apoptosis is enhanced, and the growth of the cancer cells is inhibited (13). Furthermore, it has been reported that this drug can inhibit the proliferation of various human cancer cell lines (34,35). Theoretically, lapatinib has the advantage of being a dual TKI (for EGFR/ERBB2) (8). ...
... Radiolabeling of poly ADP ribose polymerase (PARP) inhibitors is gaining interest with numerous preclinical studies and an ongoing clinical trial in GB patients using [36,[39][40][41][42][43]. The first clinical results of [ 18 F]-Fluciclovine ([ 18 F]F-ACBC) for GB imaging were promising and radiolabelling of receptor tyrosine kinase inhibitors and mammalian target of rapamycin (mTOR) pathway inhibitors has also shown potential [44][45][46][47][48][49][50][51]. It is noted that PET imaging using the deoxycytidine kinase substrate [ 18 F]F-clofarabin has been shown to be a good imaging tool to localise and quantify responses in GB patients undergoing immunotherapy [52]. ...
... For CED, a catheter system, stereotactically placed intratumourally or into the post-surgical cavity, employs a pump to provide continuous positive pressure for local drug delivery (ranging from 0.1 to 10 μl/min) (Figure 3 and Figure 4) instead of a bolus injection [128,129]. This was proved to be a safe and effective drug delivery [46,54,55,[59][60][61][62]76,77,84,86,93,96,105,106,114,126,149,150,[165][166][167]184,227,229,231,[233][234][235][236][237][238][239][240][241][242][243][244]280 method, reaching a higher concentration of the drug within the GB tumour, and lack of systemic toxicity. This is especially favourable for α-particle emitters with relatively short half-lifes, such as bismuth-213 (45 min) or astatine-211 (7.2 h), as most of the radioactive decay will occur within the relevant cavity before being distributed throughout the body via the systemic and lymphatic systems [130]. ...
Article
Full-text available
Despite numerous clinical trials and pre-clinical developments, the treatment of glioblastoma (GB) remains a challenge. The current survival rate of GB averages one year, even with an optimal standard of care. However, the future promises efficient patient-tailored treatments, including targeted radionuclide therapy (TRT). Advances in radiopharmaceutical development have unlocked the possibility to assess disease at the molecular level allowing individual diagnosis. This leads to the possibility of choosing a tailored, targeted approach for therapeutic modalities. Therapeutic modalities based on radiopharmaceuticals are an exciting development with great potential to promote a personalised approach to medicine. However, an effective targeted radionuclide therapy (TRT) for the treatment of GB entails caveats and requisites. This review provides an overview of existing nuclear imaging and TRT strategies for GB. A critical discussion of the optimal characteristics for new GB targeting therapeutic radiopharmaceuticals and clinical indications are provided. Considerations for target selection are discussed, i.e. specific presence of the target, expression level and pharmacological access to the target, with particular attention to blood-brain barrier crossing. An overview of the most promising radionuclides is given along with a validation of the relevant radiopharmaceuticals and theranostic agents (based on small molecules, peptides and monoclonal antibodies). Moreover, toxicity issues and safety pharmacology aspects will be presented, both in general and for the brain in particular.
... To date, PET imaging of T cells has been described in preclinical models targeting T cell surface markers CD3, CD4 and CD8, and uptake was correlated with response to immunotherapy [21][22][23]. Recent preclinical studies have demonstrated that the use of mAbs for Tcell imaging can impair Tcell function, despite being administered at low doses [24], [17]. These functional effects of mAbs are likely due to their bivalent nature and interaction with speci c Fc receptors. ...
Preprint
Full-text available
Background CD103 is an integrin specifically expressed on the surface of cancer-reactive T cells. CD103 has been linked with better disease-specific survival in patients with ovarian, cervical, and endometrial cancer. The number of CD103⁺ T cells significantly increases during successful immunotherapy across human malignancies and therefore might be an attractive biomarker for non-invasive immune PET imaging of T cell infiltration. Indeed, we previously demonstrated that zirconium-89 (⁸⁹Zr) radiolabeled anti-CD103 antibodies could be used for PET imaging of CD103⁺ T cells at relevant cell densities. However, the long half-life of antibodies precluded repeat imaging of CD103⁺ T cell dynamics early in therapy, and is associated with a significant radiation burden. Methods Two different anti-human CD103 Fab fragments radiolabeled with ⁸⁹Zr or ⁶⁸Ga were developed, namely ⁸⁹Zr- hCD103 Fab and ⁶⁸Ga-hCD103 Fab respectively. In vivo evaluation of these tracers was performed in nude mice (BALB/cOlaHsd-Foxn1nu) with established CD103-expressing CHO (CHO.CD103) or CHO wildtype (CHO.K1) xenografts, followed by serial PET imaging and ex vivo bio-distribution. Results In vivo, both ⁸⁹Zr- and ⁶⁸Ga- hCD103 Fab tracers showed high target-to-background ratios, high target site selectivity and high sensitivity in human CD103 positive xenografts. Conclusion We conclude that the two novel human CD103 immuno-PET tracers may be used for future non-invasive assessment of cancer reactive T cell infiltration. Consequently, both ⁸⁹Zr and ⁶⁸Ga- hCD103 Fab PET tracers should be explored in the clinical setting for stratification of patients who could benefit from immune checkpoint inhibition therapy.
... Positron emission tomography (PET) is a molecular imaging technique that allows repeated and non-invasive clinical assessment. [38][39][40][41][42][43][44][45][46] PET is characterized by a high spatial resolution, sensitivity, and possibility to quantify the imaging signal and thus ideally suited for determining whole-body T RM load using radiolabeled mAbs. The PET isotope zirconium-89 ( 89 Zr; t 1/2 =78.4 hours) is favorable for radiolabeling mAbs, as its physical half-life matches the time mAbs require for optimal target-to-background signals. ...
Article
Full-text available
Purpose CD103, an integrin specifically expressed on the surface of cancer-reactive T cells, is significantly increased during successful immunotherapy across human malignancies. In this study, we describe the generation and zirconium-89 ( ⁸⁹ Zr) radiolabeling of monoclonal antibody (mAb) clones that specifically recognize human CD103 for non-invasive immune positron-emission tomography (PET) imaging of T cell infiltration as potential biomarker for effective anticancer immune responses. Experimental design First, to determine the feasibility of anti-CD103 immuno-PET to visualize CD103-positive cells at physiologically and clinically relevant target densities, we developed an ⁸⁹ Zr-anti-murine CD103 PET tracer. Healthy, non-tumor bearing C57BL/6 mice underwent serial PET imaging after intravenous injection, followed by ex vivo biodistribution. Tracer specificity and macroscopic tissue distribution were studied using autoradiography combined with CD103 immunohistochemistry. Next, we generated and screened six unique mAbs that specifically target human CD103 positive cells. Optimal candidates were selected for ⁸⁹ Zr-anti-human CD103 PET development. Nude mice (BALB/cOlaHsd-Foxn1nu) with established CD103 expressing Chinese hamster ovary (CHO) or CHO wild-type xenografts were injected with ⁸⁹ Zr-anti-human CD103 mAbs and underwent serial PET imaging, followed by ex vivo biodistribution. Results ⁸⁹ Zr-anti-murine CD103 PET imaging identified CD103-positive tissues at clinically relevant target densities. For human anti-human CD103 PET development two clones were selected based on strong binding to the CD103 ⁺ CD8 ⁺ T cell subpopulation in ovarian cancer tumor digests, non-overlapping binding epitopes and differential CD103 blocking properties. In vivo, both ⁸⁹ Zr-anti-human CD103 tracers showed high target-to-background ratios, high target site selectivity and a high sensitivity in human CD103 positive xenografts. Conclusion CD103 immuno-PET tracers visualize CD103 T cells at relevant densities and are suitable for future non-invasive assessment of cancer reactive T cell infiltration.
Article
Overexpression of the epidermal growth factor receptor (EGFR, erbB1) has been observed in a wide range of solid tumors and has frequently been associated with poor prognosis. As a result, EGFR inhibition has become an attractive anticancer drug design strategy, and a large number of small molecular inhibitors have been developed. Despite the widespread clinical use of EGFR tyrosine kinase inhibitors (TKIs), their drug resistance, inadequate accumulation in tumors, and severe side effects have spurred the search for better antitumor drugs. Metal complexes have attracted much attention because of their different mechanisms compared with EGFR‐TKIs. Therefore, the combination of metals and inhibitors is a promising anticancer strategy. For example, Ru and Pt centers are introduced to design complexes with double or multiple targets, while Au complexes are combined with inhibitors to overcome drug resistance. Co complexes are designed as prodrugs with weak side effects and enhanced targeting by the hypoxia activation strategy, and other metals such as Rh and Fe enhance the anticancer effect of the complexes. In addition, the introduction of Ga center is beneficial to the development of nuclear imaging tracers. In this paper, metal EGFR‐TKI complexes in the last 15 years are reviewed, their mechanisms are briefly introduced, and their advantages are summarized.
Article
Full-text available
Significant evidence suggests that the failure of clinically tested epidermal growth factor receptor (EGFR) tyrosine kinase inhibitors (e.g. erlotinib, lapatinib, gefitinib) in glioblastoma (GBM) patients is primarily attributed to insufficient brain penetration, resulting in inadequate exposure to the targeted cells. Molecular imaging tools can facilitate GBM drug development by visualizing drug biodistribution and confirming target expression and localization. To assess brain exposure via PET molecular imaging, we synthesized fluorine-18 isotopologues of two brain-penetrant EGFR tyrosine kinase inhibitors developed specifically for GBM. Adapting our recently reported radiofluorination of N-arylsydnones, we constructed an ortho-disubstituted [¹⁸F]fluoroarene as the key intermediate. The radiotracers were produced on an automated synthesis module in 7–8% activity yield with high molar activity. In vivo PET imaging revealed rapid brain uptake in rodents and tumor accumulation in an EGFR-driven orthotopic GBM xenograft model.
Article
Bruton's tyrosine kinase (BTK) is a target for treating B-cell malignancies and autoimmune diseases. To aid in the discovery and development of BTK inhibitors and improve clinical diagnoses, we have developed a positron emission tomography (PET) radiotracer based on a selective BTK inhibitor, remibrutinib. [18F]PTBTK3 is an aromatic, 18F-labeled tracer that was synthesized in 3 steps with a 14.8 ± 2.4% decay-corrected radiochemical yield and ≥99% radiochemical purity. The cellular uptake of [18F]PTBTK3 was blocked up to 97% in JeKo-1 cells using remibrutinib or non-radioactive PTBTK3. [18F]PTBTK3 exhibited renal and hepatobiliary clearance in NOD SCID (non-obese diabetic/severe combined immunodeficiency) mice, and the tumor uptake of [18F]PTBTK3 in BTK-positive JeKo-1 xenografts (1.23 ± 0.30% ID/cc) was significantly greater at 60 min post injection compared to the tumor uptake in BTK-negative U87MG xenografts (0.41 ± 0.11% ID/cc). In the JeKo-1 xenografts, tumor uptake was blocked up to 62% by remibrutinib, indicating the BTK-dependent uptake of [18F]PTBTK3 in tumors.
Article
Full-text available
Poly (ADP-ribose) polymerases (PARPs) constitute of 17 members that are associated with divergent cellular processes and play a crucial role in DNA repair, chromatin organization, genome integrity, apoptosis, and inflammation. Multiple lines of evidence have shown that activated PARP1 is associated with intense DNA damage and irritating inflammatory responses, which are in turn related to etiologies of various neurological disorders. PARP1/2 as plausible therapeutic targets have attracted considerable interests, and multitudes of PARP1/2 inhibitors have emerged for treating cancer, metabolic, inflammatory, and neurological disorders. Furthermore, PARP1/2 as imaging targets have been shown to detect, delineate, and predict therapeutic responses in many diseases by locating and quantifying the expression levels of PARP1/2. PARP1/2-directed noninvasive positron emission tomography (PET) has potential in diagnosing and prognosing neurological diseases. However, quantitative PARP PET imaging in the central nervous system (CNS) has evaded us due to the challenges of developing blood-brain barrier (BBB) penetrable PARP radioligands. Here, we review PARP1/2's relevance in CNS diseases, summarize the recent progress on PARP PET and discuss the possibilities of developing novel PARP radiotracers for CNS diseases.
Chapter
The goal of nuclear molecular imaging is to non-invasively visualize and quantify a specific target that is involved in a certain molecular or cellular process. This is achieved with a ligand that binds to this target, which is labeled with a radionuclide that can be visualized with single photon emission computed tomography (SPECT) or positron emission tomography (PET) imaging. An important criterium for obtaining good images is that the target is specifically expressed or active in the tissue of interest as compared to the surrounding tissues, leading to high target-to-background ratios. Furthermore, the radioligand should bind specifically to the target of interest, without activating the target or altering target function. In this chapter, various classes of targets that have been employed for nuclear imaging are introduced. In the current overview the targets are divided into (1) receptors involved in signal transduction, (2) enzymes, and (3) components of the extracellular matrix. For each class, examples of targets and corresponding radioligands that are already implemented in (preclinical) imaging are described. Furthermore, general considerations, hurdles and pitfalls when designing radioligands for specific targets such as receptors and enzymes are discussed. At last, an outlook on future potentially interesting targets, not only for imaging but also for radionuclide therapy is given.
Chapter
The rise in the number of fluorescent tracers that have become available over recent years follows the clinical demand to provide more specific detail on disease-related features during surgery. However, fluorescent tracers standardized assessment criteria that determine tracer suitability for clinical translation do not exist. In this article, an overview of the different approaches in tracer development that are currently pursued for fluorescent tracers is provided, with a special focus on receptor-targeted tracers. Based on a selected number of clinically relevant targets (Folate, EGFR, PSMA, CXCR4 and VEGF/ανβ3) key features of in vivo fluorescence imaging are highlighted. We also discuss potential improvements in the preclinical assessment of fluorescent tracers (including the influence of the dye component on the receptor affinity, in vivo biodistribution and tumor visualization). In addition, basic photophysical principles of fluorescence imaging and up-and-coming applications such as tracer assessment in large animal models help underline the translational efforts and potential in this emerging field of molecular imaging research.
Article
Full-text available
Background: Tyrosine kinase inhibitors (TKIs) have experienced a tremendous boost in the last decade, where more than 15 small molecule TKIs have been approved by the FDA. Unfortunately, despite their promising clinical successes, a large portion of patients remain unresponsive to these targeted drugs. For non-small cell lung cancer (NSCLC), the effectiveness of TKIs is dependent on the mutational status of epidermal growth factor receptor (EGFR). The exon 19 deletion as well as the L858R point mutation lead to excellent sensitivity to TKIs such as erlotinib and gefitinib; however, despite initial good response, most patients invariably develop resistance against these first-generation reversible TKIs, e.g., via T790M point mutation. Second-generation TKIs that irreversibly bind to EGFR wild-type and mutant isoforms have therefore been developed and one of these candidates, afatinib, has now reached the market. Whether irreversible TKIs differ from reversible TKIs in their in vivo tumor-targeting properties is, however, not known and is the subject of the present study. Methods: Erlotinib was labeled with carbon-11 and afatinib with fluorine-18 without modifying the structure of these compounds. A preclinical positron emission tomography (PET) study was performed in mice bearing NSCLC xenografts with a representative panel of mutations: an EGFR-WT xenograft cell line (A549), an acquired treatment-resistant L858R/T790M mutant (H1975), and a treatment-sensitive exon 19 deleted mutant (HCC827). PET imaging was performed in these xenografts with both tracers. Additionally, the effect of drug efflux transporter permeability glycoprotein (P-gp) on the tumor uptake of tracers was explored by therapeutic blocking with tariquidar. Results: Both tracers only demonstrated selective tumor uptake in the HCC827 xenograft line (tumor-to-background ratio, [(11)C]erlotinib 1.9 ± 0.5 and [(18)F]afatinib 2.3 ± 0.4), thereby showing the ability to distinguish sensitizing mutations in vivo. No major differences were observed in the kinetics of the reversible and the irreversible tracers in each of the xenograft models. Under P-gp blocking conditions, no significant changes in tumor-to-background ratio were observed; however, [(18)F]afatinib demonstrated better tumor retention in all xenograft models. Conclusions: TKI-PET provides a method to image sensitizing mutations and can be a valuable tool to compare the distinguished targeting properties of TKIs in vivo.
Article
7576 Background: Our previous study suggests that ¹¹ C-PD153035, a specific and potent inhibitor of EGFR tyrosine kinase (EGFR-TKI), is a novel PET radiotracer and ¹¹ C-PD153035 PET/CT is a promising non-invasive method for in vitro and vivo imaging of EGFR in NSCLC. We report a pilot study evaluating the efficacy of ¹¹ C-PD153035 PET/CT imaging as a “molecular fingerprint” to the EGFR-TKI in advanced NSCLC (stage IIIB/IV). Methods: Patients with pathologically proved advanced NSCLC were enrolled. ¹¹ C-PD153035 PET/CT was performed before (PET/CT 1) and 4 weeks after (PET/CT 2) received EGFR-targeted therapy (erlotinib) with assessment of maximum standardized uptake values (SUVmax) for correlation with response. Treatment response was also expressed as CT results on the basis of traditional Response Evaluation Criteria in Solid Tumors (RECIST) criteria. Image fusion and calculation of tumor subvolumes was performed on Xeleris workstation. Radioactivity concentrations, derived from regions of interest, were analyzed to the SUVmax. Results: 12 patients (5 men, 7 women; age range, 60–79 years) have been enrolled in this study from August 2008, including 3 cases of squamous cell carcinoma, 1 case of large cell carcinoma, and the other of adenocarcinoma. Tumor uptake of ¹¹ C-PD153035 before received therapy was observed in 8 patients and SUVmax ranged from 3.1 to 5.9. There was significant difference in SUVmax (p < 0.01) between tumor and normal tissue. Tumor/lung ratio at 20 min was 4.14 ± 1.80. The SUVmax did not correlate with tumor marker before therapy in our limited data. After 4 weeks therapy, a total of 7 patients underwent the 11C-PD153035 PET /CT again and 5 patients only received CT scanning. Tumor responders had a higher baseline pre-therapy SUVmax than nonresponders (p < 0.01). However, there was no significant difference between pre- and post-therapy SUVmax among 7 patients (p > 0.05). Conclusions: Our data indicated that 11C-PD153035 PET/CT before therapy may play a vital role for identification of patients that could be suitable for EGFR targeted therapies. Thus, it offers opportunities to individualize and optimize target therapy for NSCLC patients. No significant financial relationships to disclose.
Article
Due to their increased activity against activating mutants, first-generation epidermal growth factor receptor (EGFR) kinase inhibitors have had remarkable success in treating nonsmall cell lung cancer (NSCLC) patients, but acquired resistance, through a secondary mutation of the gatekeeper residue, means that clinical responses are short-lived. Addressing this unmet medical need requires agents which can target both of the most common double mutants: T790M/L858R (TMLR) and T790M/del(746-750) (TMdel). Herein we describe how a noncovalent double-mutant selective lead compound was optimized using a strategy focused on the structure-guided increase in potency without added lipophilicity or reduction of 3-dimensional character. Following successive rounds of design and synthesis it was discovered that cis-fluoro substitution on 4-hydroxy and 4-methoxy piperidinyl groups provided synergistic, substantial, and specific potency gain either through direct interaction with the enzyme and/or effects on the proximal ligand oxygen atom. Further development of the fluoro-hydroxy piperidine series resulted in the identification of a pair of diastereomers which showed 50-fold enzyme and cell based selectivity for T790M mutants over wild-type EGFR (wtEGFR) in vitro and pathway knock-down in an in vivo xenograft model.
Article
Introduction: Survival of patients after resection of colorectal cancer liver metastasis (CRCLM) is 36%-58%. Positron emission tomography (PET) tracers, imaging the expression of prognostic biomarkers, may contribute to assign appropriate management to individual patients. Aurora kinase A (AURKA) expression is associated with survival of patients after CRCLM resection. Methods: We synthesized [(3)H]alisertib and [(11)C]alisertib, starting from [(3)H]methyl nosylate and [(11)C]methyl iodide, respectively. We measured in vitro uptake of [(3)H]alisertib in cancer cells with high (Caco2), moderate (A431, HCT116, SW480) and low (MKN45) AURKA expression, before and after siRNA-mediated AURKA downmodulation, as well as after inhibition of P-glycoprotein (P-gp) activity. We measured in vivo uptake and biodistribution of [(11)C]alisertib in nude mice, xenografted with A431, HCT116 or MKN45 cells, or P-gp knockout mice. Results: [(3)H]Alisertib was synthesized with an overall yield of 42% and [(11)C]alisertib with an overall yield of 23%±9% (radiochemical purity ≥99%). Uptake of [(3)H]alisertib in Caco2 cells was higher than in A431 cells (P=.02) and higher than in SW480, HCT116 and MKN45 cells (P<.01). Uptake in A431 cells was higher than in SW480, HCT116 and MKN45 cells (P<.01). Downmodulation of AURKA expression reduced [(3)H]alisertib uptake in Caco2 cells (P<.01). P-gp inhibition increased [(3)H]alisertib uptake in Caco2 (P<.01) and MKN45 (P<.01) cells. In vivo stability of [(11)C]alisertib 90min post-injection was 94.7%±1.3% and tumor-to-background ratios were 2.3±0.8 (A431), 1.6±0.5 (HCT116) and 1.9±0.5 (MKN45). In brains of P-gp knockout mice [(11)C]alisertib uptake was increased compared to uptake in wild-type mice (P<.01) CONCLUSIONS: Radiolabeled alisertib can be synthesized and may have potential for the imaging of AURKA, particularly when AURKA expression is high. However, the exact mechanisms underlying alisertib accumulation need further investigation. Advances in knowledge and implications for patient care: Radiolabeled alisertib may be used for non-invasively measuring AURKA protein expression and to stratify patients for treatment accordingly.
Article
Methods: Wild-type and Abcb1a/b and/or Abcg2 knockout mice underwent (11)C-erlotinib PET/MR scans, with or without co-injection of a pharmacological dose of erlotinib (10 mg/kg) or after pretreatment with the ABCB1/ABCG2 inhibitor elacridar (10 mg/kg). Integration plot analysis was used to determine organ uptake (CLuptake) and biliary excretion (CLbile) clearances of radioactivity. Results: (11)C-erlotinib distribution to the brain was restricted by Abcb1a/b and Abcg2 and CLuptake into brain was only significantly increased when both Abcb1a/b and Abcg2 were absent (wild-type mice: 0.017±0.004 mL/min/g tissue, Abcb1a/b(-/-)Abcg2(-/-) mice: 0.079±0.013 mL/min/g tissue, P<0.001). Pretreatment of wild-type mice with elacridar increased CLuptake,brain to comparable levels as in Abcb1a/b(-/-)Abcg2(-/-) mice (0.090±0.007 mL/min/g tissue, P<0.001). Absence of Abcb1a/b and Abcg2 led to a 2.6-fold decrease in CLbile (wild-type mice: 0.025±0.005 mL/min/g tissue, Abcb1a/b(-/-)Abcg2(-/-) mice: 0.0095±0.001 mL/min/g tissue, P<0.001). There were pronounced differences in distribution of (11)C-erlotinib to brain, liver, kidney and lung and hepatobiliary excretion into intestine between animals injected with a microdose and pharmacological dose of erlotinib. Conclusion: ABCG2, ABCB1 and possibly other transporters influence in vivo disposition of (11)C-erlotinib and thereby affect its distribution to normal and potentially also tumor tissue. Saturable transport of erlotinib leads to non-linear pharmacokinetics which may compromise the prediction of erlotinib tissue distribution at therapeutic doses from PET with a microdose of (11)C-erlotinib. Inhibition of ABCB1 and ABCG2 is a promising approach to enhance brain distribution of erlotinib to increase its efficacy in the treatment of brain tumors.