You are on page 1of 7

pubs.acs.

org/biochemistry Article

Nanomolar, Noncovalent Antagonism of Hedgehog Cholesterolysis:


Exception to the “Irreversibility Rule” for Protein Autoprocessing
Inhibition
Andrew G. Wagner, Robert T. Stagnitta, Zihan Xu, John L. Pezzullo, Nabin Kandel, José-Luis Giner,
Douglas F. Covey, Chunyu Wang, and Brian P. Callahan*
Cite This: https://doi.org/10.1021/acs.biochem.1c00697 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
Downloaded via UNIV ESTADUAL PAULISTA on January 22, 2022 at 02:45:25 (UTC).

ABSTRACT: Hedgehog (Hh) signaling ligands undergo carboxy


terminal sterylation through specialized autoprocessing, called
cholesterolysis. Sterylation is brought about intramolecularly in a
single turnover by an adjacent enzymatic domain, called HhC,
which is found in precursor Hh proteins only. Previous attempts to
identify antagonists of the intramolecular activity of HhC have
yielded inhibitors that bind HhC irreversibly through covalent
mechanisms, as is common for protein autoprocessing inhibitors.
Here, we report an exception to the “irreversibility rule” for
autoprocessing inhibition. Using a fluorescence resonance energy transfer-based activity assay for HhC, we screened a focused library
of sterol-like analogues for noncovalent inhibitors and identified and validated four structurally related molecules, which were then
used for structure−activity relationship studies. The most effective derivative, tBT-HBT, inhibits HhC noncovalently with an IC50 of
300 nM. An allosteric binding site for tBT-HBT, encompassing residues from the two subdomains of HhC, is suggested by kinetic
analysis, mutagenesis studies, and photoaffinity labeling. The inhibitors described here resemble a family of noncovalent, allosteric
inducers of HhC paracatalysis which we have described previously. The inhibition and the induction appear to be mediated by a
shared allosteric site on HhC.

■ INTRODUCTION
Hedgehog (Hh) ligands, the extracellular signaling factors
two subdomains, the Hedgehog intein (HINT) and the sterol
recognition region (SRR), only the HINT has its three-
involved in embryo development and many types of dimensional structure resolved by experimental structural
cancers,1−8 appear unique in their post-translational mod- methods.22−24
ification by cholesterol.9,10 Lipidation occurs during Hh HhC and the protein substrate, HhN, are covalently linked
biosynthesis through an autoprocessing reaction called in the Hh precursor protein, and this presents a special
cholesterolysis. A partially characterized ∼25 kDa C-terminal challenge for antagonism. There are substantial chelate-type
domain, HhC, present in the precursor forms of Hh serves as connectivity effects between the HhC catalyst and HhN
the intramolecular catalyst for cholesterolysis (Figure 1, lef t).
substrate that must be overcome.25−28 Additional challenges
HhN-chol, the cholesterylated product, is released from HhC
and undergoes additional lipidation as it matures into the Hh arise from the limited turnover inherent to protein
ligand. 10,11 Deactivating mutations in HhC result in autoprocessing, typically a single turnover. Suppressing a single
endoplasmic reticulum (ER) retention and ER-associated turnover by an intramolecular catalyst demands inhibitors that
protein degradation of the Hh precursor,12,13 thereby are bound tightly; dissociation permits autoprocessing to
suppressing downstream Hh signaling.12,14 begin, potentially diminishing the inhibitor recognition.
We have begun pursuing selective antagonists of HhC that Accordingly, small-molecule antagonists of Hh cholesterolysis
mimic the effects of deactivating HhC mutations.15−17 We and related forms of protein autoprocessing nearly all depend
view HhC inhibitors as novel directions for therapeutic
intervention in human cancers, particularly for malignancies
driven by the Hh ligand and where existing Hh blocking drugs Received: October 20, 2021
have lost efficacy.7,18,19 Inhibitors of HhC cholesterolysis hold Revised: November 29, 2021
additional promise as chemical tools to understand the basic
biology of Hh signaling. HhC inhibitors may also stabilize
HhC’s dynamic structure,20 enabling the structural character-
ization of this unusual autoprocessing element.21 Of HhC’s

© XXXX American Chemical Society https://doi.org/10.1021/acs.biochem.1c00697


A Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

Figure 1. HhC functions as a dedicated self-cleaving, protein carboxy sterol ligase of the Hh ligand. (Left) Hh autoprocessing has two steps: (Step
1) the side-chain thiol of a conserved cysteine residue at position 1 of HhC attacks the amide bond of the preceding residue, glycine, to form an
internal thioester and (Step 2) binding of cholesterol to HhC triggers the transesterification to the lipid’s 3β-OH group, generating a
cholesterylated Hh ligand and displacing HhC. The two subdomains of HhC are indicated as follows: HINT (Hedgehog/Intein) and SRR. (Right)
Representative structures of HhC inhibitors reported here and of HhC paracatalytic inducers described previously.

Figure 2. Library screening identifies structurally related noncovalent HhC inhibitors. (A) Tripartite HhC reporter construct, C−H−Y, where HhC
is flanked by CFP and YFP. HhC cholesterolysis results in a loss of the FRET signal as H−Y and cholesterylated C separate. (B) Preliminary
inhibitory activity with compound 1. Traces show screen data with C−H−Y in the presence of 1.5 μM cholesterol and 50 μM 1 (red), along with
controls: C−H−Y in buffer only (black, solid) and C−H−Y with cholesterol (black dotted). (C) Chemical structures and IC50 values of screen hit
compounds 1−4. (D) Dose−response curves of inhibitor compounds 1−4 (green, blue, orange, and red, respectively) against C−H−Y in the presence
of 1.5 μM cholesterol.

on binding the precursor protein irreversibly through covalent


interactions.16,17,29−32
■ RESULTS AND DISCUSSION
Library Screening Identifies a Family of HhC
Here, we describe an exception to the empirical “irrever- Inhibitors. We tested 1187 steroid analogues from Chem-
sibility rule” for protein autoprocessing inhibition. Screening a Bridge as potential inhibitors of Hh cholesterolysis using our
focused library of sterol-like analogues, we identified the first continuous light-to-dark fluorescence resonance energy trans-
noncovalent, tight-binding inhibitors of HhC cholesterolysis. fer (FRET) activity assay.15−17 The FRET reporter construct,
The most effective inhibitor has an IC50 value of 300 × 10−9 C−H−Y, has the Drosophila melanogaster HhC flanked by cyan
M. A mixed inhibition mode is suggested by kinetic analysis fluorescent protein (CFP) (C) and yellow fluorescent protein
and by photoaffinity labeling (PAL). Computational modeling (YFP) (Y) as the FRET donor and acceptor, respectively.
and mutagenesis of HhC suggest that inhibitors bind an FRET is measured as the ratio of emissions of 540 nm/460 nm
allosteric site composed of residues from HhC’s two after excitation at 400 nm. Cholesterolysis results in the loss of
subdomains. We note a striking resemblance between the FRET as C−H−Y autoprocesses into HhC-YFP (H−Y) and
inhibitors of HhC cholesterolysis described here and a family sterylated CFP (C-chol) (Figure 2A).
of noncovalent, inducers of HhC paracatalysis, which we have Standard reaction conditions for compound screening using
described previously (Figure 1, right).15 The inhibitor/ C−H−Y in 96-well plates have been described.15 Briefly, wells
activator duality appears to be mediated by the same allosteric contained Ni-NTA purified C−H−Y (0.1 μM) in cholester-
site in HhC. olysis buffer (Tris, 20 mM, pH 7.1; Fos-Choline 12, 1.5 mM;
B https://doi.org/10.1021/acs.biochem.1c00697
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

Figure 3. Inhibition kinetics and photoactive sterol labeling suggest that the substrate and inhibitor can be bound by HhC simultaneously. (A)
Mixed-type inhibition kinetics apparent in the presence of inhibitor 1. Initial velocity of C−H−Y cholesterolysis with increasing concentrations of
compound 1 shows progressively lower maximum rate values. (B) Cravatt’s tSP, used here as a photoactivatable substrate analogue of HhC. (C)
Photolabeling of HhC by tSP persists with saturating concentrations of 1, consistent with noncompetitive interactions. All samples contained 5 μM
HhC and 50 μM azide-linked fluorophore; other components were varied as indicated. (Top) Photolabeling was assessed by SDS-PAGE and
fluorescent gel imaging. (Bottom) Coomassie stain for total HhC protein. Controls: Lanes 1−5, 8, and 9. Experimental: Lane 6, photolabeled HhC
by tSP in the absence of 1. Lane 7, diminished photolabeling of HhC by tSP with saturating cholesterol. Lane 10, persistent tSP PAL in the presence
of saturating compound 1.

ethylenediaminetetraacetic acid (EDTA), 5 mM; and dithio- as a function of inhibitor concentration (Figure 2D) were used
threitol (DTT), 1 mM) in a total volume of 100 μl. Library to calculate the IC50 values (Supporting Information). IC50
compounds were added from 10 mM dimethyl sulfoxide values for 1−4, obtained from ≥3 independent experiments,
(DMSO) stocks to a final concentration of 50 μM (0.5% are as follows: 1, 2.0 μM; 2, 4.0 μM; 3, 12 μM; and 4, 81 μM.
DMSO v/v). Prior to initiating the reaction with cholesterol, Kinetic Evidence of Mixed-Type Inhibition: Com-
plates were incubated for 30 min at 30 °C while FRET pound 1 Diminishes Catalysis through Effects on Both
readings were recorded every 2 min. Data collected during this kmax and KM. We assessed the changes in the kinetic
delay were monitored to flag compounds that interfered with parameters, kmax and KM, for cholesterolysis using C−H−Y
the C−H−Y signal. Cholesterolysis was initiated by adding in the presence of 1 to understand the mode of inhibition. The
cholesterol from an ethanol stock (1.5 μM final, 4% ethanol v/ above dose−response experiments indicated an IC50 value of
v), and FRET was recorded every 2 min for 1.5 h. All plates 2.0 μM. In the kinetic experiments described here, 1 was added
contained 16 controls wells, 8 with C−H−Y plus cholesterol, to C−H−Y at concentrations of 0.8, 2.5, 5.0, 10, and 25 μM,
representing 100% activity, and 8 wells with C−H−Y minus corresponding to 0.4−12.5 × the IC50 value. To those wells,
cholesterol, representing 100% inhibition. A total of five we titrated substrate cholesterol from 0.2 to 100 μM. In Figure
compounds from the library displayed inhibitory activity 3A, the initial velocity is plotted as a function of increasing
toward C−H−Y. In Figure 2B, we show kinetic traces for cholesterol concentration. Initial velocity data were fit to a
C−H−Y in the presence of 1, the most effective inhibitor from standard Michaelis−Menten equation (Supporting Informa-
the screen along with intraplate controls. The remaining hit tion). Analysis of the changing kinetic values ±1 suggested
molecules possessed structural features similar to that of mixed-type inhibition, where both KM and kmax values are
compound 1: N-tert butyl pyrrole, aliphatic cyclic amine, and depressed by the inhibitor. Calculated KM values under the
an aryl group (Figure 2C). conditions tested were as follows: no inhibitor, KM 1.2 μM;
Four of the five preliminary hits 1−4 were available for with 0.8 μM 1, KM 1.1 μM; with 2.5 μM 1, KM 1.6 μM; with 10
repurchase and were retested by dose−response inhibition μM 1, KM 2.6 μM; and with 25 μM 1, KM 4.0 μM. The
experiments using C−H−Y. The compounds were also calculated KM values appear modestly sensitive to increasing
evaluated by end-point kinetic analysis using sodium dodecyl concentrations of 1, with an increase in KM of 3-fold following
sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) (Fig- a 25-fold increase in the concentration of 1. There was a
ure S1). For the dose−response experiments, compounds were clearer concentration-dependent suppression of kmax with the
dissolved in DMSO to 10 mM and then diluted from 100 to addition of 1. In the absence of an inhibitor, kmax for
0.015 μM into wells containing C−H−Y (0.1 μM) over a 9- cholesterolysis is 1.9 × 10−3 s−1. With 0.8 μM 1, kmax is reduced
point titration. To initiate the assay, cholesterol was added at to 1.3 × 10−3 s−1, and at 2.5 μM 1, kmax is reduced further to
its KM value of 1.5 μM,33 and FRET readings were collected 0.9 × 10−3 s−1. Increasing the concentration of 1 to 10 and 25
every 2 min. Initial velocity values were calculated as the slope μM decreases kmax to 0.48 × 10−3 s−1 and 0.29 × 10−3 s−1,
of the linear phase of the FRET decay at each inhibitor respectively. In summary, inhibitor 1 interferes mainly with the
concentration. Relative velocity (RV) was calculated by turnover rate along with a lesser but reproducible increase in
dividing the initial velocity with the inhibitor by the initial the KM for substrate cholesterol. These results are consistent
velocity in the absence of the inhibitor. Plots of the RV values with mixed-type inhibition.
C https://doi.org/10.1021/acs.biochem.1c00697
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

Figure 4. Nanomolar HhC inhibitor and provisional binding site by computational modeling and point mutagenesis. (A) Chemical structures of
HhC inhibitor analogues, along with IC50 values toward cholesterolysis. (B) Proposed ternary complex of the substrate, the inhibitor, and HhC.
Color key: inhibitor 8 or tBT-HBT, yellow; cholesterol, magenta; HINT subdomain, cyan; and SRR, purple. Point mutations and corresponding KD
values are shown along with each mutant’s cholesterolysis activity (active/inactive). We note that mutations have been made at several homologous
positions by Ondrus and her associates (refs 20 and 21) in human Sonic HhC (C1, H72, L73, L128, C143, Y144, H154, R161, H193, W194, and
Y195), and the resulting effects on cholesterolysis activity accord well with the behavior of the mutants evaluated here. Mutation color key: red,
complete loss of inhibitor binding; orange, 60−300 fold loss of inhibitor binding; and gray, <10-fold loss of inhibitor binding affinity.

Photolabeling of HhC by a Diazirine-Modified site maps to the SRR of HhC. In summary, tSP furnishes an
Substrate Analogue Persists with Saturating Concen- active-site PAL probe for HhC.
trations of 1, Suggesting a Ternary HhC−Substrate− Confident in the utility of tSP, we carried out an
Inhibitor Complex. Mixed-type inhibitors can be bound by experimental test for HhC-tSP-1 ternary complex formation.
enzymatic catalysts both in the presence and in the absence of We sought to determine whether the addition of 1 would
the substrate. To further probe for ternary “E-S-I” complex diminish the active-site photolabeling of HhC by tSP. We
formation, suggested by the above kinetic analysis, we applied mixed HhC (5 μM) and tSP (10 μM) ± saturating inhibitor 1
PAL with the trans-sterol probe (tSP)34 as an active-site (200 μM), followed by UV irradiation to activate PAL and
modifier of HhC (Figure 3B). We reasoned that if inhibitor 1 then click chemistry labeling to attach an azide-modified
and tSP were bound by HhC at separate sites, PAL of HhC by fluorophore (50 μM) to the alkyne group on tSP. A covalent
tSP would likely be insensitive to added 1; if, however, HhC/tSP/fluorophore complex was detected by SDS-PAGE
inhibitor 1 and tSP bind HhC competitively, the addition of with UV gel imaging. In Figure 3C, the results of the control
saturating 1 should diminish the PAL of HhC. tSP contains a reactions with tSP, HhC, and the azide-modified fluorophore
diazirine group at C-6 of the sterol nucleus. Alkyl diazirine are shown first in lanes 1−7. Out of this set, a fluorescence
groups decompose under UV light to produce N2 and a signal is apparent in lane 6 only. The sample in lane 6 was
reactive diazo intermediate or, alternatively, a carbene.35 subjected to UV irradiation and click chemistry. The covalent
Reaction with nearby protein residues generates a covalent HhC/tSP/fluorophore species detected in lane 6 is absent in
protein/probe adduct.20,36,37 The sterol side chain in tSP lane 3. The lane 3 sample, which lacked UV irradiation, shows
includes an alkyne group to facilitate the analysis of PAL that HhC in complex with tSP dissociates without PAL. The
through the use of copper click chemistry.34,38 control sample in lane 7 is identical to the lane 6 sample except
We established the utility of tSP as a cholesterol analogue for that it also included cholesterol (200 μM). The absence of the
HhC in two ways. First, we tested tSP as an alternative HhC HhC/tSP/fluorophore complex in lane 7 demonstrates that
substrate using the C−H−Y system. The results demonstrated active site binding by cholesterol competitively blocks the PAL
that tSP was readily accepted as a substrate, with a KM value of by the substrate analogue, tSP. In the second sample set, Figure
6.8 μM and a kmax value of 4.2 × 10−3 s−1 (Figure S2A,B), 3C lane 8−10, we examined the influence of inhibitor 1 on
differing by < 2-fold in kmax/KM compared with cholesterol. HhC/tSP photolabeling. Samples in lane 8−9 are controls to
Next, we mapped the tSP binding site on HhC using tandem test for nonspecific fluorophore labeling promoted by added 1;
mass spectrometry (MS/MS) analysis of an HhC/tSP adduct. no evidence was observed. The key experimental sample is in
Earlier truncation studies of HhC indicated that the SRR lane 10, containing HhC, tSP, plus saturating 1, and an azide-
resides in the final 70−90 amino acids of HhC.22 To simplify modified fluorophore. Despite saturating 1, this inhibitor of
the mapping study, we used a product form of HhC, post- HhC cholesterolysis did not significantly diminish active site
autoprocessing. Details on the preparation of product HhC photolabeling by tSP. The persistence of photolabeling of HhC
and photolabeling to form the HhC/tSP adduct are provided by substrate analogue tSP in the presence of inhibitor 1 is
in the Supporting Information (Figure S2C). Preliminary MS/ consistent with the mode-of-inhibition analysis above:
MS analysis of purified HhC/tSP indicated the tSP attachment cholesterolysis antagonism involves a catalytically inert ternary
to HhC peptide EQLHSSPK (HhC residues 174−181). This complex of HhC, sterol, and 1.
D https://doi.org/10.1021/acs.biochem.1c00697
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

Analogue Testing Identifies tBT-HBT as the First 4B, gray); between 60- and 300-fold loss of the binding affinity
Nanomolar Inhibitor of HhC Cholesterolysis. To explore (Figure 4B, orange); and abolished inhibitor binding (Figure
the structure−activity relationship (SAR) with this novel class 4B, red). The most severe loss in tBT-HBT binding mapped to
of HhC inhibitors, we prepared and tested several analogues of mutations in the HINT domain (L73, F86, C143, and Y144,
1−4 (Figure 4A). The first analogue, 5, incorporates the colored red) lining the predicted tBT-HBT binding site in our
benzothiazole and N-t-butyl pyrrole groups of 1 and the computational model. Mutations that produced a large but not
homopiperazine of 2. As an inhibitor of cholesterolysis, the total loss of tBT-HBT binding mapped to the SRR subdomain
molecule was improved 2−3 fold over 1 and 2, with an IC50 of (H154A, R161A, H193Y, and W194A, colored orange). Two
1.2 μM toward C−H−Y in the FRET assay. In analogue 6, we mutations in HINT (C1A and H72A, colored gray) and two in
replaced the t-butyl group of 5 with a hydrogen atom; this the SRR (G191D and Y195F, colored gray) resulted in smaller
truncation resulted in a substantially weakened inhibitor, with losses of the tBT-HBT affinity. The compromised binding in
an IC50 of 61 μM. The t-butyl group was restored in analogues H154A, R161A, and H193A, which are distal to the predicted
7 and 8, while the heterocycle was switched from the pyrrole to tBT-HBT binding site, lack clear explanation, indicating that
1,3 oxazole 7 or thiophene 8. Compound 7 was indistinguish- the computational model needs further improvement.
able from 5 as a cholesterolysis inhibitor. Thiophene 8 Collectively, the results support the idea that tBT-HBT
provided a ∼ 5-fold increase in binding compared with binding by HhC involves SRR and HINT subdomain
analogue 6. Analogue 8 is thereby the first submicromolar interactions.
noncovalent inhibitor of HhC cholesterolysis, with an IC50 of
300 nM. Four additional analogues, 9−12, based on 8 were
prepared next. The t-butyl group was truncated in 9 or
■ CONCLUDING REMARKS AND
INHIBITOR/ACTIVATOR DUALITY
substituted in 10−12. Each change diminished the inhibitory HhC, the C-terminal half of Hedgehog precursor proteins,
activity to varying degrees, reinforcing the importance of the t- possesses intramolecular cholesterolysis activity that is
butyl group for binding by HhC. Analogue 8, hereafter called necessary to release and sterylate the adjacent Hh signaling
tBT-HBT, emerged from this study as the best inhibitor, ligand, HhN. The significance of this initiating step in the Hh
improving the IC50 by ∼10-fold from the top hit of the library signaling pathway is three-fold: the transformation appears
screen. unique to Hh family proteins; it lies upstream of all known
Inhibitor Binding by HhC Requires Residues from the receptor interactions; and lastly, sterylation provides a vital
SRR and HINT Subdomains. We used the nM inhibitor tBT- membrane anchor that regulates Hh ligand diffusion and
HBT (8) along with a panel of HhC mutants to probe for the physiological signaling. Small molecules that manipulate HhC
inhibitor binding site. To guide the mutagenesis, a computa- autoprocessing could help unravel the complex biology of Hh
tional model of the ternary complex, HhC−substrate− signaling by providing an off switch near the pathway’s origin.
inhibitor, was prepared with D. melanogaster using the program The promise of HhC antagonists/agonists extends to
MODELLER (Supporting Information).39 Results of virtual therapeutics in treating sporadic tumors and congenital
docking40 of substrate cholesterol and tBT-HBT into the disorders where Hh ligand signaling is dysregulated.
predicted D. melanogaster HhC structure are shown in Figure Because of their unique action, protein autoprocessing
4B. Single-point mutations in HhC were introduced based in domains represent intrinsically hard-to-hit targets for non-
part on this predicted ternary complex as well as on residue covalent antagonism. There is a large entropic advantage
conservation across HhC sequences. associated with intramolecular catalysis that must be overcome,
The binding affinity of tBT-HBT by HhC and HhC mutants and the biological function of protein autoprocessing events
was characterized from an inhibitor-dependent enhancement such as Hh cholesterolysis requires a single turnover only.
in FRET with the C−H−Y reporter. By using ΔFRET to Published examples of autoprocessing inhibitors are dominated
report the inhibitor binding, we could evaluate a greater variety by irreversible antagonism, and the few examples of non-
of HhC mutants, including mutants that lack catalytic activity. covalent inhibitors show modest binding affinity.
In two previous studies involving irreversible inhibitors of The findings reported here represent an important exception
HhC, we observed that the covalent adduct formation with C− to the “irreversibility rule” for autoprocessing inhibition. We
H−Y caused a FRET quenching.16,17 The structural changes in speculate that the most effective compound, tBT-HBT (8), is
HhC that underlie these small molecule induced FRET bound by HhC near the interface of the SRR and HINT,
changes in C−H−Y are not yet known. The FRET trapping these two subdomains in a conformation that is
enhancement observed here from C−H−Y with added tBT- capable of binding cholesterol but incapable of catalyzing
HBT was not apparent in C−Y, a FRET control construct cholesterolysis. To our knowledge, tBT-HBT is the tightest
where HhC is omitted. A FRET enhancement was also absent binding noncovalent inhibitor of HhC cholesterolysis, and it
with C-HINT-Y and C-SRR-Y, truncation mutants of HhC appears to be unrivaled among the reported noncovalent
where one of the two subdomains of HhC is deleted. The inhibitors of related forms of protein autoprocessing.
insensitivity of C−Y, C-HINT-Y, and C-SRR-Y to added tBT- In an earlier report, we described reversibly bound small
HBT (Figure S3) suggests that inhibitor recognition requires molecules targeting HhC that activate cholesterol-independent
full-length HhC. autoproteolysis, a nonnative paracatalytic reaction. The most
Results of our mutagenesis study on tBT-HBT binding by effective HhC activator, HAC8, had an AC50 of 9 μM and
HhC are summarized in Figure 4B. Experimental data of accelerated the rate of Hh precursor autoproteolysis by 225-
ΔFRET plotted as a function of the increasing tBT-HBT fold.15 The structures of HAC8 and tBT-HBT are strikingly
concentration for wild-type and HhC point mutants is found in similar (Figure 1, right). The compounds emerged from the
Figure S4. All 13 point mutations weakened the tBT-HBT same chemical screen, and not surprisingly, HAC8 and tBT-
affinity for HhC. We grouped the mutational effects into three HBT seem to share a binding site on HhC (Figure S5). Thus,
general categories: <10-fold loss of the binding affinity (Figure depending on the bound effector, HhC can accelerate the
E https://doi.org/10.1021/acs.biochem.1c00697
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

paracatalytic autoproteolysis of the precursor, as with HAC8, Author Contributions


or suppress the rate of native cholesterolysis, as with tBT-HBT. The manuscript was written through contributions of all
A single allosteric site that can toggle between activation and authors. All authors have given approval to the final version of
inhibition depending on the effector is intriguing and finds the manuscript.
precedent in selected metabolic and signaling enzymes.41−43 Funding
Despite the descriptions of protein autoprocessing domains as National Cancer Institute (grant R01 CA206592); Department
primitive or protozymes, left over from an earlier age, we of Defense (grant W81XWH-14-1-0155); NSF (grant CHE-
continue to find close parallels with extant enzymes in terms of 1048516); and NIH grant (S10 OD012254).
catalytic mechanisms and inhibition strategies. Notes


The authors declare no competing financial interest.

*
ASSOCIATED CONTENT
sı Supporting Information

The Supporting Information is available free of charge at


■ ACKNOWLEDGMENTS
We acknowledge the generous support from the NIH:
CA206592 to B.P.C. and C.W.; GM143714 to JLG;
https://pubs.acs.org/doi/10.1021/acs.biochem.1c00697.
HL067773 to D.F.C.; and a CURE fellowship from the NCI
Compound screening; protein purification; kinetic to A.G.W. We thank Callahan lab members Xiaoyu Zhang and
analysis; creation of HhC mutants; chemical synthesis; Dan Ciulla for help and encouragement and Dr. Juergen
and compound NMR and liquid chromatography mass Schulte for expert advice with NMR. We are grateful to Dr.
spectroscopy (PDF) Alistair Lees and Dr. Lynn Schmitt for access to the
photolabeling equipment. We thank Carl Smith and Polina
Accession Codes Holubovska for help with initial compound screening. We
C-terminal domain of Drosophila melanogaster Hedgehog acknowledge the aid of the Proteomics and Metabolomics
Core Facility at Weill Cornell Medicine for proteomics MS/
protein: UniProtKBQ0293


MS experiments.
AUTHOR INFORMATION
Corresponding Author
■ ABBREVIATIONS
Hh, Hedgehog; HhC, Hedgehog C-terminal domain; HhN,
Brian P. Callahan − Department of Chemistry, Binghamton Hedgehog N-terminal domain; HINT, Hedgehog intein
University, State University of New York, Binghamton, New subdomain; SRR, sterol recognition region; C−H−Y, CFP−
York 13902, United States; orcid.org/0000-0003-4826- HhC−YFP; RV, relative velocity; PAL, photoaffinity labeling;
0162; Email: callahan@binghamton.edu tSP, trans sterol probe; SAR, structure−activity relationship;
tBT-HBT, compound 8; DMSO, dimethyl sulfoxide


Authors
Andrew G. Wagner − Department of Chemistry, Binghamton
University, State University of New York, Binghamton, New REFERENCES
York 13902, United States; orcid.org/0000-0003-4786- (1) Liu, Z.; Xu, J.; He, J.; Zheng, Y.; Li, H.; Lu, Y.; Qian, J.; Lin, P.;
7747 Weber, D. M.; Yang, J.; Yi, Q. A critical role of autocrine sonic
Robert T. Stagnitta − Department of Chemistry, Binghamton hedgehog signaling in human CD138+ myeloma cell survival and drug
University, State University of New York, Binghamton, New resistance. Blood 2014, 124, 2061−2071.
(2) Chen, M.; Carkner, R.; Buttyan, R. The hedgehog/Gli signaling
York 13902, United States
paradigm in prostate cancer. Expert Rev. Endocrinol. Metab. 2011, 6,
Zihan Xu − Department of Chemistry, Binghamton University, 453−467.
State University of New York, Binghamton, New York 13902, (3) Scales, S. J.; de Sauvage, F. J. Mechanisms of Hedgehog pathway
United States activation in cancer and implications for therapy. Trends Pharmacol.
John L. Pezzullo − Department of Chemistry, SUNY-ESF, Sci. 2009, 30, 303−312.
Syracuse, New York 13210, United States (4) di Magliano, M. P.; Hebrok, M. Hedgehog signalling in cancer
Nabin Kandel − Department of Biological Sciences, Center for formation and maintenance. Nat. Rev. Cancer 2003, 3, 903−911.
Biotechnology and Interdisciplinary Studies, Rensselaer (5) Sims-Mourtada, J.; Yang, D.; Tworowska, I.; Larson, R.; Smith,
Polytechnic Institute, Troy, New York 12180, United States D.; Tsao, N.; Opdenaker, L.; Mourtada, F.; Woodward, W. Detection
José-Luis Giner − Department of Chemistry, SUNY-ESF, of canonical hedgehog signaling in breast cancer by 131-iodine-
labeled derivatives of the sonic hedgehog protein. J. Biomed.
Syracuse, New York 13210, United States; orcid.org/ Biotechnol. 2012, 2012, 639562.
0000-0002-8092-5139 (6) Fei, D. L.; Sanchez-Mejias, A.; Wang, Z.; Flaveny, C.; Long, J.;
Douglas F. Covey − Department of Developmental Biology, Singh, S.; Rodriguez-Blanco, J.; Tokhunts, R.; Giambelli, C.; Briegel,
Taylor Family Institute for Innovative Psychiatric Research, K. J.; Schulz, W. A.; Gandolfi, A. J.; Karagas, M.; Zimmers, T. A.;
St. Louis, Missouri 63110, United States; orcid.org/0000- Jorda, M.; Bejarano, P.; Capobianco, A. J.; Robbins, D. J. Hedgehog
0002-9316-8683 signaling regulates bladder cancer growth and tumorigenicity. Cancer
Chunyu Wang − Department of Biological Sciences, Center for Res. 2012, 72, 4449−4458.
Biotechnology and Interdisciplinary Studies, Rensselaer (7) Bissey, P.-A.; Mathot, P.; Guix, C.; Jasmin, M.; Goddard, I.;
Polytechnic Institute, Troy, New York 12180, United States; Costechareyre, C.; Gadot, N.; Delcros, J.-G.; Mali, S. M.; Fasan, R.;
orcid.org/0000-0001-5165-7959 Arrigo, A.-P.; Dante, R.; Ichim, G.; Mehlen, P.; Fombonne, J.
Blocking SHH/Patched Interaction Triggers Tumor Growth Inhib-
Complete contact information is available at: ition through Patched-Induced Apoptosis. Cancer Res. 2020, 80,
https://pubs.acs.org/10.1021/acs.biochem.1c00697 1970−1980.

F https://doi.org/10.1021/acs.biochem.1c00697
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

(8) Petrova, E.; Matevossian, A.; Resh, M. D. Hedgehog (27) Kirby, A. J.; Dutta-Roy, N.; da Silva, D.; Goodman, J. M.; Lima,
acyltransferase as a target in pancreatic ductal adenocarcinoma. M. F.; Roussev, C. D.; Nome, F. Intramolecular General Acid
Oncogene 2015, 34, 263−268. Catalysis of Phosphate Transfer. Nucleophilic Attack by Oxyanions
(9) Jeong, J.; McMahon, A. P. Cholesterol modification of on the PO32- Group. J. Am. Chem. Soc. 2005, 127, 7033−7040.
Hedgehog family proteins. J. Clin. Invest. 2002, 110, 591−596. (28) Page, M. I.; Jencks, W. P. Entropic contributions to rate
(10) Porter, J. A.; Young, K. E.; Beachy, P. A. Cholesterol accelerations in enzymic and intramolecular reactions and the chelate
modification of hedgehog signaling proteins in animal development. effect. Proc. Natl. Acad. Sci. U.S.A. 1971, 68, 1678−1683.
Science 1996, 274, 255−259. (29) Li, Z.; Tharappel, A. M.; Xu, J.; Lang, Y.; Green, C. M.; Zhang,
(11) Buglino, J. A.; Resh, M. D. Identification of conserved regions J.; Lin, Q.; Chaturvedi, S.; Zhou, J.; Belfort, M.; Li, H. Small-molecule
and residues within Hedgehog acyltransferase critical for palmitoyla- inhibitors for the Prp8 intein as antifungal agents. Proc. Natl. Acad. Sci.
tion of Sonic Hedgehog. PLoS One 2010, 5, No. e11195. U.S.A. 2021, 118, No. e2008815118.
(12) Chen, X.; Tukachinsky, H.; Huang, C.-H.; Jao, C.; Chu, Y.-R.; (30) Li, Z.; Fu, B.; Green, C. M.; Liu, B.; Zhang, J.; Lang, Y.;
Tang, H.-Y.; Mueller, B.; Schulman, S.; Rapoport, T. A.; Salic, A. Chaturvedi, S.; Belfort, M.; Liao, G.; Li, H. Cisplatin protects mice
Processing and turnover of the Hedgehog protein in the endoplasmic from challenge of Cryptococcus neoformans by targeting the Prp8
reticulum. J. Cell Biol. 2011, 192, 825−838. intein. Emerg. Microb. Infect. 2019, 8, 895−908.
(13) Huang, C.-H.; Hsiao, H.-T.; Chu, Y.-R.; Ye, Y.; Chen, X. (31) Chan, H.; Pearson, C. S.; Green, C. M.; Li, Z.; Zhang, J.;
Derlin2 protein facilitates HRD1-mediated retro-translocation of Belfort, G.; Shekhtman, A.; Li, H.; Belfort, M. Exploring Intein
sonic hedgehog at the endoplasmic reticulum. J. Biol. Chem. 2013, Inhibition by Platinum Compounds as an Antimicrobial Strategy. J.
288, 25330−25339. Biol. Chem. 2016, 291, 22661−22670.
(14) Tang, H.-Y.; Huang, C.-H.; Zhuang, Y.-H.; Christianson, J. C.; (32) Roland, K. L.; Little, J. W. Reaction of LexA repressor with
diisopropyl fluorophosphate. A test of the serine protease model. J.
Chen, X. EDEM2 and OS-9 are required for ER-associated
Biol. Chem. 1990, 265, 12828−12835.
degradation of non-glycosylated sonic hedgehog. PLoS One 2014, 9,
(33) Ciulla, D. A.; Jorgensen, M. T.; Giner, J.-L.; Callahan, B. P.
No. e92164.
Chemical Bypass of General Base Catalysis in Hedgehog Protein
(15) Smith, C. J.; Wagner, A. G.; Stagnitta, R. T.; Xu, Z.; Pezzullo, J.
Cholesterolysis Using a Hyper-Nucleophilic Substrate. J. Am. Chem.
L.; Giner, J.-L.; Xie, J.; Covey, D. F.; Wang, C.; Callahan, B. P. Soc. 2018, 140, 916−918.
Subverting Hedgehog Protein Autoprocessing by Chemical Induction (34) Hulce, J. J.; Cognetta, A. B.; Niphakis, M. J.; Tully, S. E.;
of Paracatalysis. Biochemistry 2020, 59, 736−741. Cravatt, B. F. Proteome-wide mapping of cholesterol-interacting
(16) Owen, T. S.; Ngoje, G.; Lageman, T. J.; Bordeau, B. M.; Belfort, proteins in mammalian cells. Nat. Methods 2013, 10, 259−264.
M.; Callahan, B. P. Fö rster resonance energy transfer-based (35) West, A. V.; Muncipinto, G.; Wu, H.-Y.; Huang, A. C.;
cholesterolysis assay identifies a novel hedgehog inhibitor. Anal. Labenski, M. T.; Jones, L. H.; Woo, C. M. Labeling Preferences of
Biochem. 2015, 488, 1−5. Diazirines with Protein Biomolecules. J. Am. Chem. Soc. 2021, 143,
(17) Owen, T. S.; Xie, X. J.; Laraway, B.; Ngoje, G.; Wang, C.; 6691−6700.
Callahan, B. P. Active site targeting of hedgehog precursor protein (36) Budelier, M. M.; Cheng, W. W. L.; Bergdoll, L.; Chen, Z.-W.;
with phenylarsine oxide. Chembiochem 2015, 16, 55−58. Janetka, J. W.; Abramson, J.; Krishnan, K.; Mydock-McGrane, L.;
(18) Amakye, D.; Jagani, Z.; Dorsch, M. Unraveling the therapeutic Covey, D. F.; Whitelegge, J. P.; Evers, A. S. Photoaffinity labeling with
potential of the Hedgehog pathway in cancer. Nat. Med. 2013, 19, cholesterol analogues precisely maps a cholesterol-binding site in
1410−1422. voltage-dependent anion channel-1. J. Biol. Chem. 2017, 292, 9294−
(19) Yauch, R. L.; Dijkgraaf, G. J. P.; Alicke, B.; Januario, T.; Ahn, C. 9304.
P.; Holcomb, T.; Pujara, K.; Stinson, J.; Callahan, C. A.; Tang, T.; (37) Cheng, W. W. L.; Chen, Z.-W.; Bracamontes, J. R.; Budelier, M.
Bazan, J. F.; Kan, Z.; Seshagiri, S.; Hann, C. L.; Gould, S. E.; Low, J. M.; Krishnan, K.; Shin, D. J.; Wang, C.; Jiang, X.; Covey, D. F.; Akk,
A.; Rudin, C. M.; de Sauvage, F. J. Smoothened mutation confers G.; Evers, A. S. Mapping two neurosteroid-modulatory sites in the
resistance to a Hedgehog pathway inhibitor in medulloblastoma. prototypic pentameric ligand-gated ion channel GLIC. J. Biol. Chem.
Science 2009, 326, 572−574. 2018, 293, 3013−3027.
(20) Mafi, A.; Purohit, R.; Vielmas, E.; Lauinger, A. R.; Lam, B.; (38) Presolski, S. I.; Hong, V. P.; Finn, M. G. Copper-Catalyzed
Cheng, Y.-S.; Zhang, T.; Huang, Y.; Kim, S.-K.; Goddard, W. A., III; Azide-Alkyne Click Chemistry for Bioconjugation. Curr. Protoc. Chem.
Ondrus, A. E. Hedgehog proteins create a dynamic cholesterol Biol. 2011, 3, 153−162.
interface. PLoS One 2021, 16, No. e0246814. (39) Webb, B.; Sali, A. Comparative Protein Structure Modeling
(21) Purohit, R.; Peng, D. S.; Vielmas, E.; Ondrus, A. E. Dual roles Using MODELLER. Curr. Protoc. Bioinf. 2016, 54, 5.6.1−5.6.37.
of the sterol recognition region in Hedgehog protein modification. (40) Trott, O.; Olson, A. J. AutoDock Vina: improving the speed
Commun. Biol. 2020, 3, 250. and accuracy of docking with a new scoring function, efficient
(22) Hall, T. M. T.; Porter, J. A.; Young, K. E.; Koonin, E. V.; optimization, and multithreading. J. Comput. Chem. 2010, 31, 455−
Beachy, P. A.; Leahy, D. J. Crystal structure of a Hedgehog 461.
autoprocessing domain: homology between Hedgehog and self- (41) Laufkötter, O.; Hu, H.; Miljković, F.; Bajorath, J. Structure- and
splicing proteins. Cell 1997, 91, 85−97. Similarity-Based Survey of Allosteric Kinase Inhibitors, Activators, and
(23) Xie, J.; Du, Z.; Callahan, B.; Belfort, M.; Wang, C. H, (1)(3)C Closely Related Compounds. J. Med. Chem. 2021, DOI: 10.1021/
and (1)(5)N NMR assignments of a Drosophila Hedgehog acs.jmedchem.0c02076.
autoprocessing domain. Biomol. NMR Assignments 2014, 8, 279−281. (42) Meng, H.; McClendon, C. L.; Dai, Z.; Li, K.; Zhang, X.; He, S.;
(24) Zhao, J.; Ciulla, D. A.; Xie, J.; Wagner, A. G.; Castillo, D. A.; Shang, E.; Liu, Y.; Lai, L. Discovery of Novel 15-Lipoxygenase
Zwarycz, A. S.; Lin, Z.; Beadle, S.; Giner, J.-L.; Li, Z.; Li, H.; Banavali, Activators To Shift the Human Arachidonic Acid Metabolic Network
N.; Callahan, B. P.; Wang, C. General Base Swap Preserves Activity toward Inflammation Resolution. J. Med. Chem. 2016, 59, 4202−
and Expands Substrate Tolerance in Hedgehog Autoprocessing. J. Am. 4209.
(43) Cockrell, G. M.; Zheng, Y.; Guo, W.; Peterson, A. W.; Truong,
Chem. Soc. 2019, 141, 18380−18384.
J. K.; Kantrowitz, E. R. New paradigm for allosteric regulation of
(25) Fersht, A. R.; Kirby, A. J. Intramolecular nucleophilic catalysis
Escherichia coli aspartate transcarbamoylase. Biochemistry 2013, 52,
in the hydrolysis of substituted aspirin acids. J. Am. Chem. Soc. 1968,
8036−8047.
90, 5826−5832.
(26) Fersht, A. R.; Kirby, A. J. Hydrolysis of aspirin. Intramolecular
general base catalysis of ester hydrolysis. J. Am. Chem. Soc. 1967, 89,
4857−4863.

G https://doi.org/10.1021/acs.biochem.1c00697
Biochemistry XXXX, XXX, XXX−XXX

You might also like