You are on page 1of 13

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Articles

Cite This: ACS Chem. Biol. 2018, 13, 2758−2770 pubs.acs.org/acschemicalbiology

Quantitative Live-Cell Kinetic Degradation and Mechanistic Profiling


of PROTAC Mode of Action
Kristin M. Riching,† Sarah Mahan,† Cesear R. Corona,‡ Mark McDougall,‡ James D. Vasta,†
Matthew B. Robers,† Marjeta Urh,† and Danette L. Daniels*,†

Promega Corporation, 2800 Woods Hollow Road, Madison, Wisconsin 53711, United States

Promega Biosciences Incorporated, 277 Granada Drive, San Luis Obispo, California 93401, United States
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: A new generation of heterobifunctional small


molecules, termed proteolysis targeting chimeras (PROTACs),
targets proteins for degradation through recruitment to E3 ligases
Downloaded via 83.132.123.215 on September 20, 2019 at 14:51:11 (UTC).

and holds significant therapeutic potential. Despite numerous


successful examples, PROTAC small molecule development remains
laborious and unpredictable, involving testing compounds for end-
point degradation activity at fixed times and concentrations without
resolving or optimizing for the important biological steps required
for the process. Given the complexity of the ubiquitin proteasomal
pathway, technologies that enable real-time characterization of
PROTAC efficacy and mechanism of action are critical for accelerating compound development, profiling, and improving
guidance of chemical structure−activity relationship. Here, we present an innovative, modular live-cell platform utilizing
endogenous tagging technologies and apply it to monitoring PROTAC-mediated degradation of the bromodomain and extra-
terminal family members. We show comprehensive real-time degradation and recovery profiles for each target, precisely
quantifying degradation rates, maximal levels of degradation (Dmax), and time frame at Dmax. These degradation metrics show
specific PROTAC and family member-dependent responses that are closely associated with the key cellular protein interactions
required for the process. Kinetic studies show cellular ternary complex stability influences potency and degradation efficacy.
Meanwhile, the level of ubiquitination is highly correlated to degradation rate, indicating ubiquitination stemming from
productive ternary complex formation is the main driver of the degradation rate. The approaches applied here highlight the
steps at which the choice of E3 ligase handle can elicit different outcomes and discern individual parameters required for
degradation, ultimately enabling chemical design strategies and rank ordering of potential therapeutic compounds.

T argeting key drivers of disease for loss or degradation has


been a desirable outcome for numerous therapeutic
treatments and is often achieved indirectly by modulation of
which could be used as recruiter molecules for PROTAC
design.4,5,12,14,16−20 Of those, currently only two proteins, von
Hippel Lindau (VHL) and Cereblon (CRBN), both of which
upstream signaling pathways, transcriptional programs, and/or are substrate adaptor components of larger E3 Cullin-RING
epigenetic events.1−3 The first examples of compounds complexes, have shown significant and broad success against a
affecting degradation were demonstrated many years ago diverse set of targets.14,17,21−31
using small molecules or peptides to bridge interactions Chemical conversion of previously characterized inhibitors
between a target protein with components of the ubiquitin to PROTAC degradation compounds have shown therapeutic
proteasomal system (UPS).4−8 More recently, the approach advantages, including improved potency and prolonged
has been refined for efficacy and efficiency, comprising a pharmacodynamics,14,21,23,24,28,30,32−35 and in some cases
heterobifunctional compound termed proteolysis targeting have unexpectedly introduced specificity not observed in the
chimera (PROTAC), also known as SNIPER or degrono- parental pan-selective inhibitor.26,30,36,37 Despite these early
mid.2−6,9−15 Chemically, these compounds consist of a small successes, multiple challenges still exist in the functional
molecule target binder or inhibitor on one side which is characterization of PROTACs, including assessment of cellular
chemically linked to an E3 ligase recruiter compound on the permeability, ternary complex formation (target:PROTAC:E3
other side1,2 (Figure 1A). PROTACs induce degradation by ligase component), functional ubiquitination, and degradation.
simultaneously binding the target protein and the E3 ligase To optimize the likelihood of finding an active compound, an
complex proteins, bringing the target protein into proximity for extensive chemical series is typically developed with two main
ubiquitination and targeting it for degradation through the
UPS1,2 (Figure 1A). While there are hundreds of E3 ligases in Received: July 25, 2018
eukaryotes and hundreds more proteins involved in active E3 Accepted: August 23, 2018
ligase complexes, very few have known binding compounds Published: August 23, 2018

© 2018 American Chemical Society 2758 DOI: 10.1021/acschembio.8b00692


ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

Figure 1. Generation of HiBiT-tagged endogenous BET family members. (A) Schematic showing strategy for monitoring PROTAC-mediated
degradation of endogenously tagged HiBiT-BET family in live cells. HiBiT was inserted via CRISPR/Cas9 to the N-terminus of BET family
proteins BRD2, BRD3, and BRD4 in a HEK293 cell line stably expressing LgBiT, which complements with HiBiT to form the luminescent
NanoBiT luciferase. Following treatment with PROTACs, protein degradation is monitored by the loss of luminescence. Ternary complex
formation or ubiquitination is measured with NanoBRET using the HiBiT-BET protein complemented with LgBiT as an energy donor and
HaloTag fused to an E3 ligase component or ubiquitin as the respective energy acceptor. (B) HiBiT blot of CRISPR edited clonal cell lines
expressing endogenous HiBiT-BRD2, HiBiT-BRD3, or HiBiT-BRD4. Single isoforms of BRD2 and BRD3 were found, while both short and long
isoforms of BRD4 were detected. (C) PROTAC-induced degradation of HiBiT-BET family members after 4 h treatment with 1 μM dBET1 or
MZ1. Data are represented as mean RLU values (n = 3) of representative experiments. (D) Bioluminescence imaging using an Olympus LV200
microscope of endogenously tagged HiBiT-BRD4 in LgBiT expressing HEK293 cells treated with 1 μM MZ1 for 2 h.

variables: linker composition and the E3 ligase complex (NanoBRET)42,43 allows for kinetic measurements of intra-
recruitment handle.2−4,12,16,19 Compounds are then screened cellular protein interactions along the degradation pathway
for function by assessing target protein levels ± PROTAC such as ternary complex formation, ubiquitination, and
treatment at a given time and various concentrations using PROTAC-target engagement (Figure 1A). In these studies
end-point techniques such as Western blots or mass we apply this approach to characterize the cellular mechanism
spectrometry. These approaches, however, cannot be done in of action of two PROTACs, dBET1 and MZ1, which were
live cells, lack adaptability for high-throughput screening, and shown to degrade bromodomain and extra-terminal (BET)
are difficult to perform in a quantitative fashion. More family members BRD2, BRD3, and BRD4. 14,30 Both
importantly, if degradation is not observed in these measure- compounds contain the pan-BET inhibitor JQ1,44 serving as
ments, they provide no information as to the point of failure or the target-recruiting moiety, conjugated via an amide bond on
which steps in the process could benefit by compound the same attachment point.14,30 Similar to other BET family
optimization through further structure−activity relationship PROTACs, both MZ1 and dBET1 demonstrate greater
(SAR) efforts. potency as degradation compounds relative to parental BET
In this study, we present a modular live-cell platform to inhibitors.14,17,30,32,34,35,44−46 These two PROTACs were
monitor the critical steps of PROTAC mode of action (Figure chosen for this study due to availability, the use of the same
1A). At the core of the platform, we combine CRISPR/ JQ1 ligand with identical conjugation but to different E3
Cas938,39 endogenous tagging and luminescent technology40,41 handles, and because they are reported to have differential
to kinetically measure target protein levels with high precision degradation patterns.14,30,37 dBET1, which recruits CRBN
and without disruption of native expression levels or through a thalidomide handle, shows equivalent degradation
transcriptional regulation (Figure 1A). Coupling this technol- between BRD2, BRD3, and BRD4.14 Recent structural studies
ogy with optimized bioluminescence resonance energy transfer with related variants of dBET1 show noncooperative binding
2759 DOI: 10.1021/acschembio.8b00692
ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

Figure 2. BET family member kinetic degradation profiles and quantitation of degradation parameters. (A) HEK293 cells containing endogenously
tagged HiBiT-BET family member and coexpressing LgBiT were treated with an 8-point concentration series of 1 μM dBET1 or MZ1 added at t =
0. Luminescence (RLU) was continuously monitored in 5 min intervals over a 24 h time period in the presence of a stabilized furimazine substrate,
Endurazine. Profiles are plotted as mean fractional RLU values by normalizing to DMSO control. Error bars are represented as SEM of n = 3
experiments. (B) Representative schematic and equations to determine degradation parameters: degradation rate, Dmax and time at Dmax. Summary
of all degradation parameters tabulated for each BET family member treated with the various concentrations of dBET1 and MZ1 from kinetic
degradation profiles in (A). The initial degradation portion of each concentration curve was fit using GraphPad Prism to a one-component
exponential decay model (red line), and best fit parameters ƛ (degradation rate) and plateau (minimum remaining fraction) were obtained. The
maximum degraded fraction, Dmax, was calculated as 1− plateau. Time at Dmax was defined as the amount of time during which protein levels
remained below the threshold plateau + 10% of Dmax. (C) Degradation rate and (D) Dmax values expressed as percent degradation from (B) plotted
against PROTAC concentration for each family member and PROTAC. (D) DC50 values were obtained by fitting Dmax values to a variable slope
dose−response model in GraphPad Prism. Variability is expressed as SD of the mean from n = 3 experiments. (E) Time at Dmax, as determined by
the formula in (B), for HiBiT-BET family members treated with 1 μM dBET1 or MZ1 PROTAC. Data are represented in singlicate taken from the
combined (n = 3) kinetic traces.

between BET family members with CRBN.47 In contrast, structural and biophysical studies indicate cooperative binding
MZ1, which recruits VHL through a VHL ligand handle,48 between VHL and a bromodomain of BRD4 in the presence of
shows preferential degradation of BRD4.30 Supporting MZ1.37 These structural and biochemical studies yield
2760 DOI: 10.1021/acschembio.8b00692
ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

considerable and meaningful insights into the stability and scope showed highly uniform expression, nuclear localization,
formation of ternary complexes in vitro but are done with and rapid degradation of HiBiT-BRD4, as evidenced by loss of
isolated BET bromodomains and limited components of the luminescence over 2 h in the presence of 1 μM MZ1 (Figure
respective multiprotein E3 ligase complexes.37,47 Full-length 1D). Together, these data show expected physiological and
BET family members are large proteins, however, each PROTAC responses of the endogenously tagged HiBiT-BET
contains two bromodomains capable of separately binding proteins, indicating insertion of the HiBiT tag at their
compound.44,45 Therefore, in the cell, positioning of the family respective genomic loci did not impact function.
members within the E3 ligase complexes could be very Determination of Degradation Profiles and Quanti-
different, potentially leading to critical differences in fication of Key Degradation Parameters. To investigate
degradation activities as well. Here, we complement the the time-dependent effects of PROTAC-mediated degradation
biochemical work and provide additional live-cell context by on each BET family member, comprehensive degradation
characterizing the kinetic mechanistic differences mediated via profiles of biological triplicates were generated by monitoring
the two possible E3 ligase ternary complexes as well as further luminescence in 5 min continuous intervals over a 24 h period
understanding of family member specificities. Significant (Figure 2A). Treatment of HiBiT-BRD2, BRD3, and BRD4
insights into the degradation process and differential modes with an 8-point dilution series ranging from 1 nM to1 μM
of action of MZ1 and dBET1 are revealed and also show how dBET1 or MZ1 showed this concentration window resulted in
these technologies could be applied broadly to enhance a broad range of responses from very little degradation to near
understanding of any PROTAC or process which impacts complete degradation as detected by luminescence (Figure
protein homeostasis. 2A). Strikingly, each family member showed a unique

■ RESULTS AND DISCUSSION


Generation of Endogenously Tagged HiBiT-BET
signature, which was both PROTAC- and concentration-
dependent (Figure 2A). The degradation profiles in general
contained three phases: an initial degradation curve showing
Family Members. Protein overexpression can impart defects loss of the target, a Dmax, and then an eventual upward rise
on normal protein turnover and homeostasis; therefore, we indicating target recovery. Recovery curves were highly
developed a means to monitor protein levels in real-time variable, often multiphasic, and not quantifiable. Fortunately,
without disruption of endogenous expression or regulation. To other parameters such as initial degradation rate, Dmax, and
this end, we used CRISPR/Cas9 genome editing38,39 to time at Dmax as depicted in Figure 2B were all amenable for
append an 11 amino acid peptide, termed HiBiT,41 to the N- quantitation from these profiles.
terminus of BET family members natively expressed in To determine degradation rate and D max at each
HEK293 cells, BRD2, BRD3, and BRD4 (Figure 1A and concentration, only the starting curve representing the time
Supporting Information Table 1). The HiBiT peptide, which frame to maximal degradation was considered, as defined in
shows high efficiency for CRISPR insertion given its small size, Figure 2B. This initial curve fit a single-component exponential
has pM affinity for and spontaneously complements with the decay model, yielding parameters for cellular degradation rate
18 kDa LgBiT protein, forming the luminescent and bright and Dmax for concentrations of each PROTAC and BET family
NanoBiT luciferase41 (Figure 1A). LgBiT can be introduced member summarized in Figure 2B. Trends showed that
for luminescent detection of HiBiT fusion proteins in a variety increasing concentration of PROTAC increased rate of
of ways, but to enable live cell detection, we performed degradation, eventually reaching a plateau at higher concen-
CRISPR endogenous tagging in HEK293 cells stably trations (Figure 2C). This trend indicates that beyond the
expressing LgBiT. CRISPR clonal selection was performed, plateau, concentration of compound is not rate-limiting.
and genomic analysis by Sanger sequencing revealed Moreover, it is expected that very high concentrations will
homozygous HiBiT allelic insertions for BRD3 and BRD4, become inhibitory, consistent with observed hook effects for
and a heterozygous insertion for BRD2 (data not shown). PROTACs17,30 Degradation rate plots revealed emergence of
Confirmation of size and expression for each HiBiT-BET family member differences, particularly for MZ1 treatment,
family member was confirmed by SDS-PAGE with luminescent which showed fast degradation of BRD2 and BRD4, while
blotting as described in the Materials and Methods. Single BRD3 exhibited much slower degradation (Figure 2C). Such
bands for BRD2 and BRD3 were observed, while both long family differences were not observed with dBET1, which
and short isoforms of BRD4 were detected, as expected given showed similar rates of degradation for BRD2, BRD3, and
the placement of HiBiT on the N-terminus (Figure 1B). BRD4 (Figure 2C). These results demonstrate how not only
Comparative luminescence expression levels of each family PROTAC-mediated recruitment to either VHL or CRBN E3
member showed similar expression levels of BRD2 and BRD4, ligase complexes can impact the degradation rate, but also how
while expression of BRD3 was approximately 60% lower different family members can show varying rates to the same
relative to these (Figure 1C and Supporting Information Table PROTAC (Figure 2C).
2). End-point analysis of each endogenously tagged HIBiT- Despite having differences in the initial degradation rate,
BET family protein treated with 1 μM dBET1 or MZ1 showed family members were able to achieve similar Dmax if given
substantial degradation at 4 h (Figure 1C). To show specificity enough time, and total amount of degradation increased with
of degradation, HiBiT-BRD4 cells treated with either dBET1 increasing concentrations of PROTAC (Figure 2A and B).
or MZ1 for 3 h were then treated with a 10-fold molar excess Previous dose response curves (DRCs) to determine cellular
of JQ1 (Supporting Information Figure 1). This resulted in DC50 potency of PROTAC compounds have calculated
increased rates of recovery of BRD4 in the presence of JQ1, degradation at fixed time points,14,30,37,47 overlooking the
the parental competitive binding inhibitor which does not time dependency to reach Dmax (Figure 2A). We find that for
induce degradation (Supporting Information Figure 1). To any given family member at any concentration, the time at
assess proper localization and response to PROTAC treatment, which Dmax is reached varies greatly at each concentration
bioluminescence imaging using an Olympus LV200 micro- (Figure 2A). Therefore, DRCs at fixed times could easily
2761 DOI: 10.1021/acschembio.8b00692
ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

misrepresent DC50 potency values depending on the time found at both 100 nM and 10 nM treatments (Supporting
chosen for the measurement. For example, measuring Information Figures 3A and B). Furthermore, the time at Dmax
degradation of BRD2 at 2 h would show either PROTAC is for each family member followed the trends in family member
potent for degradation, but at 20 h it would appear that neither potency for a particular PROTAC (Figure 2D and E). For
PROTAC had any effect, because degradation had already example, BRD4 showed greatest MZ1 potency and greatest
occurred and transitioned to recovery (Figure 2A). With these time at Dmax, while BRD2 had lowest MZ1 potency and
understandings and to remove the impact of time, we sought a shortest time at Dmax, despite having a fast initial degradation
more accurate method to calculate DC50 values by determining rate (Figure 2C, D, and E).
Dmax at every concentration irrespective of time (Figure 2B and These data suggest many factors play a role in time at Dmax
D). This approach reflects and takes into account the actual and recovery. The finding that BET family members show
cellular response and the true capacity for degradation. longer times at Dmax after treatment with MZ1 as compared to
Comparison of DC50 values calculated in this fashion allowed dBET1 (Figures 2A, B, and E and Supporting Information
for rank ordering potency, with MZ1 showing BRD4 > BRD3 Figure 3) support previous reports of reduced cellular
> BRD2 and with dBET1, BRD3 > BRD4 > BRD2 (Figure compound stability of dBET1.14 The differential family
2D). The window of family member potencies exhibited a member response however to the same PROTAC (Figure
broader range with MZ1 (9X) as compared to dBET1 (4X), 2A, B, and E and Supporting Information Figure 3) indicate
but significant differences in degradation capabilities only other events are impacting these phases. This is most apparent
emerged at very low concentrations of PROTAC (Figure 2D), for BRD2, which shows faster recovery as compared to BRD3
matching previous studies with the endogenous proteins and BRD4 (Figure 2A); therefore, we investigated whether
analyzed by Western blotting.30 inhibition alone of BET proteins could lead to compensatory
As other PROTAC cellular studies have utilized ectopic changes in BRD2 protein levels. To do this, we performed
expression of fluorescent proteins to determine compound kinetic analysis of HiBiT-BRD2 after treatment with JQ1
degradation potencies,47 we wanted to examine whether this alone. In these studies, we saw a rapid and significant increase
approach influenced the kinetic degradation rate and Dmax. To in HiBiT-BRD2 levels (Supporting Information Figure 4A),
do so, comparative experiments were performed with Nano- not observed for HiBiT-BRD3 and HiBiT-BRD4 (Supporting
Luc40 (the parental luciferase of the HiBiT+LgBiT comple- Information Figures 4B and C) similarly treated with JQ1,
mentation technology) fused to BRD4 and transiently indicating potential feedback on BRD2 synthesis as a
transfected in HEK293 cells. In comparison to the profile for consequence of BET inhibition. These results and others
endogenous HiBiT-BRD4 treated with MZ1 in Figure 2A, that have shown a similar effect on endogenous BRD2 by mass
ectopically expressed NanoLuc-BRD4 treated with MZ1 shows spectrometry49 suggest that the rapid increase in BRD2
a significantly different degradation signature (Supporting following BET family degradation may be due to compensa-
Information Figure 2A). Overall, the degradation window was tory feedback mechanisms, and demonstrate how these can
compressed, showing only a small population of the NanoLuc- outcompete degradation to drive faster recovery of specific
BRD4 was degraded, even at high concentrations of MZ1 family members over others.
(Supporting Information Figure 2A). Determination of the Kinetic Monitoring of Ternary Complex Formation
degradation rate showed degradation of the ectopic NanoLuc- and Ubiquitination. To relate the degradation profiles to the
BRD4 was slower (Supporting Information Figure 2B), and specific interactions in the degradation pathway, we inves-
the DC50 showed a 30× reduction in apparent potency tigated the kinetics of intracellular ternary complex formation.
(Supporting Information Figure 2C). Moreover, the profiles To do so, we utilized proximity-based NanoBRET technol-
showed more rapid protein recovery, not matching the ogy42 comprising the same endogenously tagged HiBiT-BET
prolonged time at Dmax observed for endogenously tagged family members complemented with LgBiT as luminescent
BRD4 (Figure 2A). Because constitutive expression from a energy donors, and ectopically expressed HaloTag-CRBN or
non-native promoter lacks the relevant epigenetic and HaloTag-VHL fusions as energy acceptors. (Figure 1A). Due
transcriptional control found at the endogenous target loci to its ratiometric nature,42 NanoBRET is particularly suited for
and ectopic expression will produce excess target protein levels these studies as loss of the donor (target protein) will not
in the cells, degradation profiles cannot be meaningfully impact the readout of ternary complex formation (Supporting
interpreted (Supporting Information Figure 2A), and quanti- Information Figure 5), and is capable of detecting transient
tative parameters will be skewed (Supporting Information interactions. To first demonstrate ternary complex specificity
Figures 2B and C). by NanoBRET, we examined ternary complex formation of
While the degradation rate and DC50 values are readouts of ectopically expressed NanoLuc-BRD4 with either HaloTag-
efficiency and potency, they do not capture the time at Dmax in VHL or HaloTag-CRBN in the presence of either PROTAC.
the cells, which speaks to the time of sustained compound Only the specific pairings of NanoLuc-BRD4 with the
efficacy. To quantify this time frame, we defined the time at respective E3 ligases and PROTACs, showed a dose-depend-
Dmax as the period in which protein levels were within a ent increase in ternary complex formation, as measured by
threshold window of 10% of Dmax: (plateau + 0.1(Dmax)) NanoBRET (Supporting Information Figure 5A). While
(Figure 2B). Calculation of time at Dmax in this fashion showed detection of ternary complexes within short periods of time
this parameter increased with increasing concentration of (30−60 min) was possible, longer term kinetic analyses were
PROTAC, with a few outliers were observed at the extremely challenging due to the significant loss of the target proteins. To
low levels of degradation where the threshold window mitigate this loss and study the ternary complex stability over
approached the level of experimental noise (Figure 2B). several hours, MG132, a proteasomal inhibitor, was added to
From this analysis, we found that MZ1 treatment showed prevent degradation, and resulted in a more robust assay
longer times at Dmax for all BET family members as compared window (Supporting Information Figures 5B and C). Because
to dBET1 treatment at 1 μM (Figure 2E), with similar trends MG132 can be toxic to cells, we tested cell viability by
2762 DOI: 10.1021/acschembio.8b00692
ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

Figure 3. NanoBRET ternary complex formation and ubiquitination of BET family members. (A and B) HEK293 cells containing endogenously
tagged HiBiT-BET family members and expressing LgBiT were transiently transfected with either HaloTag-VHL or HaloTag-CRBN (A) or
HaloTag-Ubiquitin (B), which encodes for the first ubiquitin repeat, amino acids 1−76, of the polyubiquitin-B precursor protein. Forty-eight hour
post-transfection, NanoBRET ratios were calculated to measure intracellular ternary complex formation (A) or target ubiquitination (B) in the
presence of 1 μM dBET1 or MZ1 (t = 0). Kinetic monitoring of NanoBRET was performed for 6 (A) or 2 (B) h in the presence of both stabilized
furimazine substrate, Vivazine, and NanoBRET618 fluorescent ligand. Experiments were also set up for all assays without NanoBRET618
fluorescent ligand as a control and allowed for subtraction of background within the NanoBRET assay, as detailed in the Materials and Methods.
Data are represented as fold increase in NanoBRET by normalizing to DMSO control. Variability expressed as SEM from n = 3 experiments. (C)
HEK293 cells containing endogenously tagged HiBiT-BET family members and expressing LgBiT were treated at the indicated times with 1 μM
dBET1 or MZ1 PROTACs. Cells were lysed with digitonin and incubated for 10 min with both primary polyclonal anti-Ubiquitin and Alexa-594
fluorescent secondary antibody to determine NanoBRET ratios. Data are represented as fold increase in NanoBRET by normalizing to t = 0 time
point. Variability expressed as SEM from n = 3 experiments. (D) Degradation rate at 1 μM PROTAC from Figure 2C is plotted against
ubiquitination fold increase at the 1 h time point shown in (C).

CellTiter-Glo after 10 μM MG132 treatment over 6 h and members treated with dBET1, ubiquitination signals were
found no impact on cell number (Supporting Information similar in fold change over time and in general slower as
Figure 5D). We then performed kinetic analysis of ternary compared with MZ1 treatments (Figure 3B). As an orthogonal
complex formation in the presence of MG132 for 6 h with approach, NanoBRET was also used to probe endogenous
endogenously tagged HiBiT-BET family members comple- ubiquitination of HiBiT-BET family members, utilizing a
mented with LgBiT. Complexes between all BET family polyclonal ubiquitin primary antibody and Alexa594-con-
members and CRBN in the presence of dBET1 showed very jugated secondary antibody as an acceptor. These data show
rapid association kinetics, peaking at 1 h and then either identical trends to those observed in the live-cell kinetic
holding steady or declining (Figure 3A). In contrast, experiments with HaloTag-Ubiquitin (Figure 3C). The
complexes of BET family members with VHL and MZ1 different NanoBRET ratios obtained for the same target
exhibited slower association kinetics, yet continued to increase treated with either MZ1 or dBET1 indicate that ubiquitination,
over time showing longer complex stability (Figure 3A). The either in pattern or extent, is different dependent on the E3
prolonged stability of the ternary complexes mediated by MZ1 ligase complex to which it is recruited. The degradation rates
compared to dBET1 for all BET family members agreed well determined in Figure 2C show striking similarity to the trends
with the trend that MZ1 treatment resulted in longer time at in ubiquitination for respective family members and
Dmax (Figures 2A and E), indicating ternary complex stability PROTACs. Plotting degradation rates against the relative
directly influenced the duration of degradation efficacy. increase in ubiquitination at the same concentration of
Because our readout of ternary complex formation is only a PROTAC revealed a positive and linear correlation, indicating
measurement of binding and not activity of the complex, we the step of ubiquitination is the predictive step of degradation
measured downstream target ubiquitination. Intracellular and controls the rate (Figure 3D).
ubiquitination kinetics were measured by NanoBRET using Lytic and Live Cell Target Engagement and Cell
the same HiBiT-BET family members as donors, but paired Permeability Assessment. Given the large molecular size of
with HaloTag-Ubiquitin as an acceptor. Robust and rapid heterobifunctional degraders, a significant step in the process
ubiquitination of BRD2 and BRD4 was observed with MZ1 of chemical optimization is characterization of cell perme-
treatment, while BRD3 showed both slower kinetics and ability, target occupancy, and binding affinities to both the
reduced ubiquitination levels (Figure 3B). For BET family target and E3 ligase complex. To facilitate these studies, we
2763 DOI: 10.1021/acschembio.8b00692
ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

wanted to transition from monitoring protein:protein inter- VHL protein (Kd = 66 nM for MZ137 and 90 nM for
actions to protein:small molecule interactions. To this end, we VH29848). These results, however, were not recapitulated in
applied NanoBRET technology using ectopic NanoLuc fusions live cells, with MZ1 showing a >40-fold reduction in binding
as donors and fluorescently labeled small molecule tracers as affinity relative to that of VH298 (Figures 4C and F).
acceptors.43 This approach interrogates the separate events of Similar live cell target engagement studies were done with
PROTAC binding to either target or E3 ligase component NanoLuc-CRBN and a CRBN tracer and showed a 10-fold
(Figure 4A), with a key advantage in that it can be performed reduction in binding of dBET1 as compared to that of
in both live cell or lysate formats to assess permeability.43 thalidomide (Supporting Information Figures 6A−C). To
Competitive displacement of the VHL tracer molecule interrogate the live cell binding affinities and kinetics to BRD4
revealed identical binding affinities between MZ1 and a related by these PROTACs, competitive displacement studies were
potent monovalent VHL binder, VH298,48 in lytic format done using a BET family tracer with NanoLuc-BRD4. Results
(Figure 4B), in agreement with biophysical measurements in showed dBET1 and MZ1 had an 8- and 15-fold right-shifted
vitro which showed identical binding affinity with recombinant reduction in binding affinity, respectively, compared to that for
JQ1 in live cells (Figures 4D and F). Intracellular kinetics were
performed and found that both PROTACs had significantly
slower target engagement compared to JQ1 in live cells and
apparent IC50 values which decreased over time (Figure 4E).
Together, these results indicate reduced and slow cellular
permeability of MZ1 and dBET1, yet they remain highly
efficacious for degradation at substoichiometric binding affinity
concentrations, supporting a catalytic mechanism of action by
these PROTACs.
Response of cMyc in MV4;11 Cell Lines to Inhibitors
and PROTAC Treatments. One of the primary consequences
and desired outcomes of BET family inhibition is down-
regulation of key oncogenic targets, including cMyc.44,45 To
quantitatively study cMyc levels post-treatment with BET
inhibitors as compared to PROTACs, we tagged endogenous
cMyc with HiBiT on the C-terminus using CRISPR/Cas9 in a
relevant leukemia line, MV4;11 (Supporting Information Table
1). Treatment of these cells with 10 nM JQ1, dBET1, or MZ1
over a 24 h time period showed differential levels of cMyc
expression knockdown, from minimal loss with JQ1, to 50 and
80% loss with dBET1 and MZ1, respectively (Figure 5A).
Correlation to cell viability assays at each time point
revealed that only treatment with MZ1 resulted in measurable
cellular death at this concentration (Figure 5B). At a higher
concentration, both compounds showed significant reduction
in cMyc expression and concurrent cellular death (Figures 5C
and D). While cMyc is one of many transcriptional targets
impacted by BET inhibition44,45 and PROTAC-mediated
degradation,14 the use of HiBiT endogenous tagging as a
downstream reporter of function allowed for understanding of
the amount of loss needed to elicit the desired phenotypic
outcome.
Figure 4. NanoBRET target engagement with PROTAC compounds.
(A) Schematic showing NanoBRET target engagement strategy with
either E3 ligase component or BRD4 using NanoLuc fusions as
energy donors and fluorescent small molecule tracers as energy
■ CONCLUSION
Generation of heterobifunctional PROTAC compounds, either
acceptors, which can be competitively displaced by unlabeled de novo or from known inhibitors, presents an exciting new
compounds. (B and C) Competitive displacement profiles of direction for drug discovery.1,2,10,12 BET family PROTACs,
HEK293 cells transiently transfected with NanoLuc-VHL and including those studied here, have shown advantages over BET
incubated with the VHL fluorescent tracer in the presence of serial inhibitors and hold great promise for potential treatment of a
titrations of either MZ1 or VH298 in digitonin-permeabilized lysate wide variety of diseases, including cancer and inflamma-
(B) or live cells (C). (D) Live cell competitive displacement profiles tion.14,17,30,32,34,35,44,45 In this work, we further understanding
of HEK293 cells transiently transfected with NanoLuc-BRD4 and of the cellular mechanism of action of dBET1 and MZ1 using a
incubated with the BRD4 fluorescent tracer in the presence of serial suite of cell-based assays combined with CRISPR/Cas938,39
titrations of either MZ1, dBET1, or JQ1 in live cells. Data are endogenous tagging, allowing for precise and quantitative
represented as NanoBRET ratios normalized to zero compound.
Error bars are expressed as SD of the mean (n = 3) of a representative
kinetic analysis of key degradation events. Importantly, we also
experiment. (E) NanoBRET kinetic monitoring of IC50 values for assess which stages of the degradation process are impacted by
JQ1, dBET1, and MZ1 binding to BRD4 in live cells. Data are recruitment to different E3 ligase complexes, a critical
expressed as singlicate IC50 values from RLU (n = 3) of a consideration and variable in the development of PROTAC
representative experiment. (F) Tabulated IC50 values for indicated compounds. Currently the primary E3 ligase components
target, PROTAC, and assay format. utilized are VHL and CRBN, but active research efforts are
2764 DOI: 10.1021/acschembio.8b00692
ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

Figure 5. PROTAC modulation of cMyc-HiBiT expression and viability in MV4;11 cells (A and B) MV4;11 cells containing endogenously tagged
cMyc-HiBiT were treated 10 (A) or 100 (B) nM JQ1, dBET1, or MZ1. (C and D) Cell viability was measured by CellTiter-Glo following
treatment with 10 (C) or 100 (D) nM for each respective compound. All data are represented as RLU normalized to DMSO control. Error bars are
expressed as SD of the mean (n = 3) of a representative experiment.

Figure 6. Model of degradation phases and contributing mechanisms Schematic showing the three phases of degradation: initiation of degradation,
degradation maximum (Dmax), and recovery correlated to the key mechanistic processes listed in black for each phase. The target protein response
to PROTACs at each phase is represented pictorially, showing first introduction of PROTAC and target protein, followed by loss of target protein,
and then recovery. Listed in red are the parameters identified to regulate each phase with arrows depicting the processes which to monitor or
optimize.

underway to identify other efficacious E3 ligase components to complex with MZ137 or related dBET compounds47 and their
serve as recruiters.2−4,12,16,19 This will result in an expansion of respective E3 ligase partners. In the case of MZ1/VHL,
options in the future, directly translating to increased chemical positive cooperativity in binding was observed,37 in agreement
development and furthering the need for robust and relevant with our finding that these complexes were more stable over
technologies to efficiently profile and triage PROTAC cellular time in cells. In contrast, for related dBET compounds with
activity. CRBN, a lack of cooperativity as well as multiple
To fully characterize and compare PROTAC-mediated conformations have been reported,47 suggestive of more
degradation via the VHL or CRBN E3 ligase complexes, dynamic and transient ternary complex formation for this
quantifiable parameters of the cellular degradation profile were system. Indeed, we see this behavior mirrored inside the cell
determined in these studies. Across all BET family targets with complexes forming quickly but not showing increasing
studied here, we found that the MZ1/VHL recruitment system stabilization over time with any of the BET family members.
elicits faster rates of degradation, slower, but more stable Our studies were done in HEK293 cells, and it is worth
ternary complex formation, enhanced ubiquitination, greater noting that degradation profiles could vary in other cell lines
compound potency, and longer-lasting degradation efficacy as that may have different expression levels of E3 ligase
compared to the dBET1/CRBN system. Because both components or BET family members. However, we would
compounds contain JQ1 as the target specific handle, we not expect overall trends to be different because ternary
conclude these differences are directly attributed to ternary complex stability and activity of specific complexes would likely
complex stability and relative positioning within the different be the same. Because thorough studies comparing functional
E3 ligase systems of each BET target protein for ubiquitina- outcomes imparted by use of different E3 ligase handles have
tion. Our ternary complex kinetic data correlate well with not been previously performed, it is difficult to predict whether
recent structural studies of bromodomains of BRD4 proteins in changes in the E3 recruitment system will be favorable or
2765 DOI: 10.1021/acschembio.8b00692
ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

unfavorable for any given target. Additional BET family capabilities of high sensitivity over a broad dynamic range.41 In
PROTACs,17 including more potent versions of dBET1, have our studies, we were are able to accurately and continuously
been reported.47,50 It would be intriguing to determine which monitor degradation over a 3-log range with high reproduci-
cellular steps and parameters contribute to the improvement in bility, indicating this approach can enable detection of proteins
potency for this class of compounds, and then translate this that have low levels of native expression and/or for
understanding to help guide further chemical design as well as characterization of early development PROTAC compounds,
to enhance SAR for BET degraders in general. which might not show fast or robust degradation.
More broadly, we seek to use the functional relationships Other approaches are currently also used to study
determined from analysis of dBET1- and MZ1-mediated protein:protein and protein:small molecule interactions but
degradation to build a mechanistic map of primary do not have the same advantages for degradation research as
contributing events for each phase of the PROTAC cellular compared to NanoBRET.42 Given its proximity-based and
degradation profile; degradation initiation, Dmax, and recovery ratiometric nature, NanoBRET is suited for studying
(Figure 6). This model can be used as a guide for compound interactions that are transient, and measurement of the
rank-order profiling, design, and optimization, showing which interaction is not influenced by changes in the level of donor
measurements will be most predictive in improving any given if this occurs.42 This is why we have specifically chosen the
phase in degradation. The important contributors to the initial target protein as the energy donor in the configuration of the
degradation phase are PROTAC compound permeability, assay, allowing for determination of complex formation even
binding affinities to targets and E3 components, and with active loss of the target. In contrast, other protein
recruitment not simply into a ternary complex, but a interaction technologies are reliant on a single readout, such as
productive complex to achieve ubiquitination. Ultimately the fluorescence or luminescent complementation and lack the
level of target ubiquitination, and not necessarily the efficiency ability to discern whether changes in signal are due to the
of ternary complex formation dictates the rate of degradation. modulation of the interaction or changes in levels of one or
For Dmax, we find, perhaps not unexpectedly, that this phase is both partners. A further advantage of NanoBRET is the dual
related to the long-term stability of the ternary complex and is readout capability, which allows simultaneous monitoring of
also highly dependent on any downstream cellular feedback to both the interaction (protein:protein or protein:small mole-
loss of the target which would promote recovery. This cule) by calculation of NanoBRET ratios as well as target loss
interplay is most evident in our data for BRD2, which shows by direct luminescent monitoring, provided the target protein
fast rate of degradation, stabilized ternary complex formation, is the donor.
but by far the shortest time at Dmax. We found this to be due to The platform of technologies presented here to characterize
competing upregulation of protein levels, a response stimulated the cellular mechanism of action of two BET bromodomain
by JQ1 treatment alone that appeared to be amplified in PROTACs, dBET1 and MZ1, not only advance our under-
response to PROTAC treatment from the added layer of BET standing of these chemical probes, but more broadly provide a
family degradation compared to inhibition. The protein comprehensive analysis to advance PROTAC drug discovery
recovery phase, which is challenging to study given its and cellular activity assessment. Our cellular analysis shows
multiphasic and more complex profiles, will likely vary greatly that both ternary complex stability and subsequent ubiquiti-
between targets, as we have found for BET family members, nation are key steps, but they influence and regulate different
and will also depend on the cellular chemical stability of the phases of degradation (Figure 6), correlations not previously
PROTAC compound. Together, these correlations help break described due to studies done primarily with in vitro assays.
down some of the barriers in properly assessing PROTAC These understandings combined with our technology platform
activity within the highly complex degradation process, to study each parameter in live cells will help further
providing useful entry points and important parameters to compound optimization, including fine-tuning family member
monitor for optimization of key steps required for degradation. specificity if desired. While the focus of this work was on
Our understanding of the degradation process, particularly PROTAC characterization, the approach presented here can
the quantitative and kinetic determination of the three phases be applied to study any process that impacts protein
shown in Figure 6, would not have been possible without the homeostasis, as demonstrated with cMyc (Figure 5) and in
use of the highly sensitive endogenous tagging system, our other studies of therapeutic targets HIF1α and β-catenin
HiBiT,41 and complementary intracellular protein interaction (data not shown). The diverse capabilities and applications of
studies using NanoBRET.42,43 Prior to our studies presented these technologies coupled with endogenous tagging via
here, understanding of PROTAC degradation profiles have CRISPR/Cas9 present a highly enabling and physiologically
been highly granular with limited quantitation, using primarily relevant means to study protein degradation, providing unique
Western blot analysis to observe target loss and recovery at insight into the individual and highly complex mechanisms
fixed time points and concentrations. Moreover, other regulating the degradation process.


approaches have focused on monitoring ectopically expressed
fusion tagged proteins or domains in the cells.47 While this
MATERIALS AND METHODS
allows for qualitative and visual determination of degradation
as well as study of protein interactions as we have shown here, Expression Plasmids. Clones expressing N-terminal HaloTag
its application for quantitative degradation analysis resulted in fusions of human VHL (NM̅ 000551) and Cereblon (AAH17419.1)
were obtained from Kazusa DNA Research Institute (Kisarazu, Japan)
alteration of degradation rate, reduced Dmax, right-shifted DC50
as pFN21A HaloTag CMV Flexi Vectors (Promega). The N-terminal
values, and non-native recovery profiles. Therefore, to maintain HaloTag-Ubiquitin constructed consisted of the first ubiquitin repeat,
endogenous expression level and regulation, the pairing of amino acids 1−76, of the human polyubiquitin-B precursor
HiBiT luminescent technology and CRISPR/Cas9 endogenous (NM̅ 018955) and was cloned into the pFN28K (Promega) without
tagging was critical. Compared to other tags that could also be appending any additional residues to the native C-terminal sequence.
used for endogenous insertion, HiBiT provides additional N-terminal NanoLuc fusions of human VHL (NM̅ 000551), Cereblon

2766 DOI: 10.1021/acschembio.8b00692


ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

(AAH17419.1), and BRD4, long isoform (NM̅ 058243) were cloned (20 mM Tris-HCl, pH 7.5, 150 mM NaCl, and 0.1% Tween 20) with
into pFN31K vectors (Promega). 1 mM dithiothreitol (DTT) before replacing buffer with NanoGlo
Reagents and Cell Culture. dBET1 was purchased from Cayman Blotting Buffer containing 100 nM purified LgBiT protein (Promega)
Chemical, JQ1 was purchased from Xcessbio Biosciences, MZ1 and and incubating at 4 °C overnight with shaking. NanoGlo Luciferase
VH298 were generously donated by A. Ciulli (University of Dundee), Assay Substrate (Promega) was added, and luminescence was
and MG132 was purchased from InvivoGen. HEK293 and MV4;11 captured on an ImageQuant LAS 4000 digital imaging system (GE
cells were obtained from the American Type Culture Collection and Healthcare Life Sciences).
cultured at 37 °C, 5% CO2 in DMEM (Gibco) or IMDM (American Bioluminescence Imaging. Clonal HiBiT-BRD4 edited cells
Type Culture Collection), respectively, containing 10% fetal bovine were plated in 8-well glass bottom chamber slides (Nunc Lab-Tek II)
serum (Seradigm). at a density of 8 × 104 per well in 200 μL of growth medium and
Generation of HEK293 LgBiT Stable Cell Line. HEK293 cells incubated at 37 °C, 5% CO2 overnight. Following incubation,
(8 ×106) plated in a 10 cm2 Petri dish were transfected with 20 μg medium was replaced with Opti-MEM (Gibco) containing 20 μM
LgBiT CMV plasmid for 24 h and grown in 2 mg mL−1 G418 Vivazine extended release substrate (Promega) and incubated for an
selection antibiotic (Promega) until large colonies were present. additional 1 h at 37 °C, 5% CO2 to allow equilibration of substrate
Clones were isolated by limiting dilution into 96-well plates and after and stabilization of luminescent signal. Real-time imaging was
expansion were screened for LgBiT expression by triplicate plating performed using the LV200 bioluminescence imaging system
with one plate receiving HiBiT Lytic Detection Reagent (Promega) (Olympus) equipped with an ImagEM X2 EM-CCD camera
supplemented with purified HiBiT-HaloTag and another plate (Hamamatsu) and a temperature-controlled stage. Images were
receiving CellTiter-Glo (Promega). A clone with high luminescence collected for a period of 2 h with cellSens software (Olympus)
relative to CellTiterGlo was expanded from the third plate and using a 60× oil-immersion objective, electron multiplying EM gain of
maintained in 800 μg/mL G418. 1200, and 10 s exposure times. Average projections of 10 sequential
CRISPR/Cas9-Editing. One nanomole Alt-R CRISPR RNA images were generated, and dynamic range adjustments and
(crRNA) and 1 nmol Alt-R trans-activating crRNA (tracrRNA) pseudocolor rendering was performed using FIJI.
were assembled in 50 μL of Nuclease-Free Duplex Buffer (Integrated Kinetic Live Cell HiBiT Detection. Clonal cell lines edited for
DNA Technologies) by incubation at 95 °C for 5 min and cooling to HiBiT insertion to BET family members were plated in white 96-well
RT to generate gRNA for BRD2, BRD4, and cMyc. CRISPRevolution tissue culture plates at a density of 2 × 104 cells per well in 100 μL of
sgRNA EZ kit (Synthego) was used as the gRNA for BRD3. Single- growth medium and incubated overnight at 37 °C, 5% CO2. For
stranded ultramer DNA oligonucleotides (IDT) were used as the experiments with transient expression of NanoLuc-BRD4, 8 × 105
ssODN donor templates. Ribonucleoprotein (RNP) complexes with HEK293 cells were first transfected with 0.02 μg of NanoLuc-BRD4
recombinant Streptococcus pyogenes Cas9 with an N-terminal tag diluted in 2 μg of promoterless carrier DNA using FuGENE HD
containing a histidine-nuclear localization signal-myc sequence (Promega), and the following day, were plated in white 96-well tissue
(Aldevron) were assembled as previously described41 by incubating culture plates at a density of 2 × 104 cells per well in 100 μL of growth
100 pmol of Cas9 and 120 pmol of gRNA for 10 min at ambient medium and incubated overnight at 37 °C, 5% CO2. For end point
temperature. For HiBiT knock-in to BRD2, BRD3, and BRD4, 2 × live cell detection, wells were treated with dBET1 or MZ1 compounds
105 HEK293 cells were resuspended in 20 μL of 4D Nucleofector for the indicated time frames before addition of NanoGlo Live Cell
solution SF, and RNP complexes along with 100 pmol ssODN Substrate (Promega) and reading luminescence on a GloMax
template were electroporated into cells with the 4D Nucleofector Discover. For continuous live cell detection out to 24 h, medium
System (Lonza) using program CM-130. For HiBiT knock-in to was replaced with CO2-independent medium (Gibco) containing 20
cMyc, MV4;11 cells were resuspended in 20 μL of 4D Nucleofector μM Endurazine, an extended time-released substrate (Promega), and
solution SE, and RNP complexes along with 100 pmol ssODN plates were incubated at 37 °C, 5% CO2, for 2.5 h before addition of a
template were electroporated into cells with the 4D Nucleofector 3-fold serial dilution of 1 μM final concentration dBET1, MZ1, or
System (Lonza) using program DJ-100. Immediately following JQ1 compounds. Plates retaining the plate lids were then read every 5
electroporation, cells were incubated at ambient temperature for 10 min for a period of 24 h on a GloMax Discover (Promega) set to 37
min before transferring to a 12-well plate for culturing. Edited pools °C.
were analyzed for HiBiT insertion by assaying for luminescence on a Quantitation of Degradation Kinetics. Degradation rate and
GloMax Discover (Promega) 48−72 h postelectroporation. degradation plateau were calculated by fitting only the initial
Generation and Validation of Clonal HiBiT Insertions. Clonal degradation portion of each kinetic concentration curve to the
populations of edited cells were obtained by sorting live singlets using equation:
a FACSMelody cell sorter (Becton Dickinson) into 96- or 384-well
plates. Clones were expanded and screened by luminescence, and y = (y0 − plateau)e−ƛt + plateau
genomic DNA was isolated using the Wizard Genomic DNA
Purification kit (Promega) from 8 to 10 clones per edited pool. where ƛ = degradation rate in units of hr−1. The degraded fraction,
Genomic DNA was amplified using primers designed to anneal ∼200 Dmax, was calculated as 1 − plateau. For each curve, the first few data
bases either upstream or downstream from the inserted ssODN points before onset of degradation were excluded from the fits. Time
template sequence, subcloned into a pF5K vector backbone at Dmax was determined by calculating the length of time during which
(Promega), transformed into JM109 cells (Promega), and 24 the fractional population remained below a threshold defined as
individual colonies per clonal population were sequenced by Sanger plateau + 0.1(Dmax).
sequencing. Clones were selected based on 100% sequence Lytic HiBiT Detection and Viability of MV4;11 Cells. Clonal
conformity with no insertions or deletions in the HiBiT-containing cMyc-HiBiT edited cells were replicate plated at a density of 5 × 104
open reading frame. cells per well in solid, white 96-well tissue culture plates (Corning
HiBiT Blotting. HEK293 clonal cells (2 ×105) expressing HiBiT- Costar #3917), and cultured overnight. The following day, cells were
BET family members were plated in 24-well dishes overnight and then treated with either 10 nM or 100 nM JQ1, dBET1, or MZ1 for the
lysed in 100 μL of mammalian lysis buffer (Promega) supplemented indicated time frames. Following each treatment time point, one
with Protease Inhibitor Cocktail (Promega) and RQ1 RNase-Free replicate plate was assayed for luminescence of cMyc-HiBiT by
DNase (Promega) on ice. One hundred microliters of 2× SDS-PAGE addition of an equal volume of NanoGlo HiBiT Lytic Reagent
gel loading buffer (120 mM Tris-HCl, pH 6.8, 1.5 mM bromophenol (Promega N3030) to each well, and another replicate plate was
blue, 25% glycerol, 200 mM dithiothreitol, and 1% SDS) was added to assayed for cell viability by addition of an equal volume of CellTiter-
each sample and incubated for 5 min at 95 °C. Lysates were separated Glo 2.0 to each well (Promega). Plates were shaken on an orbital
by SDS-PAGE and transferred to a nitrocellulose membrane. shaker for 10−20 min before reading luminescence on a GloMax
Following transfer, the membrane was briefly incubated in TBST Discover (Promega).

2767 DOI: 10.1021/acschembio.8b00692


ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

NanoBRET Live Cell Ternary Complex and Ubiquitination. NanoGlo Substrate and Extracellular NanoLuc Inhibitor (Promega)
HEK293 clonal cells (8 ×105) expressing the HiBiT-BET family were added according to the manufacturer’s recommended protocol,
members were transfected with FuGENE HD (Promega) and 2 μg of and filtered luminescence was measured on a GloMax Discover
HaloTag-CRBN, HaloTag-VHL, or HaloTag-UBB in 6-well plates. luminometer equipped with 450 nm BP filter (donor) and 600 nm LP
For transient NanoBRET experiments with NanoLuc-BRD4, 8 × 105 filter (acceptor) using 0.3 s integration time. For real-time target
HEK293 cells were transfected with FuGENE HD (Promega) and engagement analysis, NanoBRET NanoGlo Substrate and Extrac-
0.02 μg of NanoLuc-BRD4 with either 2 μg of HaloTag-CRBN, or ellular NanoLuc Inhibitor (Promega) and 1× NanoBRET tracer were
HaloTag-VHL. The following day, 2 × 104 transfected cells were added to the cells 30 min prior test compound addition. To measure
replated into white 96-well tissue culture plates in the presence or NanoBRET in permeabilized cells, digitonin was added to the cells to
absence of HaloTag NanoBRET 618 Ligand (Promega) and a final concentration of 50 ug/mL and Extracellular Nluc inhibitor
incubated overnight at 37 °C, 5% CO2. The following day, medium was omitted during the detection step. Milli-BRET units (mBU) are
was replaced with Opti-MEM (Gibco) containing 20 μM Vivazine, an calculated by multiplying the raw NanoBRET values by 1000.
extended time-released substrate (Promega), and plates were
incubated at 37 °C, 5% CO2, for 1 h before addition of DMSO or
1 μM final concentration dBET1 or MZ1 compounds. Plates were
then read every 3 min for a period of 6 h on a CLARIOstar equipped

*
ASSOCIATED CONTENT
S Supporting Information

with an atmospheric control unit (BMG Labtech) set to 37 °C and


The Supporting Information is available free of charge on the
5% CO2. Dual filtered luminescence was collected with a 460/80 nm ACS Publications website at DOI: 10.1021/acschem-
bandpass filter (donor, NanoBiT-BET protein) and a 610 nm long bio.8b00692.
pass filter (acceptor, HaloTag NanoBRET ligand) using an Tables of CRISPR guide and template sequences and
integration time of 0.5 s. Background subtracted NanoBRET ratios comparative HiBiT-BET family expression; figures of
expressed in milliBRET units were calculated from the equation: kinetic competition of dBET1 and MZ1 with JQ1,
ji acceptor channel acceptor channel (no ligand) zyz
mBRET ratio = jjjj zz1000
impact of ectopic expression on degradation, BET family
k donor channel (no ligand) z{
− member time at Dmax at 10 and 100 nM PROTAC, BET
donor channel
family protein level profiles after JQ1 inhibition, ternary
Fold increase in BRET was calculated by normalizing mBRET ratios complex specificity and impact of proteasomal inhib-
to the average mBRET ratios for DMSO controls. ition, and NanoBRET target engagement of dBET1 with
NanoBRET Ubiquitination ImmunoAssay. Clonal cell lines NanoLuc-CRBN, (PDF)
edited for HiBiT insertion to BET family members were plated in
white 96-well tissue culture plates at a density of 2 × 104 cells per well
in 100 μL of growth medium and incubated overnight at 37 °C, 5%
CO2. The following day, 1 μM dBET1 or MZ1 was added to the plate
■ AUTHOR INFORMATION
Corresponding Author
for the indicated time frames before replacing medium with Opti- *Tel: +1-608-274-4330; Fax: +1-608-277-2601; E-mail:
MEM (Gibco) containing 200 μg/mL digitonin, 1:200 dilution of danette.daniels@promega.com.
primary anti-Ub antibody (Enzo Life Sciences, BML-PW8810), 1:500 ORCID
dilution of secondary antimouse Alexa 594 antibody (Cell Signaling Danette L. Daniels: 0000-0002-7659-3020
Technologies, 8890), and 20 μM NanoGlo substrate. Additional
control wells received no antibodies (control for background Notes
NanoBRET) or no primary antibody (control for specificity). Plates The authors declare the following competing financial
were placed on an orbital shaker for 10 min and NanoBRET interest(s): All authors are employees of Promega Corporation
measurements were collected on a CLARIOstar (BMG Labtech). and Promega Corporation is the commercial owner by
NanoBRET calculations were made according to the equation in the assignment of patents of the HaloTag, NanoLuc, NanoBRET
section: NanoBRET Live Cell Ternary Complex and Ubiquitination. target engagement, NanoBiT, and HiBiT technologies and
Fold increase in NanoBRET was calculated by normalizing mBRET their applications.


ratios to the untreated, t = 0 wells.
NanoBRET Target Engagement. NanoLuc-CRBN, VHL-Nano- ACKNOWLEDGMENTS
Luc, and NanoLuc-BRD4 were transfected in HEK-293 cells using
FuGENE HD (Promega) according to the manufacturer’s protocol. We wish to thank A. Ciulli for providing MZ1 and VH298
Briefly, NanoLuc/target fusion constructs were diluted into Trans- compounds, T. Machleidt and M. Schwinn for guidance on
fection Carrier DNA (Promega) at a mass ratio of 1:10 (mass/mass), HiBiT CRISPR/Cas9 editing and development of LgBiT
after which FuGENE HD was added at a ratio of 1:3 (μg DNA: μL stable cell line, M. Slater for advice on CRISPR clonal
FuGENE HD). One part (vol) of FuGENE HD complexes thus selection, C. Zimprich for help with target engagement, T.
formed were combined with 20 parts (vol) of HEK-293 cells Worzella for Glo-Max Discover instrumentation support, and
suspended at a density of 2 × 105, followed by incubation in a
G. Tarpley for thoughtful advice and continued support.


humidified, 37 °C/5% CO2 incubator for 20 h. Following transfection,
cells were washed and resuspended in Opti-MEM. NanoBRET assays
were performed in white, 96-well plates (Corning 3600) at a density REFERENCES
of 2 × 104 cells/well. All chemical inhibitors were prepared as (1) Deshaies, R. J. (2015) Protein degradation: Prime time for
concentrated stock solutions in DMSO (Sigma-Aldrich) and diluted PROTACs. Nat. Chem. Biol. 11, 634−635.
in Opti-MEM (unless otherwise noted) to prepare working stocks. (2) Lai, A. C., and Crews, C. M. (2017) Induced protein
For end point analysis of target engagement, cells were equilibrated degradation: an emerging drug discovery paradigm. Nat. Rev. Drug
for 2 h with energy transfer probes and test compound prior to Discovery 16, 101−114.
NanoBRET measurements. NanoBRET tracers were prepared at a (3) Churcher, I. (2018) Protac-Induced Protein Degradation in
working concentration of 20× in tracer dilution buffer (12.5 mM Drug Discovery: Breaking the Rules or Just Making New Ones? J.
HEPES, 31.25% PEG-400, pH 7.5). VHL NanoBRET Tracer was Med. Chem. 61, 444−452.
added to the cells at a final concentration of 1 μM, CRBN NanoBRET (4) Collins, I., Wang, H., Caldwell, J. J., and Chopra, R. (2017)
tracer was added to cells at a final concentration of 0.5 μM, and Chemical approaches to targeted protein degradation through
NanoBRET BRD Tracer-02 was added to cells at a final concentration modulation of the ubiquitin-proteasome pathway. Biochem. J. 474,
of 0.25 μM. To measure NanoBRET in live cells, NanoBRET 1127−1147.

2768 DOI: 10.1021/acschembio.8b00692


ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

(5) Sakamoto, K. M., Kim, K. B., Kumagai, A., Mercurio, F., Crews, Advantages of Targeted Protein Degradation Over Inhibition: An
C. M., and Deshaies, R. J. (2001) Protacs: chimeric molecules that RTK Case Study. Cell Chem. Biol. 25, 67−77.
target proteins to the Skp1-Cullin-F box complex for ubiquitination (24) Gechijian, L. N., Buckley, D. L., Lawlor, M. A., Reyes, J. M.,
and degradation. Proc. Natl. Acad. Sci. U. S. A. 98, 8554−8559. Paulk, J., Ott, C. J., Winter, G. E., Erb, M. A., Scott, T. G., Xu, M., Seo,
(6) Sakamoto, K. M., Kim, K. B., Verma, R., Ransick, A., Stein, B., H. S., Dhe-Paganon, S., Kwiatkowski, N. P., Perry, J. A., Qi, J., Gray,
Crews, C. M., and Deshaies, R. J. (2003) Development of Protacs to N. S., and Bradner, J. E. (2018) Functional TRIM24 degrader via
target cancer-promoting proteins for ubiquitination and degradation. conjugation of ineffectual bromodomain and VHL ligands. Nat. Chem.
Mol. Cell. Proteomics 2, 1350−1358. Biol. 14, 405−412.
(7) Schneekloth, J. S., Jr., Fonseca, F. N., Koldobskiy, M., Mandal, (25) Gustafson, J. L., Neklesa, T. K., Cox, C. S., Roth, A. G., Buckley,
A., Deshaies, R., Sakamoto, K., and Crews, C. M. (2004) Chemical D. L., Tae, H. S., Sundberg, T. B., Stagg, D. B., Hines, J., McDonnell,
genetic control of protein levels: selective in vivo targeted D. P., Norris, J. D., and Crews, C. M. (2015) Small-Molecule-
degradation. J. Am. Chem. Soc. 126, 3748−3754. Mediated Degradation of the Androgen Receptor through Hydro-
(8) Neklesa, T. K., and Crews, C. M. (2012) Chemical biology: phobic Tagging. Angew. Chem., Int. Ed. 54, 9659−9662.
Greasy tags for protein removal. Nature 487, 308−309. (26) Huang, H. T., Dobrovolsky, D., Paulk, J., Yang, G., Weisberg, E.
(9) Carmony, K. C., and Kim, K. B. (2012) PROTAC-induced L., Doctor, Z. M., Buckley, D. L., Cho, J. H., Ko, E., Jang, J., Shi, K.,
proteolytic targeting. Methods Mol. Biol. 832, 627−638. Choi, H. G., Griffin, J. D., Li, Y., Treon, S. P., Fischer, E. S., Bradner, J.
(10) Cromm, P. M., and Crews, C. M. (2017) Targeted Protein E., Tan, L., and Gray, N. S. (2018) A Chemoproteomic Approach to
Degradation: from Chemical Biology to Drug Discovery. Cell Chem. Query the Degradable Kinome Using a Multi-kinase Degrader. Cell
Biol. 24, 1181−1190. Chem. Biol. 25, 88−99.
(11) Gu, S., Cui, D., Chen, X., Xiong, X., and Zhao, Y. (2018) (27) Nabet, B., Roberts, J. M., Buckley, D. L., Paulk, J., Dastjerdi, S.,
PROTACs: An Emerging Targeting Technique for Protein Degrada- Yang, A., Leggett, A. L., Erb, M. A., Lawlor, M. A., Souza, A., Scott, T.
tion in Drug Discovery. BioEssays 40, 1700247. G., Vittori, S., Perry, J. A., Qi, J., Winter, G. E., Wong, K. K., Gray, N.
(12) Hughes, S. J., and Ciulli, A. (2017) Molecular recognition of S., and Bradner, J. E. (2018) The dTAG system for immediate and
ternary complexes: a new dimension in the structure-guided design of target-specific protein degradation. Nat. Chem. Biol. 14, 431−441.
chemical degraders. Essays Biochem. 61, 505−516. (28) Olson, C. M., Jiang, B., Erb, M. A., Liang, Y., Doctor, Z. M.,
(13) Ottis, P., and Crews, C. M. (2017) Proteolysis-Targeting Zhang, Z., Zhang, T., Kwiatkowski, N., Boukhali, M., Green, J. L.,
Chimeras: Induced Protein Degradation as a Therapeutic Strategy. Haas, W., Nomanbhoy, T., Fischer, E. S., Young, R. A., Bradner, J. E.,
ACS Chem. Biol. 12, 892−898. Winter, G. E., and Gray, N. S. (2017) Pharmacological perturbation of
(14) Winter, G. E., Buckley, D. L., Paulk, J., Roberts, J. M., Souza, A., CDK9 using selective CDK9 inhibition or degradation. Nat. Chem.
Dhe-Paganon, S., and Bradner, J. E. (2015) DRUG DEVELOP-
Biol. 14, 163−170.
MENT. Phthalimide conjugation as a strategy for in vivo target (29) Schiedel, M., Herp, D., Hammelmann, S., Swyter, S., Lehotzky,
protein degradation. Science 348, 1376−1381.
A., Robaa, D., Olah, J., Ovadi, J., Sippl, W., and Jung, M. (2018)
(15) Neklesa, T. K., Winkler, J. D., and Crews, C. M. (2017)
Chemically Induced Degradation of Sirtuin 2 (Sirt2) by a Proteolysis
Targeted protein degradation by PROTACs. Pharmacol. Ther. 174,
Targeting Chimera (PROTAC) Based on Sirtuin Rearranging Ligands
138−144.
(SirReals). J. Med. Chem. 61, 482−491.
(16) Lohbeck, J., and Miller, A. K. (2016) Practical synthesis of a
(30) Zengerle, M., Chan, K. H., and Ciulli, A. (2015) Selective Small
phthalimide-based Cereblon ligand to enable PROTAC development.
Molecule Induced Degradation of the BET Bromodomain Protein
Bioorg. Med. Chem. Lett. 26, 5260−5262.
(17) Lu, J., Qian, Y., Altieri, M., Dong, H., Wang, J., Raina, K., Hines, BRD4. ACS Chem. Biol. 10, 1770−1777.
(31) Zhang, C., Han, X. R., Yang, X., Jiang, B., Liu, J., Xiong, Y., and
J., Winkler, J. D., Crew, A. P., Coleman, K., and Crews, C. M. (2015)
Hijacking the E3 Ubiquitin Ligase Cereblon to Efficiently Target Jin, J. (2018) Proteolysis Targeting Chimeras (PROTACs) of
BRD4. Chem. Biol. 22, 755−763. Anaplastic Lymphoma Kinase (ALK). Eur. J. Med. Chem. 151, 304−
(18) Lu, M., Liu, T., Jiao, Q., Ji, J., Tao, M., Liu, Y., You, Q., and 314.
Jiang, Z. (2018) Discovery of a Keap1-dependent peptide PROTAC (32) DeMars, K. M., Yang, C., Castro-Rivera, C. I., and Candelario-
to knockdown Tau by ubiquitination-proteasome degradation path- Jalil, E. (2018) Selective degradation of BET proteins with dBET1, a
way. Eur. J. Med. Chem. 146, 251−259. proteolysis-targeting chimera, potently reduces pro-inflammatory
(19) Ottis, P., Toure, M., Cromm, P. M., Ko, E., Gustafson, J. L., and responses in lipopolysaccharide-activated microglia. Biochem. Biophys.
Crews, C. M. (2017) Assessing Different E3 Ligases for Small Res. Commun. 497, 410−415.
Molecule Induced Protein Ubiquitination and Degradation. ACS (33) Kerres, N., Steurer, S., Schlager, S., Bader, G., Berger, H.,
Chem. Biol. 12, 2570−2578. Caligiuri, M., Dank, C., Engen, J. R., Ettmayer, P., Fischerauer, B.,
(20) Bulatov, E., and Ciulli, A. (2015) Targeting Cullin-RING E3 Flotzinger, G., Gerlach, D., Gerstberger, T., Gmaschitz, T., Greb, P.,
ubiquitin ligases for drug discovery: structure, assembly and small- Han, B., Heyes, E., Iacob, R. E., Kessler, D., Kolle, H., Lamarre, L.,
molecule modulation. Biochem. J. 467, 365−386. Lancia, D. R., Lucas, S., Mayer, M., Mayr, K., Mischerikow, N., Muck,
(21) Bondeson, D. P., Mares, A., Smith, I. E., Ko, E., Campos, S., K., Peinsipp, C., Petermann, O., Reiser, U., Rudolph, D., Rumpel, K.,
Miah, A. H., Mulholland, K. E., Routly, N., Buckley, D. L., Gustafson, Salomon, C., Scharn, D., Schnitzer, R., Schrenk, A., Schweifer, N.,
J. L., Zinn, N., Grandi, P., Shimamura, S., Bergamini, G., Faelth- Thompson, D., Traxler, E., Varecka, R., Voss, T., Weiss-Puxbaum, A.,
Savitski, M., Bantscheff, M., Cox, C., Gordon, D. A., Willard, R. R., Winkler, S., Zheng, X., Zoephel, A., Kraut, N., McConnell, D.,
Flanagan, J. J., Casillas, L. N., Votta, B. J., den Besten, W., Famm, K., Pearson, M., and Koegl, M. (2017) Chemically Induced Degradation
Kruidenier, L., Carter, P. S., Harling, J. D., Churcher, I., and Crews, C. of the Oncogenic Transcription Factor BCL6. Cell Rep. 20, 2860−
M. (2015) Catalytic in vivo protein knockdown by small-molecule 2875.
PROTACs. Nat. Chem. Biol. 11, 611−617. (34) Raina, K., Lu, J., Qian, Y., Altieri, M., Gordon, D., Rossi, A. M.,
(22) Buckley, D. L., Raina, K., Darricarrere, N., Hines, J., Gustafson, Wang, J., Chen, X., Dong, H., Siu, K., Winkler, J. D., Crew, A. P.,
J. L., Smith, I. E., Miah, A. H., Harling, J. D., and Crews, C. M. (2015) Crews, C. M., and Coleman, K. G. (2016) PROTAC-induced BET
HaloPROTACS: Use of Small Molecule PROTACs to Induce protein degradation as a therapy for castration-resistant prostate
Degradation of HaloTag Fusion Proteins. ACS Chem. Biol. 10, cancer. Proc. Natl. Acad. Sci. U. S. A. 113, 7124−7129.
1831−1837. (35) Sun, B., Fiskus, W., Qian, Y., Rajapakshe, K., Raina, K.,
(23) Burslem, G. M., Smith, B. E., Lai, A. C., Jaime-Figueroa, S., Coleman, K. G., Crew, A. P., Shen, A., Saenz, D. T., Mill, C. P.,
McQuaid, D. C., Bondeson, D. P., Toure, M., Dong, H., Qian, Y., Nowak, A. J., Jain, N., Zhang, L., Wang, M., Khoury, J. D., Coarfa, C.,
Wang, J., Crew, A. P., Hines, J., and Crews, C. M. (2018) The Crews, C. M., and Bhalla, K. N. (2018) BET protein proteolysis

2769 DOI: 10.1021/acschembio.8b00692


ACS Chem. Biol. 2018, 13, 2758−2770
ACS Chemical Biology Articles

targeting chimera (PROTAC) exerts potent lethal activity against and selective chemical probe of hypoxic signalling downstream of
mantle cell lymphoma cells. Leukemia 32, 343−352. HIF-alpha hydroxylation via VHL inhibition. Nat. Commun. 7, 13312.
(36) Bondeson, D. P., Smith, B. E., Burslem, G. M., Buhimschi, A. (49) Savitski, M. M., Zinn, N., Faelth-Savitski, M., Poeckel, D., Gade,
D., Hines, J., Jaime-Figueroa, S., Wang, J., Hamman, B. D., Ishchenko, S., Becher, I., Muelbaier, M., Wagner, A. J., Strohmer, K., Werner, T.,
A., and Crews, C. M. (2018) Lessons in PROTAC Design from Melchert, S., Petretich, M., Rutkowska, A., Vappiani, J., Franken, H.,
Selective Degradation with a Promiscuous Warhead. Cell Chem. Biol. Steidel, M., Sweetman, G. M., Gilan, O., Lam, E. Y. N., Dawson, M.
25, 78−87. A., Prinjha, R. K., Grandi, P., Bergamini, G., and Bantscheff, M.
(37) Gadd, M. S., Testa, A., Lucas, X., Chan, K. H., Chen, W., (2018) Multiplexed Proteome Dynamics Profiling Reveals Mecha-
Lamont, D. J., Zengerle, M., and Ciulli, A. (2017) Structural basis of nisms Controlling Protein Homeostasis. Cell 173, 260−274.
PROTAC cooperative recognition for selective protein degradation. (50) Winter, G. E., Mayer, A., Buckley, D. L., Erb, M. A., Roderick, J.
Nat. Chem. Biol. 13, 514−521. E., Vittori, S., Reyes, J. M., di Iulio, J., Souza, A., Ott, C. J., Roberts, J.
(38) Cong, L., Ran, F. A., Cox, D., Lin, S., Barretto, R., Habib, N., M., Zeid, R., Scott, T. G., Paulk, J., Lachance, K., Olson, C. M.,
Hsu, P. D., Wu, X., Jiang, W., Marraffini, L. A., and Zhang, F. (2013) Dastjerdi, S., Bauer, S., Lin, C. Y., Gray, N. S., Kelliher, M. A.,
Multiplex genome engineering using CRISPR/Cas systems. Science Churchman, L. S., and Bradner, J. E. (2017) BET Bromodomain
339, 819−823. Proteins Function as Master Transcription Elongation Factors
(39) Jinek, M., Chylinski, K., Fonfara, I., Hauer, M., Doudna, J. A., Independent of CDK9 Recruitment. Mol. Cell 67, 5−18.
and Charpentier, E. (2012) A programmable dual-RNA-guided DNA
endonuclease in adaptive bacterial immunity. Science 337, 816−821.
(40) Hall, M. P., Unch, J., Binkowski, B. F., Valley, M. P., Butler, B.
L., Wood, M. G., Otto, P., Zimmerman, K., Vidugiris, G., Machleidt,
T., Robers, M. B., Benink, H. A., Eggers, C. T., Slater, M. R.,
Meisenheimer, P. L., Klaubert, D. H., Fan, F., Encell, L. P., and Wood,
K. V. (2012) Engineered luciferase reporter from a deep sea shrimp
utilizing a novel imidazopyrazinone substrate. ACS Chem. Biol. 7,
1848−1857.
(41) Schwinn, M. K., Machleidt, T., Zimmerman, K., Eggers, C. T.,
Dixon, A. S., Hurst, R., Hall, M. P., Encell, L. P., Binkowski, B. F., and
Wood, K. V. (2018) CRISPR-Mediated Tagging of Endogenous
Proteins with a Luminescent Peptide. ACS Chem. Biol. 13, 467−474.
(42) Machleidt, T., Woodroofe, C. C., Schwinn, M. K., Mendez, J.,
Robers, M. B., Zimmerman, K., Otto, P., Daniels, D. L., Kirkland, T.
A., and Wood, K. V. (2015) NanoBRET−A Novel BRET Platform for
the Analysis of Protein-Protein Interactions. ACS Chem. Biol. 10,
1797−1804.
(43) Robers, M. B., Dart, M. L., Woodroofe, C. C., Zimprich, C. A.,
Kirkland, T. A., Machleidt, T., Kupcho, K. R., Levin, S., Hartnett, J. R.,
Zimmerman, K., Niles, A. L., Ohana, R. F., Daniels, D. L., Slater, M.,
Wood, M. G., Cong, M., Cheng, Y. Q., and Wood, K. V. (2015)
Target engagement and drug residence time can be observed in living
cells with BRET. Nat. Commun. 6, 10091.
(44) Filippakopoulos, P., Qi, J., Picaud, S., Shen, Y., Smith, W. B.,
Fedorov, O., Morse, E. M., Keates, T., Hickman, T. T., Felletar, I.,
Philpott, M., Munro, S., McKeown, M. R., Wang, Y., Christie, A. L.,
West, N., Cameron, M. J., Schwartz, B., Heightman, T. D., La
Thangue, N., French, C. A., Wiest, O., Kung, A. L., Knapp, S., and
Bradner, J. E. (2010) Selective inhibition of BET bromodomains.
Nature 468, 1067−1073.
(45) Dawson, M. A., Prinjha, R. K., Dittmann, A., Giotopoulos, G.,
Bantscheff, M., Chan, W. I., Robson, S. C., Chung, C. W., Hopf, C.,
Savitski, M. M., Huthmacher, C., Gudgin, E., Lugo, D., Beinke, S.,
Chapman, T. D., Roberts, E. J., Soden, P. E., Auger, K. R., Mirguet,
O., Doehner, K., Delwel, R., Burnett, A. K., Jeffrey, P., Drewes, G.,
Lee, K., Huntly, B. J., and Kouzarides, T. (2011) Inhibition of BET
recruitment to chromatin as an effective treatment for MLL-fusion
leukaemia. Nature 478, 529−533.
(46) Wurz, R. P., Dellamaggiore, K., Dou, H., Javier, N., Lo, M. C.,
McCarter, J. D., Mohl, D., Sastri, C., Lipford, J. R., and Cee, V. J.
(2018) A “Click Chemistry Platform” for the Rapid Synthesis of
Bispecific Molecules for Inducing Protein Degradation. J. Med. Chem.
61, 453−461.
(47) Nowak, R. P., DeAngelo, S. L., Buckley, D., He, Z., Donovan, K.
A., An, J., Safaee, N., Jedrychowski, M. P., Ponthier, C. M., Ishoey, M.,
Zhang, T., Mancias, J. D., Gray, N. S., Bradner, J. E., and Fischer, E. S.
(2018) Plasticity in binding confers selectivity in ligand-induced
protein degradation. Nat. Chem. Biol. 14, 706.
(48) Frost, J., Galdeano, C., Soares, P., Gadd, M. S., Grzes, K. M.,
Ellis, L., Epemolu, O., Shimamura, S., Bantscheff, M., Grandi, P.,
Read, K. D., Cantrell, D. A., Rocha, S., and Ciulli, A. (2016) Potent

2770 DOI: 10.1021/acschembio.8b00692


ACS Chem. Biol. 2018, 13, 2758−2770

You might also like