ACS Publications. Most Trusted. Most Cited. Most Read
Illuminating Neuropeptide Y Y4 Receptor Binding: Fluorescent Cyclic Peptides with Subnanomolar Binding Affinity as Novel Molecular Tools
My Activity

Figure 1Loading Img
  • Open Access
Article

Illuminating Neuropeptide Y Y4 Receptor Binding: Fluorescent Cyclic Peptides with Subnanomolar Binding Affinity as Novel Molecular Tools
Click to copy article linkArticle link copied!

  • Jakob Gleixner
    Jakob Gleixner
    Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, Germany
  • Sergei Kopanchuk
    Sergei Kopanchuk
    Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, Estonia
  • Lukas Grätz
    Lukas Grätz
    Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, Germany
    More by Lukas Grätz
  • Maris-Johanna Tahk
    Maris-Johanna Tahk
    Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, Estonia
  • Tõnis Laasfeld
    Tõnis Laasfeld
    Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, Estonia
  • Santa Veikšina
    Santa Veikšina
    Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, Estonia
  • Carina Höring
    Carina Höring
    Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, Germany
  • Albert O. Gattor
    Albert O. Gattor
    Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, Germany
  • Laura J. Humphrys
    Laura J. Humphrys
    Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, Germany
  • Christoph Müller
    Christoph Müller
    Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, Germany
  • Nataliya Archipowa
    Nataliya Archipowa
    Institute of Biophysics and Physical Biochemistry, Faculty of Biology and Preclinical Medicine, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, Germany
  • Johannes Köckenberger
    Johannes Köckenberger
    Department of Chemistry and Pharmacy, Molecular and Clinical Pharmacy, Friedrich-Alexander-University Erlangen-Nürnberg, Nikolaus-Fiebiger-Straße 10, D-91058 Erlangen, Germany
  • Markus R. Heinrich
    Markus R. Heinrich
    Department of Chemistry and Pharmacy, Molecular and Clinical Pharmacy, Friedrich-Alexander-University Erlangen-Nürnberg, Nikolaus-Fiebiger-Straße 10, D-91058 Erlangen, Germany
  • Roger Jan Kutta
    Roger Jan Kutta
    Institute of Physical and Theoretical Chemistry, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93053 Regensburg, Germany
  • Ago Rinken*
    Ago Rinken
    Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, Estonia
    *Email: [email protected]. Tel: (+372) 7375-249.
    More by Ago Rinken
  • Max Keller*
    Max Keller
    Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, Germany
    *Email: [email protected]. Tel: (+49) 941-9433329.
    More by Max Keller
Open PDFSupporting Information (2)

ACS Pharmacology & Translational Science

Cite this: ACS Pharmacol. Transl. Sci. 2024, 7, 4, 1142–1168
Click to copy citationCitation copied!
https://doi.org/10.1021/acsptsci.4c00013
Published March 20, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

The neuropeptide Y (NPY) Y4 receptor (Y4R), a member of the family of NPY receptors, is physiologically activated by the linear 36-amino acid peptide pancreatic polypeptide (PP). The Y4R is involved in the regulation of various biological processes, most importantly pancreatic secretion, gastrointestinal motility, and regulation of food intake. So far, Y4R binding affinities have been mostly studied in radiochemical binding assays. Except for a few fluorescently labeled PP derivatives, fluorescence-tagged Y4R ligands with high affinity have not been reported. Here, we introduce differently fluorescence-labeled (Sulfo-Cy5, Cy3B, Py-1, Py-5) Y4R ligands derived from recently reported cyclic hexapeptides showing picomolar Y4R binding affinity. With pKi values of 9.22–9.71 (radioligand competition binding assay), all fluorescent ligands (1619) showed excellent Y4R affinity. Y4R saturation binding, binding kinetics, and competition binding with reference ligands were studied using different fluorescence-based methods: flow cytometry (Sulfo-Cy5, Cy3B, and Py-1 label), fluorescence anisotropy (Cy3B label), and NanoBRET (Cy3B label) binding assays. These experiments confirmed the high binding affinity to Y4R (equilibrium pKd: 9.02–9.9) and proved the applicability of the probes for fluorescence-based Y4R competition binding studies and imaging techniques such as single-receptor molecule tracking.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2024 The Authors. Published by American Chemical Society
In humans, four functional neuropeptide Y (NPY) receptors (Y1R, Y2R, Y4R, Y5R), all members of the superfamily of G-protein-coupled receptors, mediate the action of the neuropeptides NPY, peptide YY (PYY), and pancreatic polypeptide (PP). (1) Within the family of these linear 36-amino-acid peptides, PP exhibits a clear preference for the Y4R and is thus considered the primary activator of Y4R. (2) PP and Y4R are involved in the regulation of food intake, pancreatic secretion, gastrointestinal motility, and anxiety. (3−13) Therefore, the Y4R represents a potential target for the treatment of obesity and depression. (12−15)
Development processes aiming for new Y4R ligands require appropriate binding assays to determine Y4R affinities. So far, this has been primarily accomplished by radioligand competition binding studies using 3H- or 125I-labeled peptidic Y4R ligands. Although radiochemical binding assays represent powerful methods, in particular with respect to sensitivity and quantification of bound and free labeled ligand, several drawbacks are associated with the use of radioligands: special safety precautions, the need for specialized laboratories, permits for radioisotope handling, and the high cost of radiolabeled probes. These are some of the reasons for the continuously growing use of fluorescence-based techniques to study receptor–ligand binding. Moreover, most fluorescence-based binding assays allow, in contrast to radiochemical assays, investigations under homogeneous conditions, i.e., there is no need to separate free (unbound) ligand from receptor-bound ligand prior to measurement. (16−24) In particular, this is advantageous for kinetic studies. Another factor promoting luminescence-based assays is the availability of instruments such as multimode plate readers allowing high-throughput measurements. A bottleneck with respect to the development and use of fluorescence-based binding assays is the limited availability of fluorescent ligands exhibiting, all at once, high receptor affinity, favorable physicochemical and photophysical properties, appropriate binding kinetics (reversible binding), and adequate chemical stability as well as photostability.
For the Y4R, few fluorescently labeled ligands have been reported, Sulfo-Cy5-labeled [K4]hPP, (25) Sulfo-Cy5-labeled [K4,Nle17,30]hPP (1), (26) 5-TAMRA-labeled [K18,Nle17,30]hPP (2), (27) and the UR-KK236 (28)-derived Sulfo-Cy5.5-labeled hexapeptide 3 (29) (for structures of 13 see Figure 1A). With a Kd value of 5.6 nM, (25) only Sulfo-Cy5-[K4]hPP shows a dissociation constant <10 nM among these probes (Kd of 1 and 2: 11 and 26 nM, respectively; Ki of 3: 33 nM). Compared to the PP derivatives Sulfo-Cy5-[K4]hPP, 1 and 2, fluorescent ligand 3 exhibits a considerably lower molecular weight, which is considered favorable. However, with a Ki value of 33 nM, compound 3 exhibits only moderate Y4R binding affinity for application as a molecular tool.

Figure 1

Figure 1. (A) Structures and Y4R affinities of reported Y4R fluorescent ligands. 1 (26) and 2 (27) represent derivatives of the endogenous Y4R ligand hPP labeled with sulfo-Cy5 (S0223) and 5-TAMRA, respectively. Compound 3 (29) represents a derivative of the hexapeptide UR-KK236 (28) labeled with Sulfo-Cy5.5 (lumiprobe, ref no. 7330). (B) Structures of the reported cyclic Y4R ligands 47 showing high binding affinity to Y4R (pKi > 10). (30,31)

The present study aimed to design, synthesize, and characterize fluorescent Y4R ligands with low molecular weight and higher affinity to Y4R compared to reported probes. To achieve this, we exploited the recent discovery of the cyclic hexapeptides 4 and 5 showing two-digit picomolar Y4R affinities (Figure 1B). (30) Modification of peptide 4 resulted in the amine-functionalized peptide 6 (UR-JG93), its propionylated congener UR-JG102 (7) (Figure 1B) and the radiolabeled analogue [3H]UR-JG102, all showing very high Y4R affinity comparable to that of 4 and 5. (31)
The same approach was followed for the synthesis of fluorescent Y4R ligands, accessible by conjugation of fluorescent dyes to precursor peptide 6 (Figure 2, approach 1). Additionally, a second approach was explored by replacing Arg1 in peptide 4 by an alkyne-functionalized arginine enabling conjugation to a fluorescent dye by click chemistry (Figure 2, approach 2). The synthesized fluorescent ligands were studied with respect to Y4R binding (radiochemical competition binding assay) and Y4R agonistic activity (functional assays). Moreover, Y4R binding of selected fluorescent ligands was investigated using different luminescence-based methods (flow cytometry, fluorescence anisotropy (FA), NanoBRET assay, confocal microscopy, wide-field epifluorescence, and TIRF microscopy with single-particle tracking).

Figure 2

Figure 2. Study design of the present work: synthesis and characterization of fluorescent probes with high Y4R binding affinity derived from recently reported cyclic hexapeptides (6, (31) 4 (30)).

Results and Discussion

Click to copy section linkSection link copied!

Synthesis and Chemical Stability

In addition to the previously described amino-functionalized peptide 6, (31) an alkyne-functionalized precursor peptide (11) was prepared for fluorescence labeling. The cyclic peptide 11 was prepared via the N-terminally succinylated linear peptide 10, which was synthesized by solid-phase peptide synthesis (SPPS) involving the reported Nω-carbamoylated arginine building blocks 8 (32) (position 3) and 9 (29) (position 1) (structures of 8 and 9 shown in Figure 3).

Figure 3

Figure 3. Structures of the reported Nω-carbamoylated arginines 8 (32) and 9 (29) used in SPPS for the preparation of precursor peptide 11.

The amino-functionalized arginine derived from building block 8 is required for peptide cyclization, and the incorporation of 9 provides the alkyne functional group (Scheme 1). The fully deprotected peptide 10 was cyclized in solution via the succinyl carboxylic group and the primary amino group of the carbamoylated arginine in position 3 (coupling reagents: HOBt, PyBOP, DIPEA), yielding peptide 11 in 21% yield. To note, the synthesis of the alkyne-functionalized peptide 11 is more laborious compared to the preparation of the reported amine-functionalized precursor peptide 6, (31) as the synthesis of 11 requires two different Nω-carbamoylated arginines. The major purpose for the synthesis of peptide 11 and a fluorescent dye conjugate derived from 11 was to explore how strongly this structural modification, including the attachment of a fluorescent dye by click chemistry, affects binding to Y4R.

Scheme 1

Scheme 1. Synthesis of the Cyclic Peptide 11 via SPPSa

aReagents and conditions: (a) Fmoc amino acid/HOBt/HBTU/DIPEA (5/5/4.9/10 equiv), DMF/NMP 8:2 v/v, 38 °C, 2 × 45 min (“double” coupling); Fmoc deprotection (following amino acid coupling): 20% piperidine in DMF/NMP 8:2 v/v, rt, 2 × 10 min; (b) building block 8 or 9/HOBt/HBTU/DIPEA (3/3/2.9/6 equiv), DMF/NMP 8:2 v/v, 38 °C, 16 h (single coupling); Fmoc deprotection (following amino acid coupling): as under (a); (c) succinic anhydride/DIPEA (10/10 equiv), DMF/NMP 8:2 v/v, 38 °C, 45 min; (d) (1) TFA/CH2Cl2 1:3 v/v, rt, 2 × 20 min; (2) TFA/H2O 95:5 v/v, rt, 5 h; overall yield: 37%; (e) peptide cyclization: HOBt/PyBOP/DIPEA (3/2.9/6 equiv), DMF/NMP 8:2 v/v, rt, 16 h, 21%.

For the preparation of the fluorescent peptides, different fluorescent dyes with various reactive groups were used (Scheme 2): the indolinium-type dyes Cy3B and Sulfo-Cy5 (S0223), used as succinimidyl esters (compounds 12 and 13), the pyrylium dye Py-1 (14, yielding a pyridinium-type cyanine dye after reaction with primary amines), and the azido-functionalized pyridinium-type cyanine dye 15, which was obtained from the pyrylium dye Py-5 and 3-azidopropane-1-amine (Scheme S1, Supporting Information). The fluorescently labeled ligands 1618 were prepared from precursor 6 and the activated fluorescent dyes 12, 13, and 14, respectively, using DIPEA as a base and N,N-dimethylformamide (DMF) as solvent. 1618 were obtained in 34–69% yield. Fluorescent ligand 19 was synthesized from 11 and 15 via a copper(I)-catalyzed “click-reaction” in 27% yield.

Scheme 2

Scheme 2. Synthesis of the Fluorescent Y4R Ligands 1619a

aReagents and conditions: (a) DIPEA, DMF, rt, 2 h, 34–69%; (b) CuSO4, sodium ascorbate, H2O/NMP 1:1 v/v, rt, 2 h, 27%.

The chemical stability of 1618 was investigated in phosphate-buffered saline (PBS, pH 7.4) at room temperature over 48 h. The Cy3B- and Sulfo-Cy5 labeled compounds 16 and 17 showed excellent stability (Figure S1A,B, Supporting Information). The Py-1 labeled ligand 18 showed a slow decomposition (Figure S1C, Supporting Information). However, for at least up to 6 h, the degradation was marginal allowing an investigation of this fluorescent ligand in cellular assays used in this study.

Investigation of YR Binding in Radiochemical Binding Assays

Y1, Y2, Y4, and Y5 receptor binding of 11 and 1619 was determined using reported whole-cell radioligand competition binding assays (Y1: Keller et al., (33) Y2: Konieczny et al., (34) Y4: Konieczny et al., (30) Y5: Kuhn et al. (35)). With pKi values >9, all studied compounds showed high Y4R affinities (Table 1, competition binding curves shown in Figure S2, Supporting Information). Since the Ki values were at least 500-fold lower compared to the Ki determined at Y1, Y2, and Y5 receptors, 11 and 1619 also exhibit pronounced Y4R selectivity.
Table 1. Y Receptor Binding Data of hPP, 46, 11, and 1619
 pKi ± SEM/Ki [nM]
compd.hY1RahY2RbhY4RchY5Rd
hPP6.4/440e<5.5/>3000e10.02 ± 0.06/0.10f7.8/17e
4<5.5/>3000g<5/>10000g10.36 ± 0.11/0.048g<5.5/>3000g
5<5/>10000g<5/>10000g10.48 ± 0.04/0.033g<5.5/>3000g
6<6/>1000h6.04 ± 0.07/950h10.04 ± 0.04/0.093h<6/>1000h
11<5/>10000<5/>100009.38 ± 0.04/0.42<5.5/>3000
166.91 ± 0.01/120<6/>10009.71 ± 0.09/0.216.32 ± 0.04/480
17<6/>1000<6/>10009.48 ± 0.09/0.35<6/>1000
18<6/>1000<6/>10009.22 ± 0.03/0.616.08 ± 0.05/840
19<6/>1000<6/>10009.22 ± 0.06/0.62<6/>1000
a

Determined by competition binding at SK-N-MC neuroblastoma cells using [3H]UR-MK299 (Kd = 0.054 nM, c = 0.15 nM) as radioligand.

b

Determined by competition binding with [3H]propionyl-pNPY (Kd = 0.14 nM, (34) c = 0.5 nM) at CHO-hY2R cells.

c

Determined by competition binding at CHO-hY4R-Gqi5-mtAEQ cells using [3H]UR-KK200 (Kd = 0.67 nM, (35) c = 1 nM) as radioligand.

d

Determined by competition binding at HEC-1B-hY5R cells using [3H]propionyl-pNPY (Kd = 11 nM, (26) c = 5 nM) as radioligand.

e

Berlicki et al. (reported Ki values were converted to pKi values). (36)

f

Wirth et al. (37)

g

Konieczny et al. (30)

h

Gleixner et al. (31) Data represent mean values ± SEM (pKi) or mean values (Ki) from 3 to 4 independent experiments performed in triplicate.

Consequently, incorporation of the alkyne-functionalized carbamoylated arginine instead of Arg1 in the parent compound 4 (compound 11) and coupling to a fluorescent dye (19) only slightly affects Y4R binding (Table 1). Likewise, the conjugation of fluorescent dyes to precursor peptide 6 did not alter Y4R binding, underlining the addressed position in 6 as favorable for the attachment of bulky moieties.

Functional Studies at Y4R

Y4R agonistic activities of 11 and 1619 were investigated in a Ca2+ aequorin assay, (25,34) measuring the increase in cytosolic Ca2+, and a miniGsi recruitment assay, (37) detecting binding of an artificial G-protein to the Y4R. Additionally, compounds 16 and 17 were studied in a cAMP CAMYEN assay (31) to measure changes in cytosolic cAMP levels. The CAMYEN assay uses a Nanoluciferase as the BRET donor in place of the “Renilla” luciferase found in the traditional CAMYEL BRET assay, while retaining the rest of the sensor. Due to the increased brightness of the Nanoluciferase compared to the “Renilla” luciferase, the CAMYEN biosensor has an increased bioluminescence signal readout, as well as a bigger window in the signal-to-baseline of the cAMP responses, reducing saturation of the sensor. This allows for easier and faster detection of CAMYEN sensor activation by plate readers and a greater potential for determining ligand efficacy (i.e., partial agonism) when compared to the CAMYEL assay.
To note, control experiments with hPP in the absence and presence of dummy fluorescent ligands showed that the readout of the aequorin (Ca2+-assay) and Nanoluciferase (Nluc, miniGsi assay) luminescence is only slightly affected by the fluorophore at high fluorescent ligand concentrations (Figure S6, Supporting Information).
In the Ca2+ and mini Gsi assays, the studied peptides displayed partial agonism with efficacies ranging from 56 to 71 and 43 to 72%, respectively (Table 2, concentration–response curves shown in Figures S3 and S4, Supporting Information).
Table 2. Y4R Agonistic Activities of hPP, 46, 11, and 1619a
 Ca2+-aequorinminiGsi recruitmentCAMYEN cAMP
cmpd.pEC50 ± SEM/EC50 [nM]Emax ± SEM/%pEC50 ± SEM/EC50 [nM]Emax ± SEM/%pEC50 ± SEM/EC50 [nM]Emax ± SEM/%
hPP7.90 ± 0.2/17b100b8.94 ± 0.01/1.1b100b9.85 ± 0.09/0.15b100b
48.57 ± 0.03/2.7c81 ± 1cn.d.n.d.9.79 ± 0.1/0.17b113 ± 9b
59.00 ± 0.07/0.99c84 ± 2c8.74 ± 0.04/1.861 ± 2n.d.n.d.
67.76 ± 0.05/18b69 ± 3b8.60 ± 0.03/2.5b62 ± 2bn.d.n.d.
117.01 ± 0.01/9856 ± 37.2 ± 0.1/7362 ± 1n.d.n.d.
167.8 ± 0.1/1966 ± 27.4 ± 0.1/4172 ± 99.58 ± 0.06/0.2693 ± 5
177.41 ± 0.05/4071 ± 37.07 ± 0.09/9370 ± 99.65 ± 0.05/0.2290 ± 3
187.3 ± 0.1/6561 ± 58.15 ± 0.01/7.144 ± 2n.d.n.d.
197.25 ± 0.07/5864 ± 27.38 ± 0.02/4243 ± 1n.d.n.d.
a

Agonistic potencies (pEC50, EC50) and efficacies Emax (relative to the maximum effect elicited by 1 μM hPP, Emax = 100%) determined in a Ca2+-aequorin assay (CHO-hY4R-Gqi5-mtAEQ cells), a miniGsi recruitment assay (HEK293T-NlucN-mGsi/Y4R-NlucC cells) and in a CAMYEN cAMP assay (HEK293T-CAMYEN-hY4R cells).

b

Gleixner et al. (31)

c

Konieczny et al. (30) Data represent mean values ± SEM (pEC50, Emax) or mean values (EC50) from three or four independent experiments performed in triplicate. n.d. = not determined.

As previously observed for peptidic Y4R agonists, (31) the potencies (pEC50) obtained from the Ca2+ and miniGsi assay were consistently lower than the respective binding affinities (pKi values; Tables 1 and 2), which can be attributed to the nonequilibrium conditions in the case of the functional assays. In contrast, the pEC50 values of 16 and 17, determined in the cAMP CAMYEN assay, were in good agreement with the pKi values from the radiochemical competition binding assay, which is explainable by a downstream signal amplification for this type of functional assay.

Fluorescence Characterization

The excitation and emission spectra of 1619 and the emission quantum yield were recorded in PBS and PBS supplemented with 1% bovine serum albumin (BSA, spectra shown in Figure S7, Supporting Information). The Cy3B-labeled ligand 16 showed the highest quantum yield with 67–69% (Table 3). The small spectral shifts of the maxima of the excitation and emission spectra in the presence of BSA may indicate some unspecific interactions to protein surfaces. However, as the fluorescence quantum yield is not altered in the presence of BSA, these interactions have no significant impact on the general photophysics of the Cy3B dye. Interestingly, comparison to the Cy3B derivative, which was obtained by preparation of an amide from succinimidyl ester 12 (Cy3B-SE, cf. Scheme 2) and 2-aminoethanol, (38) reveals that the peptide moiety in fluorescent ligand 16 mediates an increase in fluorescence quantum yield by ca. 10%.
Table 3. Maxima of the Excitation and Emission Spectra and Fluorescence Quantum Yields of 1619, Determined in PBS (pH = 7.4) and PBS Containing 1% BSA
  λex [nm]/λem [nm]/Δ [eV]Φ [%]
compd.dyePBSPBS with 1% BSAPBSPBS with 1% BSA
16Cy3B571/584/0.048572/586/0.0526967
17Sulfo-Cy5648/665/0.049657/670/0.0372633
18py-1517/633/0.439506/595/0.367236
19py-5496/692/0.708515/644/0.482424
For the Sulfo-Cy5 labeled compound 17, the bathochromic shift is more pronounced and a small increase in fluorescence quantum yield is observed in the presence of BSA, which was also previously observed for peptidic neurotensin receptor ligands. (39) In the case of the pyridinium-type cyanine dye-labeled ligands 18 and 19, a hypsochromic shift of the first electronic absorption and a considerably higher fluorescence quantum yield in the presence of BSA is observed (Table 3). This phenomenon, also reported for other Py-1- and Py-5-labeled receptor ligands, (40−42) can be explained by strongly reduced degrees of freedom in configurational space of the fluorophore due to binding interactions with amino acid residues on the BSA surface, which is evidenced by the marked reduction in the Stokes shift observed for 18 and 19. In contrast to the Cy3B dye in 16, the conformationally less restricted dyes in 1719 show a Stokes shift reduction and enhancement of the fluorescence quantum yield in the presence of BSA. It is tempting that in the dyes with higher flexibility (as in 1719) the excitation energy is rapidly released via pronounced geometric changes favoring internal conversion rather than fluorescence and that rigidification due to binding to BSA blocks these deactivation channels making fluorescence a competing process. Currently, detailed mechanistic studies are conducted in our laboratories to elucidate the excited state deactivation of the pyridinium-type cyanine dyes in 18 to 19 on a molecular level.

Flow Cytometric Y4R Binding Studies

Y4R binding of 1618 was investigated in flow cytometry-based binding experiments using CHO-hY4R-Gqi5-mtAEQ cells, (25) which were also used for radiochemical binding assays and the Ca2+ aequorin assay. Saturation binding with the Cy3B-labeled ligand 16 was performed in a sodium-free buffer (buffer I, composition see the Experimental Section) and in sodium-containing (137 mM Na+) DPBS (Figure 4A). This comparison was of interest as previous studies revealed marked differences in Y4R affinities of Y4R (partial) agonists depending on the presence or absence of sodium in the binding buffer (up to 20-fold lower affinity in the presence of Na+ at a physiological concentration). (26,31,35) As receptor binding studies should generally be performed under physiological-like conditions, i.e., in sodium-containing buffers, 17 and 18 were only studied in DPBS. The flow cytometric investigation of the cell viability by propidium-based live/dead staining revealed that in sodium-free buffer (buffer I) the major fraction of CHO-hY4R-Gqi5-mtAEQ cells was nonviable already after 15 min of incubation at 22 °C (Figure S8A, Supporting Information). Over time (followed up to 6 h), the fraction of intact cells further decreased. In contrast, CHO-hY4R-Gqi5-mtAEQ cells suspended in DPBS were largely intact even after 6 h of incubation (Figure S8B, Supporting Information). As the dissociation of 16 from CHO-hY4R-Gqi5-mtAEQ cells was followed over 6 h in buffer I (see below), a data analysis based on the intact cell population was not reasonable for the whole experiment. Therefore, for all flow cytometric binding experiments performed with 16 in buffer I (saturation binding, association and dissociation kinetics, competition binding), data analysis was performed based on the nonviable cell population (corresponds to population P2 in Figure S8A, Supporting Information).

Figure 4

Figure 4. Flow cytometric saturation binding of 16 (A), 17 (B), and 18 (C) at whole CHO-hY4R-Gqi5-mtAEQ cells at 22 ± 2 °C. (A) Representative saturation isotherms (red circle) of 16 from experiments performed in sodium-free buffer (buffer I) and sodium-containing buffer (DPBS, 137 mM Na+). (B, C) Representative saturation isotherms (red circle) of 17 (B) and 18 (C) from experiments performed in sodium-containing buffer (DPBS). Unspecific binding (blue squares) was determined in the presence of 1 μM hPP (A–C). Total and unspecific binding data represent mean values ± SEM. Specific binding, representing calculated values ± propagated error, were fitted according to an equation describing a hyperbolic isotherm (binding-saturation: one site-specific binding, GraphPad Prism 5).

Saturation binding studies with fluorescent ligand 16 gave a slightly lower Kd value for buffer I (Kd: 0.30 nM) compared to DPBS (Kd: 0.70 nM). Notably, analysis of the saturation binding data based on the intact cell population (performed as a control), yielded the same Kd values as obtained from the analysis based on the nonviable cell population, showing that the loss of cell integrity had no or only little impact on Y4R binding of 16 (cf. Figure S9, Supporting Information). With Kd values <1 nM, fluorescent ligand 16 showed higher Y4R binding affinity compared to previously reported fluorescently labeled Y4R ligands. (25−27,29) Saturation binding with 17 and 18 yielded Kd values of 0.93 and 0.51 nM, respectively (Table 4 and Figure 4B,C) indicating that the type of fluorophore has almost no effect on Y4R binding.
Table 4. Parameters Characterizing Y4R Binding of 1618 Determined in Flow Cytometric Binding Assays at 22 ± 2 °C Using CHO-hY4R-Gqi5-mtAEQ Cells
  saturation bindingbinding kinetics
cmpd.bufferpKd/Kd [nM]akobs(mono) [min–1]bkobs(bi,fast) [min–1]ckobs(bi,slow) [min–1]ckoff [min–1]dkon(mono) [nM–1 min–1]ekon(bi,fast) [nM–1 min–1]e kon(bi,slow) [nM–1 min–1]eKd(kin) [nM]f
16buffer I (Na+-free)9.56 ± 0.09/0.300.13 ± 0.02n.a.0.0069 ± 0.00060.41 ± 0.07n.a.0.017 ± 0.004
16DPBS (137 mM Na+)9.16 ± 0.05/0.700.18 ± 0.041.0 ± 0.20.012 ± 0.0010.24 ± 0.051.3 ± 0.3with kon(mono): 0.05 ± 0.02
0.048 ± 0.0030.050 ± 0.006with kon(bi,fast): 0.009 ± 0.002
with kon(bi,slow): 0.25 ± 0.05
17DPBS (137 mM Na+)9.03 ± 0.02/0.930.15 ± 0.011.89 ± 0.050.026 ± 0.0020.12 ± 0.011.87 ± 0.05with kon(mono): 0.21 ± 0.03
0.071 ± 0.0080.045 ± 0.009with kon(bi,fast): 0.007 ± 0.001
with kon(bi,slow): 0.29 ± 0.07
18DPBS (137 mM Na+)9.3 ± 0.1/0.510.10 ± 0.020.7 ± 0.10.010 ± 0.0010.18 ± 0.041.3 ± 0.2with kon(mono): 0.06 ± 0.02
0.034 ± 0.0010.048 ± 0.004with kon(bi,fast): 0.008 ± 0.002
with kon(bi,slow): 0.20 ± 0.05
a

Equilibrium dissociation constant expressed as pKd (mean values ± SEM) and Kd (mean values) obtained from at least three independent experiments (performed in triplicate).

b

Observed association rate constant obtained by monophasic fitting (exponential rise to a maximum); mean values ± SEM from three independent experiments (performed in duplicate).

c

Observed association rate constant obtained by biphasic fitting; mean values ± SEM from three or four independent experiments (performed in duplicate).

d

Dissociation rate constant obtained from three-parameter monophasic fits (exponential decline); mean values ± SEM from three or four independent experiments (performed in duplicate).

e

Association rate constant ± propagated error calculated from kobs(mono), kobs(bi,fast), or kobs(bi,slow) values, the respective koff value, and the ligand concentration used for the association experiments.

f

Kinetically derived dissociation constant ± propagated error calculated from koff and kon values. n.a.: not applicable

All data obtained from association binding experiments in DPBS correspond better to a biphasic than to a monophasic curve (Figure 5A,C), as supported by a statistical extra sum-of-squares F-test (one-phase association vs two-phase association, GraphPad Prism 5) giving P values <0.002. In contrast, the association of 16 in sodium-free buffer was monophasic (F-test derived P-value: 0.55) (Figure 5A). For all studied fluorescent ligands (1618), association experiments indicated that equilibrium is reached after approximately 1 h (Figure 5A,C). Worth mentioning, the structurally related radioligand [3H]UR-JG102 ([3H]7), also studied at CHO-hY4R-Gqi5-mtAEQ cells in sodium-free (buffer I) and sodium-containing (DPBS) buffer, showed similar kinetics being monophasic in buffer I and biphasic in DPBS. However, compared to 1618, the biphasic character of the association curve of [3H]7 in DPBS was considerably more pronounced. (31) As described for [3H]7, the first (fast) association phase could represent association to a subpopulation of receptors that are sodium-bound. (31) It is well known that sodium stabilizes the inactive receptor conformation of GPCRs, (43) presumably being better accessible for receptor agonists due to a more open passage to the ligand binding pocket. (43,44) The second (slow) association could represent binding to a population of nonsodium-bound receptors (note: at 150 mM sodium a GPCR population is not necessarily saturated with sodium (44)), which are prone to adapt the active receptor conformation. At first sight, this disagrees with the observed proportions of the initial (fast) and the second (slow) association phase, which varied considerably for the fluorescent ligands 1618 (values given in the caption of Figure 5). However, as the receptor conformation also depends on the coupling to the G-protein and as G-protein binding may influence ligand binding, (45) the inconsistency of the proportions could be explained by a ligand bias with respect to the G-protein coupling efficiency of the agonist-bound receptor and in terms of G-protein modulated ligand affinity. This is in accordance with the monophasic association of 16 observed in the FA assay using Y4R displaying BBVs instead of whole cells (see below). It should also be kept in mind that during the association process, dissociation takes place, which codetermines the result of an association experiment. Furthermore, provided that the receptor populations are interconvertible and that equilibria among different receptor conformations can be modulated by the ligand, (46) a readjustment of the equilibrium between distinct receptor populations may occur upon ligand binding, which can be differently pronounced for different ligands.

Figure 5

Figure 5. Binding kinetics of 1618 determined by flow cytometry at whole CHO-hY4R-Gqi5-mtAEQ cells at 22 ± 2 °C. (A) Association of 16 to the hY4R under sodium-free conditions (buffer I) and in sodium-containing buffer (DPBS, 137 mM Na+). Concentration of 16: 0.3 and 0.7 nM, respectively. Proportion of the initial fast association (two-phase association fit, GraphPad Prism 5): 38 ± 4% (mean ± SEM). (B) Dissociation of 16 from the hY4R determined in buffer I and DPBS. The dissociation was initiated after 1.5 h of preincubation with 16 (c = 1.5 nM (Na+-free) and 3.5 nM (137 mM Na+)) by the addition of an excess of hPP (1000-fold) and 5 (100-fold). Plateau values of the three-parameter fits (monophasic exponential decline): 13% (Na+-free), 22% (137 mM Na+). (C) Association of 17 (c = 1 nM) and 18 (c = 0.5 nM) to the hY4R determined in DPBS (137 mM Na+). Proportion of the initial fast association (two-phase association fit, GraphPad Prism 5): 74 ± 1% (17), 29 ± 3% (18) (mean values ± SEM). (D) Dissociation of 17 and 18 from the hY4R in DPBS. The dissociation was initiated after 1.5 h of preincubation with 17 (c = 5 nM) or 18 (c = 2.5 nM) by the addition of an excess of hPP (1000-fold) and 5 (100-fold). Plateau values of the three-parameter fits (monophasic exponential decline): 13% (17), 22% (18). Data (A–D) represent mean values ± SEM from three independent experiments performed in duplicate.

The dissociation of 16 was monophasic for the sodium-free and sodium-containing buffer (Figure 5B). As also observed for the radioligand [3H]7, the dissociation of 16 from Y4R was incomplete (plateau significantly higher than zero, P < 0.05, t test). The minor component of long-lived Y4R binding may be explained by a conformational readjustment of the receptor protein upon ligand binding (47) or by an increased rebinding capability due to simultaneous interaction with more than one binding site. (48) Moreover, as 16 is a Y4R partial agonist, the incomplete dissociation could also be caused by ligand-induced receptor internalization. The Sulfo-Cy5 labeled compound 17 and the Py-1 labeled probe 18 showed a monophasic dissociation comparable to that of 16 in terms of koff values and the minor component of long-lived Y4R binding (Figure 5D and Table 4).
The kinetically derived dissociation constants Kd(kin) of 1618 obtained in DPBS were calculated from koff and kon(bi,fast), koff, and kon(bi,slow), and additionally from koff and kon(mono), as the latter represents the whole association process and also with regard to the fact that the biphasic character of the association was only weakly pronounced. The Kd(kin) values were consistently lower than the equilibrium Kd values obtained from saturation binding experiments. For all three fluorescent ligands, the lowest discrepancy between Kd(kin) and the equilibrium Kd (factor 2.6–3.2) is obtained when kon(bi,slow) is used for the calculation (Table 4). This indicates that Y4R binding of 1618 in DPBS largely proceeds according to the law of mass action for an incubation time of approximately 1 h. Flow cytometric Y4R competition binding studies in buffer I and DPBS using 16 as labeled ligand are discussed below.

Fluorescence Anisotropy-Based Y4R Binding Studies

As Cy3B exhibits a fluorescence lifetime (τ(PBS) = 2.27 ns, τ(PBS + 1% BSA) = 2.74 ns (38)) compatible with FA measurements, the Cy3B-labeled fluorescent ligand 16 was investigated in FA-based Y4R binding assays. For these experiments, Y4R displaying budded baculovirus particles (BBVs) were used. With a uniform size of approximately 300 nm × 50 nm (length × diameter), BBVs are well suited for FA measurements. (19) To note, techniques that are not based on the use of polarized light (e.g., radiochemical, flow cytometric, and NanoBRET binding assays), require a large excess of the labeled probe relative to the amount of receptor used. In contrast, FA measurements require approximately equal concentrations of the labeled ligand and receptor. Accordingly, ligand depletion must be taken into account for FA data analysis.
For Y4R binding studies, two batches of Y4R displaying BBVs were prepared in SF9 insect cells: the Y4R was either expressed alone or in combination with enzymes catalyzing a mammalian-like receptor glycosylation (so-called “SweetBac” system (49)). In the present article, the term Y4Rnonglyco is used for the former approach, precluding a mammalian-like glycosylation of Y4R, and the term Y4RSwBac is used for the latter (we do not use the term Y4Rglyco as glycosylation was not proven).
All FA binding experiments with 16 were performed in DPBS (137 mM Na+). Equilibrium binding involving two different concentrations of 16 yielded Kd values of 0.60 nM (Y4Rnonglyco) and 0.14 nM (Y4RSwBac) (Figure 6A and Table 5), essentially being in agreement with the equilibrium Kd value of 16 obtained from flow cytometric saturation binding studies in DPBS (0.70 nM). These experiments also provide information about the Y4R concentration of the BBV stocks (plotted at the abscissa in Figure 6A, values provided in the figure caption).

Figure 6

Figure 6. Binding of 16 to hY4R displaying BBVs studied by FA measurement at 27 °C. (A) Binding isotherms of 16 obtained from experiments using fixed concentrations of 16 (0.5 or 2 nM) and increasing amounts of Y4R. Total binding is represented by filled symbols and unspecific binding (determined in the presence of 1 μM hPP) is represented by open symbols. Depicted data (mean values ± SEM from a representative experiment performed in duplicate) represent snapshots at 90 min incubation. Y4R concentrations displayed on the abscissa were calculated after global analysis of the data from three or four individual experiments by a modified version of a model described by Veiksina et al., (19) affording the estimated binding site (Y4R) concentration of the applied BBV stock which amounted to 6 ± 1 nM (Y4Rnonglyco, mean value ± SEM, n = 3) or 2.1 ± 0.1 nM (Y4RSwBac, mean value ± SEM, n = 4). (B) Association and dissociation of 16 (0.5 nM) determined in real time for three different Y4R concentrations (green, blue, and red symbols). Total binding is represented by filled symbols and unspecific binding (determined in the presence of 1 μM hPP) is represented by open symbols. Data represent mean values ± SEM from a representative experiment performed in duplicate.

Table 5. Parameters Characterizing Y4R Binding of 16 Determined in Fluorescence Anisotropy-Based Binding Assays at 27 °C Using Y4Rnonglyco and Y4RSwBac Displaying BBVs
 saturation bindingbinding kinetics
BBVpKd/Kd [nM]akon [nM–1 min–1]bkoff [min–1]cKd(kin) [nM]d
Y4Rnonglyco9.2 ± 0.1/0.600.015 ± 0.0040.0012 ± 0.00010.11 ± 0.03
Y4RSwBac9.9 ± 0.1/0.140.021 ± 0.0020.00098 ± 0.000030.052 ± 0.006
a

Equilibrium dissociation constant expressed as pKd (mean values ± SEM) and Kd (mean value) obtained from three independent experiments (performed in duplicate).

b

Association rate constant ± SEM obtained from global analysis (19) of data from three individual experiments (performed in duplicate) each involving two different concentrations of 16 (0.5 and 7.0, or 0.5 and 2.0 nM).

c

Dissociation rate constants obtained from three-parameter monophasic fits (exponential decline). Mean values ± SEM from three independent experiments (performed in triplicate).

d

Kinetically derived dissociation constants ± SEM calculated from the mean koff value and individual kon values.

Association and dissociation experiments with 16 were performed with six different receptor concentrations and 2–3 different fluorescent ligand concentrations. The association and dissociation curves obtained for a ligand concentration of 0.5 nM are shown for three selected receptor concentrations in Figure 6B. For both Y4R batches, the association and dissociation curves were monophasic and the dissociation was nearly complete with plateau values <12%.
Notably, the association of 16 to Y4R was faster compared to flow cytometric and BRET-based kinetic binding studies, reaching a plateau after about 10 min (Figures 5A,C, 6B, and 7B), regardless of minor differences in assay temperatures (flow cytometry: 22 °C, FA: 27 °C, BRET: 25 °C). This could be attributed to the different expression systems (mammalian vs Sf9 insect cells). Y4Rs displayed by BBVs (FA assay) are highly unlikely to couple with insect cell Gαi-like proteins, (50) thus a ternary complex, which would affect agonist receptor affinity, cannot be formed. As in the case of flow cytometric kinetic binding studies, the kinetically derived Kd value of 16 was lower than the equilibrium Kd (ca. a factor of 6 and 3 for Y4Rnonglyco and Y4RSwBac, respectively) (Table 5). Concerning the two variants of Y4R preparations (nonglycosylated vs putatively glycosylated), no marked differences in Y4R binding were observed. FA-based Y4R competition binding studies using 16 as a labeled ligand are discussed below.

Figure 7

Figure 7. Characterization of Y4R binding of fluorescent ligand 16 in a NanoBRET-based binding assay at 25 °C using intact HEK293T-hY4R-NLuc(intraECL2) cells. (A) Representative saturation isotherm (specific binding) from saturation binding experiments. Unspecific binding was determined in the presence of 1 μM 5. Total and unspecific binding data represent mean values ± SEM. Specific binding data represent calculated values ± propagated error. (B) Association of 16 (c = 1 nM) to Y4R. Mean values ± SEM from three independent experiments performed in triplicate. (C) Dissociation of 16 from Y4R. The dissociation was initiated after 1.5 h of preincubation with 16 (c = 3.5 nM) by the addition of a 1000-fold excess of 5. Mean values ± SEM from three independent experiments performed in duplicate. Plateau value of the three-parameter fit describing a monophasic exponential decline: 6% (note: for the biphasic fit the plateau value was not different from zero, see discussion).

NanoBRET Y4R Binding Studies

To establish a Y4R NanoBRET binding assay, HEK293T cells were stably transfected with two different hY4R-Nluc fusion constructs, either with N-terminally fused Nluc (HEK293T-Nluc-hY4R cells) or with the Nluc inserted in the extracellular loop (ECL) 2 of the receptor (HEK293T-hY4R-Nluc(intraECL2) cells). Recently, BRET-based Y1R binding of a fluorescently labeled ligand at different Nluc-Y1R constructs was described. In these studies, the N-terminal fusion construct produced no BRET signal, whereas the intraECL2 construct gave high signals allowing the determination of equilibrium Kd values and kinetic data. (51) Binding experiments with 16 at HEK293T-Nluc-hY4R cells and HEK293T-hY4R-Nluc(intraECL2) cells also resulted in considerably lower BRET signals for the former compared to the latter (data not shown), which can most likely be attributed to a higher distance between the Nluc and the Y4R ligand binding pocket with the N-terminal fusion construct. Therefore, the binding studies with 16 were performed with the HEK293T-hY4R-Nluc(intraECL2) cells using coelenterazine h or furimazine as the Nluc substrate and 5 for the determination of unspecific binding instead of hPP (used to block Y4R in radiochemical, flow cytometric and FA-based assays). This was necessary as the insertion of Nluc into ECL2 hampers Y4R binding of the large endogenous ligand hPP (see competition binding studies below, Figure 8).

Figure 8

Figure 8. Determination of Y4R affinities of Y4R reference ligands (hPP, 5, 7, UR-MK188, UR-MEK388, UR-KK200) in different fluorescence-based assays by competition binding with 16. (A) Displacement curves based on data from flow cytometric competition binding experiments performed with intact CHO-hY4R-Gqi5-mtAEQ cells in sodium-free buffer (buffer I) and sodium-containing buffer (DPBS, 137 mM Na+). (B) Displacement curves from fluorescence anisotropy-based competition binding experiments performed with Y4RSwBac-displaying BBVs in DPBS. (C) Displacement curves from NanoBRET-based competition binding experiments performed with intact HEK293T-hY4R-NLuc(intraECL2) cells in L15-HEPES (140 mM Na+). Data (A–C), representing mean values ± SEM from three to five independent experiments performed in duplicate (B) or triplicate (A, C), were fitted according to a four-parameter logistic model.

Saturation binding with 16 at HEK293T-hY4R-Nluc(intraECL2) cells yielded a Kd value of 0.94 nM in good agreement with the Kd values obtained from flow cytometry and FA-based saturation binding assays (Tables 46 and Figure 7A).
Table 6. Parameters Characterizing Y4R Binding of 16 Determined in a NanoBRET Binding Assay at 25 °C
saturation bindingbinding kinetics
pKd/Kd [nM]akobs [min–1]bkoff [min–1]ckon [nM–1min–1]dKd(kin) [nM]e
9.02 ± 0.04/0.94mono: 0.18 ± 0.04mono: 0.06 ± 0.01mono: 0.18 ± 0.07mono: 0.3 ± 0.2
bi, fast: 0.58 ± 0.02bi, fast: 0.34 ± 0.03bi, fast: 0.24 ± 0.05bi, fast: 1.4 ± 0.4
bi, slow: 0.095 ± 0.005bi, slow: 0.016 ± 0.002bi, slow: 0.078 ± 0.008bi, slow: 0.21 ± 0.05
a

Equilibrium dissociation constant expressed as pKd (mean values ± SEM) and Kd (mean value) obtained from three independent experiments (performed in triplicate).

b

Observed association rate constants obtained by monophasic and biphasic fitting. Mean values ± SEM from four independent experiments (performed in triplicate).

c

Dissociation rate constants obtained from three-parameter monophasic (Y0 constrained to 100%) and five-parameter biphasic (Y0 constrained to 100%, plateau constrained to zero) fits (exponential decline). Mean values ± SEM from three independent experiments (performed in triplicate).

d

Association rate constant ± propagated error calculated from kobs, koff, and the ligand concentration used for the association studies.

e

Kinetically derived dissociation constants ± propagated error calculated from various koff and kon values.

As in the case of flow cytometric association and dissociation studies, kinetic data from the BRET assay with 16 were analyzed by monophasic and biphasic fits (Figure 7B,C and Table 6). For both, association and dissociation experiments, the statistical Extra sum-of-squares F-test suggested that the data are better fitted by the biphasic fit compared to a monophasic analysis (two-phase association vs one-phase association and two-phase decay vs one-phase decay, GraphPad Prism 5) giving P values <0.0001. For the monophasic dissociation analysis, the plateau value was low (6%), but still significantly different from zero (P < 0.05, t test). In contrast, for the biphasic fit, the plateau value was not different from zero (P > 0.05, t test) indicating a complete dissociation of 16 from the Y4R (Figure 7B,C). Possible reasons for a biphasic association were discussed in the section describing flow cytometric binding studies (see above). Like the biphasic association, the biphasic dissociation may also be explained by the existence of different receptor states. (47) However, since a monophasic dissociation curve was obtained from flow cytometric dissociation experiments with 16 (cf. Figure 5B), the cell type (flow cytometry: CHO cells, NanoBRET: HEK293T cells) and/or the receptor variant (flow cytometry: wild-type Y4R, NanoBRET: Y4R-Nluc construct (Nluc inserted in ECL2)) seem to be relevant for the association. Recently, a metadynamics protocol for binding and unbinding of peptide ligands of class A GPCRs suggested that Y4R agonists also bind with high affinity to the vestibule-like binding site of Y4R, located on top of the orthosteric binding site. (52) Consequently, in the dissociation process, the ligand might be withheld by residing in the vestibule of the receptor. Depending on the receptor conformation and possibly additionally triggered by the presence of Nluc in the ECL2 of the Y4R, the residence time in the vestibule might vary, resulting in different dissociation rates as observed in the NanoBRET assay.
The Kd(kin) values of 16 were calculated from kon(bi,fast) and koff(bi,fast) as well as from kon(bi,slow) and koff(bi,slow), and additionally from kon(mono) and koff(mono) as the monophasic fits cover the whole association and dissociation process. Moreover, the biphasic character was generally weak (minor differences in k values, see Table 6). Interestingly, the Kd(kin) value derived from kon(bi,fast) and koff(bi,fast) was in good agreement with the Kd value from saturation binding studies. The Kd(kin) value obtained from kon(mono) and koff(mono) and the Kd(kin) calculated from kon(bi,slow) and koff(bi,slow) were similar and slightly lower (factor 3–5) compared to the Kd obtained from saturation binding experiments (Table 6). Consequently, irrespective of the weak biphasic nature of the association and dissociation, Y4R binding of 16 studied in the NanoBRET assay approximatively follows the law of mass action.
BRET-based saturation binding experiments at HEK293T-hY4R-Nluc(intraECL2) cells were also performed with the Py-1 labeled ligand 18. Although the excitation spectrum of the Py-1 conjugate 18 shows a higher overlap with the Nluc emission spectrum (λmax ca. 460 nm) (53) compared to the excitation spectrum of the Cy3B-labeled ligand 16 (cf. Table 3), considerably lower BRET signals compared to experiments carried out with 16 (data not shown) were obtained, which can be explained by the markedly higher fluorescence quantum yield of 16 compared to 18 (see Table 3). NanoBRET Y4R competition binding studies using 16 as labeled ligand are discussed below.

Fluorescence-Based Y4R Competition Binding

Using fluorescent ligand 16 as labeled probe, Y4R affinities of previously reported Y4R ligands (hPP, 5, 7 and UR-MEK388 or UR-MK188) were determined in flow cytometric, FA- and BRET-based binding assays (displacement curves shown in Figure 8). To note, the dimeric argininamide-type Y4R ligands UR-MEK388 and UR-MK188 represent enantiomers ((S,S)- and (R,R)-configuration, respectively) displaying almost equal Y4R affinity and Y4R antagonism (Ca2+ aequorin assay). (54) Additionally, the Y4R partial agonist UR-KK200 (35) was studied in the NanoBRET assay.
The obtained Y4R affinities of hPP were consistently lower than affinities determined in radiochemical assays (Table 7), with the lowest discrepancy (ca. one log unit) found in the case of the flow cytometric binding studies performed in DPBS. The lowest hPP affinity was determined in the NanoBRET-based competition binding assay, producing a shallow curve with a very low hill slope of −0.3 (Figure 8C). This shows that the insertion of Nluc into ECL2 affects the binding of hPP to Y4R. Similar to hPP, the obtained pKi values of the cyclic peptides 5 and 7 were lower by approximately 1 order of magnitude compared to reported pKi values determined in radioligand competition binding assays. The Y4R affinity of UR-KK200, determined in the NanoBRET assay, was in good agreement with the reported data (radiochemical assay). Y4R binding data, obtained for UR-MEK388 (flow cytometric and NanoBRET assays) were also in accordance with a reported pKi value determined in a flow cytometric competition binding assay using Sulfo-Cy5-labeled [K4]hPP as labeled ligand (Table 7). The Y4R affinity of UR-MK188, determined in the FA competition binding assay, was slightly higher compared to the reported binding data of UR-MK188. Generally, the lowest discrepancies between the determined and reported binding data of Y4R reference ligands were found for the NanoBRET assay. As 16 showed a complete dissociation from the Y4R only in the case of the NanoBRET assay, the incomplete dissociation observed in the flow cytometric and FA binding assay could be a reason for the observed discrepancies between reported Y4R affinities and binding data determined with 16 in flow cytometric and FA-based competition binding assays: to achieve a complete “displacement” of the fluorescent ligand 16, the competing ligand, incapable of displacing (pseudo)irreversibly bound 16 from the receptor, must be used at higher concentrations (compared to the situation where the labeled ligand shows full reversible binding), resulting in a higher receptor occupancy and thus effectively preventing binding of 16 to the receptor. This phenomenon was also clearly observed for radiolabeled muscarinic M2 receptor ligands showing different dissociation kinetics (reversible and (pseudo)irreversible binding): competition binding studies with reference ligands yielded consistently lower pKi values for the nonreversibly binding radioligand. (55)
Table 7. Overview of hY4R Binding Affinities of Y4R Reference Ligands Determined in Fluorescence-Based Competition Binding Assays Using 16 as Labeled Probe
 flow cytometryaFAbnanoBRETcliterature
 Na+ free137 mM Na+137 mM Na+140 mM Na+Na+ free137 mM Na+
cmpd.pKi
hPP8.5 ± 0.18.9 ± 0.18.0 ± 0.27.7 ± 0.110.02d10.1e
59.49 ± 0.079.15 ± 0.089.26 ± 0.048.89 ± 0.0610.48fn.a.
79.61 ± 0.068.8 ± 0.19.4 ± 0.19.07 ± 0.0710.5e9.79e
UR-KK200n.d.n.d.n.d.7.8 ± 0.18.92g7.99e
UR-MEK3886.29 ± 0.086.68 ± 0.09n.d.6.85 ± 0.056.58hn.a.
UR-MK188n.d.n.d.7.7 ± 0.1n.d.6.88h6.82g
6.18g
a

Determined at intact CHO-hY4R-Gqi5-mtAEQ cells; mean values ± SEM from three or four independent experiments performed in triplicate.

b

Determined at Y4RSwBac displaying BBVs; mean values ± SEM from three independent experiments performed in duplicate.

c

Determined at intact HEK293T-hY4R-NLuc(intraECL2) cells; mean values ± SEM from three or four independent experiments performed in triplicate.

d

Wirth et al. (37)

e

Gleixner et al. (31)

f

Konieczny et al. (30)

g

Kuhn et al. (reported Ki values were converted to pKi). (35)

h

Keller et al. (reported Ki values were converted to pKi). (54) Note: UR-MEK388 and UR-MK188 represent enantiomers exhibiting almost equal Y4R affinity. (54) n.a. not available.

Microscopy Studies

Y4R binding of 16 and 17 was visualized by confocal microscopy using intact CHO-hY4R-Gqi5-mtAEQ cells. Cellular uptake of both fluorescent probes was evident after already 10 min of incubation at 22 °C (Figure 9). After 30 min, the intracellular and plasma membrane-associated fluorescence was higher compared to 10 min after start of incubation, which is in agreement with the association kinetics determined by flow cytometry (Figure 5A). Notably, in the presence of an excess of hPP, almost no ligand′s fluorescence could be detected (unspecific binding, Figure 9). This indicates that the cellular uptake of the fluorescent ligands, which appear to be located in vesicles, is caused by Y4R-mediated endocytosis. Similar results were previously observed with the Sulfo-Cy5 labeled hPP derivative 1 studied with the same cells. (26)

Figure 9

Figure 9. Visualization of fluorescent ligand (16, 17) binding to CHO-hY4R-Gqi5-mtAEQ cells by confocal microscopy. Shown are representative images acquired after incubation of the cells with 16 or 17 (each 20 nM) at 22 °C for 10 and 30 min. Unspecific binding was determined in the presence of 1 μM hPP. Nuclei were stained with H33342 (2 μM). Fluorescence of 16 and 17 is shown in green and red, respectively. Fluorescence of H33342 is shown in blue. Scale bar: 10 μm.

Y4R binding of the Cy3B-labeled ligand 16 was also studied at transiently transfected SK-OV-3 cells using wide-field and TIRF microscopy. For these experiments, a lower concentration of 16 (1 nM), compared to confocal microscopy (20 nM), was used, resulting in a lower receptor occupancy. As also observed for binding of 16 to CHO-hY4R-Gqi5-mtAEQ cells (confocal microscopy), unspecific binding was very low (Figure 10A).

Figure 10

Figure 10. Visualization of binding of 16 to the hY4R transiently expressed by SK-OV-3 cells using wide-field and TIRF microscopy. (A) Wide-field fluorescence images acquired after incubation of the cells with 16 at 37 °C for 30 min. The two-color composite of individual focal planes after Z-stack deconvolution is shown with the green pseudocolor for 16 (561 nm excitation) and the blue pseudocolor for the nuclear stain channel (Hoechst 34580, 405 nm excitation). (B) Wide-field fluorescence and TIRF images of the same cells obtained after incubation of the cells with 16 (1 nM) at 37 °C for 30 min. Wide-field images were processed as under (A). In TIRF images, fluorescence of 16 is shown in white pseudocolor. Scale bar: 10 μm.

As an additional control, binding of 16 to nontransfected SK-OV-3 cells was investigated, also revealing very low unspecific binding. Thus, the fluorescence of 16 detected in the total binding samples represents ligand 16 bound to Y4R. The Y4R appeared to be largely associated with the plasma membrane of the cells (Figure 10A).
Using a higher magnification (60× objective), intracellular fluorescence could be detected (Figure 10B). In the TIRF images, also binding of the ligand to Y4R located in plasma membrane protrusions (filopodia) was visible. Notably, fluorescent ligand 16 proved to be sufficiently photostable and bright enough for single-molecule tracking studies by TIRF. This was demonstrated by binding of 16 to Y4R expressing SK-OV-3 cells using a low fluorescent ligand concentration of 0.1 nM (time lapse TIRF imaging sequence recorded at 30 Hz resolution; video, shown at double speed, is available as Supporting Information). Various diffusion modes could be observed along the particle tracks. Here, events of continuous confinement as well as transient confinement (when molecules diffused in a grid of impermeable barriers) were identified. The unconfined trajectories can be classified as having mostly random diffusion, except for the filopodia where directional diffusion with high linearity was observed. A deeper analysis of the trajectories is beyond the scope of this article.

Conclusions

Click to copy section linkSection link copied!

Conjugation of different fluorescent dyes to recently discovered small cyclic peptides, exhibiting high Y4R binding and selectivity, produced fluorescent Y4R ligands with unprecedented Y4R affinity (Ki, Kd < 1 nM). Notably, the cyclic peptides contained one or two Nω-carbamoylated arginines. These modified arginines were recently introduced as bioisosteric replacements of natural arginine in arginine-containing peptides enabling radio- and fluorescence labeling of bioactive peptides via arginine residues. (32,35,39,56,57) The present work further demonstrates the usefulness of Nω-carbamoylated arginines with respect to the design and preparation of labeled probes for peptidergic receptors. The study also suggests that a broad variety of fluorescent dyes and other bulky moieties can be attached to the cyclic peptides without marked impact on Y4R binding. Therefore, further tool compounds with high Y4R affinity and selectivity are conveniently accessible according to the presented strategy. The preparation of new fluorescent ligands might be of interest, for example, in improving the physicochemical properties (use of more polar fluorescent dyes) or an adjustment of the photophysical properties for the intended application. Nonetheless, the presented fluorescent probes are useful for the determination of Y4R binding affinities in fluorescence-based competition binding assays and, thus, could be used as tools in drug development programs aiming for therapeutics acting at Y4R.

Experimental Section

Click to copy section linkSection link copied!

Materials

The protected amino acids Fmoc-Tyr(tBu)–OH, Fmoc-Leu−OH, and Fmoc-Arg(Pbf)–OH were purchased from Carbolution Chemicals (St. Ingbert, Germany). HBTU, Fmoc-Sieber PS Resin, and PyBOP were obtained from Iris Biotech (Marktredwitz, Germany). Copper(II)sulfate pentahydrate, HOBt, sodium ascorbate, and Triton X-100 were from Sigma-Aldrich (Taufkirchen, Germany). NMP and DMF for peptide synthesis, anhydrous DMF, dichloromethane, piperidine, succinic anhydride, trifluoroacetic acid (TFA), and tetrahydrofuran were purchased from ACROS/Fisher Scientific (Schwerte, Germany). DIPEA was obtained from TCI (Eschborn, Germany). tert-Butyl (3-bromopropyl)carbamate (20) was from ABCR (Karlsruhe, Germany). Acetonitrile (high-performance liquid chromatography (HPLC) gradient grade) was from VWR (Ismaning, Germany). The fluorescent dye 13 (S0586, succinimidyl ester of S0223) was from FEW chemicals (Wolfen, Germany). Human pancreatic polypeptide (hPP) and porcine neuropeptide Y (pNPY) were purchased from Synpeptide (Shanghai, China). Bacitracin and bovine serum albumin (BSA) were obtained from Serva (Heidelberg, Germany). The Nluc substrate furimazine was purchased from Promega (Walldorf, Germany). Coelenterazine h was from BIOTREND (Köln, Germany). Zeocine and Puromycine were from Invivogen (San Diego, CA), Hygromycin B and propidium iodide were from Carl Roth (Karlsruhe, Germany) and Geneticine G418 was from Fisher Scientific (Schwerte, Germany). The Y2R antagonist BIIE0246 was obtained from Boehringer Ingelheim (Ingelheim, Germany). The syntheses of the Y1R antagonists BIBO3304 (41) and [3H]UR-MK299 (molar activity: 1.81 TBq/mmol), (33) and the synthesis of the Y4R ligands 4, (30) 5, (30) 6, (31) 7, (31) and [3H]UR-KK200 (molar activity: 0.98 TBq/mmol) (35) were described previously. [3H]Propionyl-pNPY (molar activity: 1.39 TBq/mmol) was prepared according to a previously reported procedure (33) with minor modifications that were described elsewhere. (40) Compounds 8 (32) and 9, (29) the fluorescent dyes 12 (Cy3B-SE), (38) 14 (Py-1), (58) and 22 (Py-5), (58) as well as the fluorescent dummy ligands 23 (38) and 24 (39) were prepared according to reported procedures. Millipore water was consistently used for the preparation of stock solutions, buffers, and aqueous eluents for HPLC. Polypropylene reaction vessels (1.5 and 2 mL) from Sarstedt (Nümbrecht, Germany) were used to keep stock solutions and for small-scale reactions (e.g., fluorescence labeling reactions and activation of Fmoc-protected amino acids for peptide synthesis).

NMR Spectroscopy

NMR spectra were recorded on a Bruker AVANCE 400 (1H, 400 MHz; 13C, 100 MHz) or on an AVANCE 600 instrument with cryogenic probe (1H: 600 MHz and 13C: 150 MHz) (Bruker, Karlsruhe, Germany). NMR spectra were calibrated based on the solvent residual peaks (1H NMR, DMSO-d6: δ = 2.50 ppm; 13C NMR: DMSO-d6: δ = 39.50 ppm), and data are reported as follows: 1H NMR: chemical shift δ in ppm (multiplicity [s = singlet, d = doublet, t = triplet, m = multiplet, and br s = broad singlet], integral, coupling constant J in Hz).

Mass Spectrometry

High-resolution mass spectrometry (HRMS) was performed with an Agilent 6540 UHD accurate-mass quadrupole time-of-flight (Q-TOF) liquid chromatography/mass spectrometry (LC/MS) system coupled to an Agilent 1290 analytical HPLC system (Agilent Technologies. Santa Clara, CA) using an ESI source and the following LC method: column: Luna Omega C18, 1.6 μm, 50 mm × 2.1 mm (Phenomenex, Aschaffenburg, Germany), column temperature: 40 °C, solvent/linear gradient: 0–4 min: 0.1% aqueous HCOOH/acetonitrile supplemented with 0.1% HCOOH 95:5–2:98, 4–5 min: 2:98, flow: 0.6 mL/min.

Preparative HPLC

Preparative HPLC was performed with a system from Knauer (Berlin, Germany) consisting of two K-1800 pumps and a K-2001 detector (in the following referred to as “system 1”) or with a Prep 150 LC system from Waters (Eschborn, Germany) comprising a Waters 2545 binary gradient module, a Waters 2489 UV/vis detector, and a Waters fraction collector III (in the following referred to as “system 2”). A Gemini NX-C18, 5 μm, 250 mm × 21 mm (Phenomenex) was used as a reversed-phase (RP) column at a flow rate of 20 mL/min using mixtures of 0.1% aqueous TFA and acetonitrile as the mobile phase. A detection wavelength of 220 nm was used throughout. Collected fractions were lyophilized using a Scanvac CoolSafe 100–9 freeze-dryer (Labogene, Allerød, Denmark) equipped with an RZ 6 rotary vane vacuum pump (Vacuubrand, Wertheim, Germany).

Analytical HPLC

Analytical HPLC analysis was performed with a system from Agilent Technologies composed of a 1290 Infinity binary pump equipped with a degasser, a 1290 Infinity autosampler, a 1290 Infinity thermostated column compartment, a 1260 Infinity diode array detector, and a 1260 Infinity fluorescence detector. A Kinetex-XB C18, 2.6 μm, 100 mm × 3 mm (Phenomenex) served as the stationary phase at a flow rate of 0.6 mL/min. Detection was performed at 220 nm, and the oven temperature was 25 °C. Mixtures of 0.04% aqueous TFA (A) and acetonitrile (B) were used as mobile phase. The following linear gradients were applied: for compounds 11 and 19: 0–14 min: A/B 90:10–70:30, 14–16 min: 70:30–5:95, 16–20 min: 5:95 (isocratic); for compounds 1618: 0–14 min: A/B 90:10–65:35, 14–16 min: 65:35–5:95, 16–20 min: 5:95 (isocratic). The injection volume was 20 μL. Retention (capacity) factors k were calculated from the retention times tR according to k = (tRt0)/t0 (t0 = dead time, 0.76 min for the system, column, and flow rate mentioned above).

Compound Characterization

The linear precursor peptide 10, synthesized by solid-phase peptide synthesis, was characterized by HRMS. Peptide 11 was characterized by HRMS, 1H NMR spectroscopy, and RP-HPLC. The azido-functionalized fluorescent dye 15 was characterized by HRMS and 1H NMR spectroscopy. The fluorescent ligands 1619 were characterized by HRMS and RP-HPLC. HPLC purities of all target compounds were ≥95% (UV detection, 220 nm).

Synthesis of Peptide 11 and the Fluorescent Ligands 1619

Nα-Succinyl-Nω[hex-5-ynylaminocarbonyl]-Arg-Tyr-Nω-[(aminobutyl)aminocarbonyl] Arg-Leu-Arg-Tyr-Amide Tetrakis(hydrotrifluoroacetate) (10)
Peptide 10 was synthesized on a Fmoc-Sieber PS resin (100 mg, 0.59 mmol/g) by manual Fmoc strategy SPPS using DMF/NMP (both anhydrous) 4:1 v/v as solvent. 5 mL NORM-JECT syringes (B. Braun Melsungen, Melsungen, Germany), equipped with a 35 μm polypropylene frit (Roland Vetter Laborbedarf, Ammerbuch, Germany), were used as reaction vessels. The resin was allowed to swell in solvent for 45 min at room temperature followed by initial Fmoc deprotection using a 20% piperidine solution in DMF/NMP 4:1 v/v (2 × 20 min at rt). The contained standard Fmoc amino acids (Tyr, Arg, Leu) were used in 5-fold excess and preactivated with HOBt/HBTU/DIPEA (5/4.9/10 equiv) in solvent (about 2.2 mL/mmol aa) for at least 5 min before addition to the resin. For the activation of building blocks 8 and 9, used in 3-fold excess, HBTU/HOBt/DIPEA (3/3/6 equiv) were used. Amino acid coupling was carried out on a shaker (Heidolph Multi Reax; Heidolph Instruments, Schwabach, Germany) covered with a thermostat-controlled (38 °C) box. In the case of standard (proteinogenic) amino acids, “double” coupling (2 × 45 min) was performed. Arginine derivatives 8 and 9 were attached by a single coupling procedure (16 h). After coupling, Fmoc deprotection was carried out using a 20% piperidine solution in DMF/NMP 4:1 v/v for 2 × 10 min at rt. The resin was washed 6 times with about 1 mL of solvent after Fmoc deprotection and 4 × 1 mL after amino acid coupling. After coupling of the last amino acid and subsequent Fmoc deprotection, the resin was treated with a solution of succinic anhydride (52.3 mg, 0.443 mmol) and DIPEA (77.1 μL, 0.443 mmol) in DMF/NMP 80:20 v/v (500 μL) at 35 °C for 30 min. The resin was washed with DMF/NMP 4:1 v/v (6×) and CH2Cl2 (3×), followed by cleavage off the resin. For this, a mixture of trifluoroacetic acid (TFA) and DCM (3:1 v/v) was added to the resin and the mixture was shaken for 20 min. The fluid was collected in a 100 mL round-bottom flask, and the procedure was repeated once. The resin was afterward washed 3 times with cleavage solution, and the volatiles were removed in vacuo. Full side-chain deprotection was carried out with 95% aq. TFA (2 mL) for 5 h at rt. The reaction mixture was concentrated in vacuo, then water (40 mL) was given to the mixture, followed by lyophilization. Purification by preparative HPLC (system 1, gradient: 0–18 min: 0.1% aq TFA/acetonitrile 95:5–55:45, tR = 15 min) yielded 10 as a white fluffy solid (37.65 mg, 37%). HRMS (ESI): m/z [M + 2H]2+ calcd. for [C58H93N19O13]2+ 631.8595, found: 631.8616. C58H91N19O13·C8H4F12O8 (1262.49 + 456.09).
(9S,12S,15S,E)-4-Amino-N-{(S)-1-[((S)-1-{[(S)-1-amino-3-(4-hydroxyphenyl)-1-oxopropan-2-yl]amino}-5-guanidino-1-oxopentan-2-yl)amino]-4-methyl-1-oxopentan-2-yl}-15-{3-[(E)-2-(hex-5-yn-1-ylcarbamoyl)guanidino]propyl}-12-(4-hydroxybenzyl)-2,11,14,17,20-pentaoxo-1,3,5,10,13,16,21-heptaazacyclopentacos-3-ene-9-carboxamide Tris(hydrotrifluoroacetate) (11)
Peptide 10 (8.89 mg, 7.13 μmol) was dissolved in anhydrous DMF (5.7 mL), a solution of HOBt (2.89 mg, 21.4 μmol) in DMF (0.6 mL) and DIPEA (7.5 μL, 42.8 μmol) were added and the mixture was stirred at rt for 5 min. A solution of PyBOP (11.1 mg, 21.4 μmol) in DMF (0.8 mL) was added dropwise and stirring was continued at rt for 16 h. 1% aqueous TFA (10 mL) was added and the product was purified by preparative HPLC (system 1, gradient: 0–18 min: 0.1% aq TFA/acetonitrile 95:5–55:45, tR = 16 min) to yield 11 as a white fluffy solid (2.39 mg, 21%) 1H NMR (600 MHz, DMSO-d6): δ (ppm) 0.82–0.91 (m, 6H), 1.33–1.49 (m, 12H), 1.49–1.59 (m, 7H), 1.59–1.70 (m, 3H), 1.70–1.79 (m, 1H), 2.15–2.2 (m, 2H), 2.29–2.41 (m, 3H), 2.68–2.74 (m, 1H), 2.74–2.81 (m, 2H), 2.81–2.89 (m, 2H), 2.89–2.97 (m, 1H), 3.02–3.08 (m, 3H), 3.09–3.14 (m, 3H), 3.17–3.21 (m, 3H), 3.28–3.32 (m, 3H), 4.05–4.12 (m, 1H), 4.16–4.23 (m, 1H), 4.25–4.42 (m, 4H), 6.44–6.59 (br s, 1H), 6.61–6.68 (m, 4H), 6.68–6.96 (br s, 2H), 6.95–7.03 (m, 4H), 7.04–7.07 (m, 1H), 7.09–7.48 (br s, 2H, interfering with the next listed signal), 7.35 (s, 1H), 7.50 (t, 1H, J 5.4 Hz), 7.54–7.62 (m, 2H), 7.75 (d, 1H, J 7.9 Hz), 7.83 (d, 1H, J 7.5 Hz), 7.87–7.94 (m, 2H), 7.97 (d, 1H, J 7.3 Hz), 8.05–8.11 (m, 1H), 8.11–8.21 (m, 1H), 8.23–8.6 (m, 3H), 8.95 (s, 1H), 9.09–9.25 (m, 3H), 9.85 (s, 1H), 10.0 (s, 1H). HRMS (ESI): m/z [M + 3H]3+ calcd. for [C58H92N19O12]3+ 415.5719, found: 415.5725 RP-HPLC (220 nm): 98% (tR = 10.7 min, k = 13.0). C58H89N19O12·C6H3F9O6 (1244.47 + 342.07).
24-{5-[(4-{[(9S,12S,15S,19R,E)-4-Amino-9-({(S)-1-[((S)-1-{[(S)-1-amino-3-(4-hydroxyphenyl)-1-oxopropan-2-yl]amino}-5-guanidino-1-oxopentan-2-yl)amino]-4-methyl-1-oxopentan-2-yl}carbamoyl)-15-(3-guanidinopropyl)-12-(4-hydroxybenzyl)-2,11,14,17,20-pentaoxo-1,3,5,10,13,16,21-heptaazacyclopentacos-3-en-19-yl]amino}-4-oxobutyl)amino]-5-oxopent-1-yn-1-yl}-5,5,27,27-tetramethyl-16-oxa-20-aza-12-azoniaheptacyclo[15.11.0.03,15.04,12.06,11.020,28.021,26]octacosa-1(28),2,4(12),6(11),7,9,21(26),22,24-nonaene-8-sulfonate Bis(hydrotrifluoroacetate) (16)
Peptide 6 (1.05 mg, 0.626 μmol) was dissolved in anhydrous DMF (100 μL) followed by the addition of DIPEA (0.88 μL, 5.00 μmol). A solution of 12 (0.49 mg, 0.626 μmol) in DMF (64 μL) was added, and the mixture was shaken at rt in the dark for 1.5 h. 100 μL of 1% aqueous TFA were added and the mixture was subjected to preparative HPLC (system 1, gradient: 0–20 min: 0.1% aq TFA/acetonitrile 90:10–50:50, tR = 11 min) to yield 16 as a purple fluffy solid (0.40 mg, 35%). HRMS (ESI): m/z [M + 2H]2+ calcd. for [C89H122N22O17S]2+ 901.9548, found: 901.9555 RP-HPLC (220 nm): 99% (tR = 10.4 min, k = 12.7). C89H120N22O17S·C4H2F6O4 (1802.14 + 228.04).
4-(2-{(1E,3E)-5-[(E)-1-(6-{[4-({(9S,12S,15S,19R,E)-4-Amino-9-[((S)-1-{[(S)-1-({(S)-1-amino-3-(4-hydroxyphenyl)-1-oxopropan-2-yl}amino)-5-guanidino-1-oxopentan-2-yl]amino}-4-methyl-1-oxopentan-2-yl)carbamoyl]-15-(3-guanidinopropyl)-12-(4-hydroxybenzyl)-2,11,14,17,20-pentaoxo-1,3,5,10,13,16,21-heptaazacyclopentacos-3-en-19-yl}amino)-4-oxobutyl]amino}-6-oxohexyl)-3,3-dimethyl-5-sulfoindolin-2-ylidene]penta-1,3-dien-1-yl}-3,3-dimethyl-3H-indol-1-ium-1-yl)butane-1-sulfonate Bis(hydrotrifluoroacetate) (17)
Peptide 6 (1.05 mg, 0.626 μmol) was dissolved in anhydrous DMF (100 μL) followed by the addition of DIPEA (0.88 μL, 5.00 μmol). A solution of 13 (0.50 mg, 0.626 μmol) in DMF (100 μL) was added, and the mixture was shaken at rt in the dark for 1.5 h. 100 μL of 1% aqueous TFA were added and the mixture was subjected to preparative HPLC (system 1, gradient: 0–20 min: 0.1% aq TFA/acetonitrile 90:10–50:50, tR = 12 min) affording 17 as a blue fluffy solid (0.82 mg, 70%). HRMS (ESI): m/z [M + 2H]2+ calcd. for [C90H132N22O19S2]2+ 944.4735, found: 944.4747. RP-HPLC (220 nm): 96% (tR = 11.6 min, k = 14.3). C90H130N22O19S2·C4H2F6O4 (1888.29 + 228.04).
1-{4-[((9S,12S,15S,19R,E)-4-Amino-9-{[(S)-1-({(S)-1-[((S)-1-amino-3-(4-hydroxyphenyl)-1-oxopropan-2-yl)amino]-5-guanidino-1-oxopentan-2-yl}amino)-4-methyl-1-oxopentan-2-yl]carbamoyl}-15-(3-guanidinopropyl)-12-(4-hydroxybenzyl)-2,11,14,17,20-pentaoxo-1,3,5,10,13,16,21-heptaazacyclopentacos-3-en-19-yl)amino]-4-oxobutyl}-2,6-dimethyl-4-[(E)-2-(2,3,6,7-tetrahydro-1H,5H-pyrido[3,2,1-ij]quinolin-9-yl)vinyl]pyridin-1-ium Bis(hydrotrifluoroacetate) Trifluoroacetate (18)
Peptide 6 (1.63 mg, 0.97 μmol) was dissolved in anhydrous DMF (150 μL) followed by the addition of DIPEA (1.1 μL, 6.31 μmol), the addition of a solution of 14 (0.47 mg, 1.16 μmol) in DMF (15 μL) and shaking of the mixture at rt in the dark for 2 h. 100 μL of 1% aqueous TFA were added, and the mixture was subjected to preparative HPLC (system 1, gradient: 0–20 min: 0.1% aq TFA/acetonitrile 90:10–50:50, tR = 14 min) to yield 18 as a red fluffy solid (0.61 mg, 34%). HRMS (ESI): m/z [M+ + 3H]4+ calcd. for [C76H113N21O12]4+ 377.9714, found: 377.9727. RP-HPLC (220 nm): 97% (tR = 13.1 min, k = 16.2). C76H110N21O12+·C6H2F9O6 (1508.86 + 341.06).
1-{3-[4-(4-{3-[(E)-Amino({3-[(9S,12S,15S,E)-4-amino-9-({(S)-1-[((S)-1-{[(S)-1-amino-3-(4-hydroxyphenyl)-1-oxopropan-2-yl]amino}-5-guanidino-1-oxopentan-2-yl)amino]-4-methyl-1-oxopentan-2-yl}carbamoyl)-12-(4-hydroxybenzyl)-2,11,14,17,20-pentaoxo-1,3,5,10,13,16,21-heptaazacyclopentacos-3-en-15-yl]propyl}amino)methylene]ureido}butyl)-1H-1,2,3-triazol-1-yl]propyl}-4-{(1E,3E)-4-[4-(dimethylamino)phenyl]buta-1,3-dien-1-yl}-2,6-dimethylpyridin-1-ium Bis(hydrotrifluoroacetate) Trifluoroacetate (19)
Peptide 11 (0.84 mg, 0.53 μmol) was dissolved in H2O (190 μL). A solution of copper(II)sulfate pentahydrate (0.2 mg, 0.64 μmol) in H2O (30 μL), a solution of sodium ascorbate (0.36 mg, 1.59 μmol) in H2O (30 μL) and a solution of 15 (0.25 mg, 0.53 μmol) in NMP (70 μL) were added followed by the addition of 180 μL of NMP to adjust the ratio of the solvents (H2O/NMP) to 1:1 v/v (final volume: ca. 500 μL). The mixture was shaken at rt in the dark for 2 h. 100 μL of 1% aqueous TFA were added, and the mixture was subjected to preparative HPLC (system 1, gradient: 0–20 min: 0.1% aq TFA/acetonitrile 80:20–50:50, tR = 9 min) to yield 19 as a red fluffy solid (0.29 mg, 27%). HRMS (ESI): m/z [M+ + 3H]4+ calcd. for [C80H120N24O12]4+ 402.2374, found: 402.2380. RP-HPLC (220 nm): 95% (tR = 13.7 min, k = 17.0). C80H117N24O12+·C6H2F9O6 (1606.97 + 341.06).

Chemical Stability

The chemical stability of the fluorescent ligands 16, 17, and 18 was investigated in PBS (adjusted to pH 7.4) at 22 °C. The incubation was started by the addition of 7.5 μL of a 2 mM stock solution (solvent: DMSO/H2O 1:1 v/v) to 142.5 μL of PBS (in the case of compound 18 supplemented with 10% DMSO due to solubility reasons) to yield a concentration of 100 μM. After time periods of 0, 6, and 48 h, an aliquot (25 μL) was withdrawn and added to 25 μL of acetonitrile/0.04% aq. TFA 1:9 v/v to obtain a peptide solution with a concentration of 50 μM. 20 μL of this solution were subjected to analytical RP-HPLC analysis using the same system and conditions as described under Analytical HPLC, but applying a different linear gradient: 0–14 min: A/B 90:10–45:55, 14–15 min: 45:55–5:95, 15–19 min: 5:95 (isocratic). A 1:1 mixture of PBS and acetonitrile/0.04% aq. TFA 1:9 v/v (20 μL) was analyzed to obtain the blank chromatogram.

Fluorescence Excitation and Emission Spectra

Excitation and emission spectra were recorded with a Cary Eclipse spectrofluorimeter (Varian, Mulgrave, Victoria, Australia) using polystyrene cuvettes (10 mm × 10 mm, reference 1961, Kartell, Noviglio, Italy). Sample solutions (2 mL) were prepared in the cuvettes. Spectra in PBS (pH 7.4) and PBS supplemented with 1% BSA were obtained for all fluorescent ligands (1619) using the following concentrations: 6 μM for 19, 5 μM for 18 (in PBS), and 1 μM for 16, 17, and 18 (in PBS with 1% BSA). Excitation spectra (cf. Figure S7) were recorded with spectral bandwidths of 5 nm (excitation slit) and 10 nm (emission slit). The spectral bandwidth applied for the emission spectra was 10 nm (excitation slit) and 5 nm (emission slit). Net spectra were obtained by subtracting the respective reference spectrum of a vehicle sample.

Determination of Fluorescence Quantum Yields

The fluorescence quantum yield of compound 19 was determined following a reported procedure using the aforementioned Cary Eclipse spectrofluorimeter for the measurement of emission spectra and a Lambda 650 UV/vis spectrophotometer (PerkinElmer, Waltham, MA) for the measurement of absorption spectra. (41) The concentration of 19 was 6 μM and cresyl violet perchlorate (Acros Organics, Geel, Belgium) was used as reference compound (concentration: 2 μM). The excitation wavelength was set close to the absorption maximum (500 nm for 19 in PBS, 515 nm for 19 in PBS with 1% BSA). Emission spectra of cresyl violet perchlorate in EtOH were recorded at an excitation wavelength of 575 nm. Sample solutions (2 mL) were prepared in polystyrene cuvettes (10 mm × 10 mm), immediately followed by measurement of the emission spectra, transfer of the solutions into acryl cuvettes (10 mm × 10 mm, reference 67.755, Sarstedt), and measurement of the absorption spectra. Emission spectra were recorded for the spectral bandwidth settings (excitation/emission) of 10/5 and 10/10 nm. The obtained quantum yields were averaged.
The fluorescence quantum yield of compounds 1618 was determined via an absolute method using an Ulbricht sphere with an inaccuracy of ca. 3% (Hamamatsu C9920–02 system equipped with a Spectralon integrating sphere) at room temperature (23 ± 1 °C). The optical density at the excitation wavelength of the sample was <0.1 (optical path length: 1 cm). Samples were measured in a 10 mm × 10 mm quartz cuvette.

Cell Culture

Cells were cultured in T75 or T175 tissue culture flasks (Sarstedt, Nümbrecht, Germany) in a humidified atmosphere (95% air, 5% CO2) at 37 °C. SK-N-MC neuroblastoma cells (obtained from the American Type Culture Collection, ATCC HTB-10) were maintained in EMEM (Sigma) supplemented with 5% FBS. CHO-hY2R cells (obtained from PerkinElmer, Rodgau, Germany) were cultured in Ham’s F-12 supplemented with 5% FBS and G418 (400 μg/mL). CHO-hY4-Gqi5-mtAEQ cells (25) were cultured in HAM’s F-12 (Sigma) supplemented with 10% FBS, hygromycin (400 μg/mL), zeocin (250 μg/mL) and G418 (400 μg/mL). HEC-1B-hY5R cells (59) were maintained in EMEM supplemented with 5% FBS and G418 (400 μg/mL). HEK293T-NlucN-mGsi/Y4R-NlucC cells (37) were cultured in DMEM (Sigma) supplemented with 10% FBS, puromycin (1 μg/mL), and G418 (600 μg/mL). HEK293T-CAMYEN-hY4R cells (31) were cultured in DMEM supplemented with 10% FBS, zeocin (100 μg/mL), and G418 (600 μg/mL). Human SK-OV-3 ovarian adenocarcinoma cells (obtained from the American Type Culture Collection, ATCC HTB-77) were maintained in McCoy’s 5A Medium (Corning) supplemented with 10% FBS, 2 mM l-glutamine, 100 U/mL penicillin, and 100 μg/mL streptomycin.

Molecular Cloning

All enzymes and reagents used for cloning were obtained from New England Biolabs (Frankfurt am Main, Germany) unless stated otherwise. Plasmids were generated using standard restriction cloning techniques, in detail: the plasmid pcDNA3.1-Nluc-hY4R encoding the N-terminally Nanoluciferase-tagged human NPY Y4 receptor, was cloned by exchanging the receptor-encoding sequence in Nluc-hH4R (in pcDNA3.1) (60) with the sequence of the hY4R. Therefore, the nucleotide sequence for the hY4R was first amplified from pcDNA3.1-hY4R (obtained from Missouri cDNA Research Center, Rolla, MO) via polymerase chain reaction (PCR, using Phusion High Fidelity Polymerase according to the manufacturer’s instructions) attaching flanking BamHI and ApaI to the 5′ and 3′ ends, respectively. Sequences of the used primers were as follows: forward, CTGAGGATCCATGAACACCTCTCACCTC; reverse, GACTGGGCCCGAAATGGGATTGGACCTGCCAC. After a PCR purification step using the FastGene Gel/PCR Extraction Kit (Nippon Genetics, Düren, Germany), both the purified PCR product and the target plasmid Nluc-hH4R were subjected to a restriction digest using BamHI and ApaI restriction enzymes overnight at 37 °C. Restriction digests were gel purified, ligated using T4 DNA ligase at room temperature for 1 h, and transformed into Top10F′ Escherichia coli (made competent in house).
For the plasmid pcDNA3.1-hY4R-Nluc(intraECL2), which encodes the Y4R with Nluc in the second extracellular loop (inserted via short glycine-serine linkers after the leucine in position 195 (L195)), pcDNA3.1-hY4R served as a vector backbone and pcDNA3.1-Nluc-Y1R(Y192) (51) was used to extract the Nluc insert. First, pcDNA3.1-hY4R was linearized after L195 via PCR adding flanking BamHI (5′ end of the linearized vector) and SacI (3′ end of the linearized vector; note: the SacI restriction site was incorporated into a -SGGGGSS- linker sequence) restriction sites using the aforementioned Phusion polymerase. Sequences of the used primers were as follows: forward: GATCGGATCCGCAGATAAGGTGGTCTGTAC; reverse: GATCAGAGCTCCCTCCACCGCCACTCAGGAACTCCAGAGCCTTG. Next steps were performed as described above: After a PCR cleanup, the purified PCR product and the pcDNA3.1-Nluc-Y1R(Y192) (used to excise the Nluc insert) were subjected to a restriction digest using BamHI and SacI restriction enzymes at 37 °C overnight. Restriction digests were gel purified, ligated, and transformed into competent E. coli, as described above. Plasmids were extracted from bacterial overnight cultures (Miniprep Kit, Nippon Genetics, Düren, Germany) and all sequences were verified by Sanger Sequencing (Eurofins Genomics, Ebersberg, Germany).
Plasmids used to prepare MultiBac(Mam) baculoviruses, (61) were generated according to the ACEMBL technology concept. (62) Acceptor and donor plasmids, used to prepare multicomponent DNA constructs, were joined by the Cre-LoxP fusion reaction. The hY4R-Nluc sequence from pcDNA3.1-hY4R-NlucC (37) (note: NlucC, fused to the C-terminus of Y4R, stands for the C-terminal fragment of Nluc consisting of 11 amino acids) was recloned into the pIMACE-CMVintP10 vector. This vector was obtained by replacing the insect polyhedrin promoter in pACEBac1 (Geneva Biotech, Pregny-Chambésy, Switzerland) with dual-host (insect/mammalian) CMVintP10 promotor taken from pSPL-IM vector (Geneva Biotech). For this purpose, pACEBac1 and pSPL-IM were both digested with the FastDigest restriction enzymes Bsu15I and XmaJI (Thermo Fisher Scientific) according to the manufacturer’s instructions. Restriction digests were gel purified using the FavorPrep Gel/PCR Purification Kit (Favorgen, Vienna, Austria), ligated by T4 DNA ligase (Thermo Fisher Scientific) at 22 °C for 1 h, and transformed into DH5α E. coli (made competent in house). After selection with gentamycin (MP Biomedicals, Eschwege, Germany), plasmids were extracted from overnight bacterial cultures (FavorPrep Plasmid DNA Extraction Kit). The resulting pIMACE-CMVintP10 was digested with HindIII and ScaI and pcDNA3.1-hY4R-Nluc was digested with HindIII and MssI to obtain the acceptor plasmid pIMACE-CMVintP10-hY4R-Nluc after gel purification and ligation. Depending on the targets of further use (mammalian, insect, or insect with mammalian-like glycosylation expression system), the acceptor plasmid pIMACE-CMVintP10-hY4R-Nluc was subjected to different types of recombination.
To obtain baculoviruses used for hY4R transduction into human SK-OV-3 ovarian adenocarcinoma cells, the acceptor plasmid pIMACE-CMVintP10-hY4R-Nluc was recombined with the donor plasmid pIDC-VSVG using Cre recombinase (New England Biolabs) and the product was propagated in pirHC E. coli strain (Geneva Biotech) under selection pressure of gentamycin and chloramphenicol (Sigma). Pseudotyping of baculovirions with vesicular stomatitis virus G-protein (VSVG) has been reported to enhance the transduction efficiency of mammalian cells by these baculoviruses. The donor plasmid pIDC-VSVG was obtained by inserting VSVG from the pUCDM-VSVG vector (Geneva Biotech) into the pIDC vector (Geneva Biotech): pIDC was digested with RruI and XmaJI, and pUCDM-VSVG was digested with MssI and XmaJI. After the cre-recombination step, the resulting multicomponent construct pIMACE-CMVintP10-hY4R-Nluc × pIDC-VSVG was subjected to restriction analysis and used for transposition into baculovirus genomes presented as bacterial artificial chromosomes in E. coli cells (DH10MultiBac, Geneva Biotech) to generate baculovirus Y4RVSVG.
To create Y4R baculoviruses for the expression of hY4R in Sf9 insect cells, pIMACE-CMVintP10-hY4R-Nluc was directly used for transposition into DH10MultiBac. To mimic the mammalian-like pattern of protein glycosylation in Sf9 cells, another type of baculovirus, encoding for Y4RSwBac, was prepared. For this purpose, the acceptor plasmid pIMACE-CMVintP10-hY4R-Nluc was recombined with the donor plasmid pIDC-SweetBac. This plasmid was obtained by recloning the SweetBac module (49) from pUDCM-GntII-GalT plasmid (Geneva Biotech) into the pIDC backbone: pUDCM-GntII-GalT was digested with MssI and XmaJI, and pIDC was digested with RruI and XmaJI. As before, the generated multicomponent construct pIMACE-CMVintP10-hY4R-Nluc × pIDC-SweetBac was restriction analyzed and used for transposition into DH10MultiBac.

Generation of the Stable HEK293T-Nluc-hY4R and HEK293T-hY4R-Nluc(intraECL2) Cell Lines

HEK293T cells were seeded into a 6-well dish in DMEM supplemented with 10% FBS at 300,000 cells/mL. The following day, cells were transfected with 2 μg of pcDNA3.1-Nluc-hY4R or pcDNA3.1-hY4R-NLuc(intraL195) using the XtremeGene HP transfection reagent (Merck, Darmstadt, Germany) according to the manufacturer’s protocol. Cells were passaged after 48 h into a T175 culture flask and selected for 2 weeks at a high G418 concentration (1000 μg/mL). The culture medium was changed three times in this time period and the cells were regularly monitored for colonial growth. Starting with the observation of the latter, cells were cultured in DMEM supplemented with 10% FBS and 600 μg/mL G418 for maintained selection pressure.

Buffers and Media Used for Binding and Functional Assays

Buffer I (used for radiochemical (Y2R, Y4R) and flow cytometric (Y4R) binding experiments): a hypotonic sodium-free HEPES buffer (25 mM HEPES, 2.5 mM CaCl2, 1 mM MgCl2, adjusted to pH 7.4 using 25% ammonia in water) supplemented with 1% BSA (Serva, Heidelberg, Germany) and 0.1 mg/mL bacitracin (Serva).
Buffer II (used for binding experiments at the Y1R and Y5R): an isotonic sodium-containing HEPES buffer (10 mM HEPES, 150 mM NaCl, 25 mM NaHCO3, 2.5 mM CaCl2, 5 mM KCl, pH 7.4) supplemented with 1% BSA.
Buffer III (used for the Ca2+ aequorin assay): an isotonic sodium-containing HEPES buffer (25 mM HEPES, 120 mM NaCl, 1.5 mM CaCl2, 1 mM MgCl2, 5 mM KCl, 10 mM d-glucose, pH 7.4).
DPBS (used for flow cytometric and fluorescence anisotropy-based Y4R binding assays): Dulbecco’s phosphate-buffered saline with calcium and magnesium (1.8 mM CaCl2, 2.68 mM KCl, 1.47 mM KH2PO4, 3.98 mM MgSO4, 137 mM NaCl, 8.06 mM Na2HPO4, pH 7.4) supplemented with 1% BSA and 0.1 mg/mL bacitracin (flow cytometry) or with 0.1% Pluronic F-127 and cOmplete EDTA-free Protease Inhibitor Cocktail (fluorescence anisotropy).
L15-HEPES (used for miniGsi recruitment, cAMP CAMYEN and NanoBRET binding assays): Leibovitz’s L-15 medium (140 mM NaCl, 1.3 mM CaCl2, 1 mM MgCl2, various amino acids and vitamins) (Fisher Scientific, Schwerte, Germany) supplemented with 10 mM HEPES and additionally with 5% FBS (miniGsi recruitment and cAMP CAMYEN assays) or 1.5% BSA (NanoBRET). Before the addition of FBS and BSA, the pH was adjusted to 7.4.

Radioligand Binding Assays

All radioligand binding assays were performed at intact Y receptor expressing cells at 23 ± 2 °C. The radiochemical competition binding assays, used to study Y1, Y2, Y4, and Y5 receptor binding of peptide 11, and fluorescent ligands 1619, were recently validated by the determination of binding affinities of the reference ligands pNPY (Y1R, Y2R, and Y5R) and hPP (Y4R), which were in good agreement with previously reported binding affinities (for determined and reference pKi or Ki values of pNPY (Y1R, Y2R, Y5R), see Konieczny et al.; (30) for determined and reference pKi values of hPP (hY4R), see Gleixner et al. (31)).

Y1R Binding

Competition binding assays at Y1R-expressing SK-N-MC neuroblastoma cells were performed as previously described using [3H]UR-MK299 as radioligand (concentration: 0.15 nM). (33) Due to low radioligand displacement, no curve fitting was performed for the investigated compounds 11 and 1719 (studied at concentrations up to 10 μM for 11 and 3 μM for 1719). For compound 16, binding data (dpm) were normalized (100% = bound radioligand in the absence of competitor; 0% = unspecific binding), plotted as % over log(concentration of competitor, M), and analyzed by a four-parameter logistic fit (log(inhibitor) vs response–variable slope; GraphPad Prism 5, GraphPad Software, San Diego, CA) to obtain pIC50 and IC50 values, which were converted to pKi and Ki values according to the Cheng–Prusoff equation (63) (logarithmic form in the case of pKi values).

Y2R Binding

Competition binding assays at CHO-hY2R cells were performed as previously described using [3H]propionyl-pNPY (Kd = 0.14 nM, concentration: 0.5 nM) as radioligand. (34) Due to low radioligand displacement, no curve fitting was performed for the investigated compounds 11 and 1619 (studied at concentrations up to 10 μM for 11 and 3 μM for 1619).

Y4R Binding

Competition binding experiments at CHO-hY4R-Gqi5-mtAEQ cells using [3H]UR-KK200 as radioligand (Kd = 0.67 nM, concentration: 1.0 nM) (35) were performed as previously described. (30) Binding data (dpm) were normalized (100% = bound radioligand in the absence of competitor; 0% = unspecific binding), plotted as % over log(concentration of competitor, M), and analyzed by a four-parameter logistic fit (log(inhibitor) vs response–variable slope; GraphPad Prism 5) to obtain pIC50 and IC50 values, which were converted to pKi and Ki values according to the Cheng–Prusoff equation (63) (logarithmic form in the case of pKi values).

Y5R Binding

Competition binding studies at HEC-1B-hY5R cells using [3H]propionyl-pNPY (Kd = 11 nM, (26) concentration: 5 nM) as radioligand were performed as previously reported. (35) Due to low radioligand displacement, no curve fitting was performed for the investigated compounds 11, 17, and 19 (studied at concentrations up to 10 μM for 11, 3 μM for 17, and 1 μM for 19). For compounds 16 and 18, binding data (dpm) were normalized (100% = bound radioligand in the absence of competitor; 0% = unspecific binding), plotted as % over log(concentration of competitor, M), and analyzed by a four-parameter logistic fit (log(inhibitor) vs response–variable slope; GraphPad Prism 5) to obtain pIC50 and IC50 values, which were converted to pKi and Ki values according to the Cheng–Prusoff equation (63) (logarithmic form in the case of pKi values).

Functional Y4R Assays

All functional assays were performed in agonist mode as all studied compounds exhibit Y4R agonistic activity.

Y4R Ca2+ Aequorin Assay

The assay was performed with live CHO-hY4R-Gqi5-mtAEQ cells as previously described (used buffer: buffer III). Measurements were carried out on a Tecan Infinite Lumi plate reader. Data analysis was performed as reported using GraphPad Prism 5 for the calculation of the area under the curve. Fractional luminescences, obtained by dividing the area of the initially occurring agonist-induced peak by the total area (sum of the areas of the agonist-induced peak and the subsequent lysis-induced peak), were normalized (100% = fractional luminescence obtained from 1 μM hPP, 0% = fractional luminescence obtained by vehicle addition (basal effect), GraphPad Prism 5). The normalized responses were plotted against log(concentration of agonist, M) and the data were fitted according to a four-parameter logistic equation (log(agonist) vs response–variable slope, GraphPad Prism 5) to obtain pEC50 values, which were converted to EC50 values. Efficacies Emax correspond to the upper plateaus of the normalized concentration–response curves.

Y4R miniGsi Recruitment Assay

The assay was performed with live HEK293T-NlucN-mGsi/Y4R-NlucC cells using a reported protocol with a minor modification: (37) on the day of the experiment, cells were not washed with FBS-free medium, i.e., the assay was performed in L15-HEPES supplemented with 5% FBS. Furimazine was used as nanoluciferase substrate. The original stock was diluted 1:1000 in L15-HEPES to obtain the feed solution for the assay. Data analysis was performed based on the areas under the curves (time window: 45 min starting with the agonist addition) using GraphPad Prism 5. Normalized responses (100% = response induced by 1 μM hPP, 0% = vehicle control) were plotted against log(concentration of agonist, M) and the data were fitted according to a four-parameter logistic equation (log(agonist) vs response–variable slope, GraphPad Prism 5) yielding pEC50 values, which were converted to EC50 values. Efficacies Emax correspond to the upper plateaus of the normalized concentration–response curves.

Y4R cAMP CAMYEN Assay

The assay was performed with live HEK293T-CAMYEN-hY4R cells applying a reported protocol. (31) Experiments were performed in triplicate. Data processing was performed with GraphPad Prism 9.0 as previously described. (31)

Flow Cytometric Y4R Binding Assays

Flow cytometry-based Y4R binding studies with the fluorescent ligands 16, 17, and 18 were performed at intact CHO-hY4R-Gqi5-mtAEQ cells with a FACSCantoII flow cytometer (Becton Dickinson), equipped with an argon laser (488 nm), a red diode laser (640 nm) and a BD High Throughput Sampler (HTS unit) for microtiter plates. The latter was used for injection of the samples of saturation and competition binding experiments. At least three individual experiments were performed in duplicate (kinetic experiments) or triplicate (saturation and competition binding). The following gain settings for forward and sideward scatter were applied throughout: FSC: 0 V, SSC: 252 V. Fluorescence was recorded using the following settings: compound 16, excitation at 488 nm, emission at 585 ± 21 nm (PE channel), gain: 470–500 V; compound 17, excitation at 640 nm, emission at 660 ± 10 nm (APC channel), gain: 450 V; compound 18, excitation at 488 nm, emission at >670 nm (PerCP-Cy5.5 channel), gain: 470–500 V, propidium (cell viability studies), excitation at 488 nm, emission at 585 ± 21 nm (PE channel), gain: 330 V. For measurements requiring the use of classical injection tubes (kinetic experiments), the medium flow rate (60 mL/min) was used. Measurements were stopped after counting of 2000–10000 gated events.
2 to 3 days prior to the experiment, cells were seeded in T75 or T175 culture flasks. On the day of the experiment, cells were detached by trypsinization, suspended in culture medium, and centrifuged. The cell pellet was resuspended in buffer I (binding studies with 16) or DPBS (binding studies with 1618), and the cell density was adjusted to 1.5 × 105 (saturation binding) or 2 × 105 cells/mL (kinetic and competition binding experiments). All receptor ligands were added to the cell suspension as 100-fold concentrated (compared to the final concentration) solutions in DMSO/H2O 1:1 v/v (fluorescent ligands, all experiments; nonlabeled ligands, competition binding studies) or in 10 mM aq. HCl (nonlabeled ligands, saturation, and kinetic experiments).
For saturation binding experiments, the wells of a polypropylene round-bottom 96-well plate (Brand, Wertheim) were prefilled with 200 μL of cell suspension followed by the addition of 2 μL of 10 mM aq. HCl (samples for total binding) or 2 μL of a 100 μM solution of hPP (samples for unspecific binding). The final concentration of hPP was 1 μM. Incubation was started by adding 2 μL of a 100-fold concentrated fluorescent ligand solution followed by gentle shaking in the dark at 22 ± 2 °C for 2 h and subsequent measurement using the HTS unit (sample volume: 45 μL, injection speed: 1.5 μL/s). For the determination of autofluorescence, three ligand-free samples were prepared by the addition of 2 μL of 10 mM aq. HCl and 2 μL of DMSO/H2O 1:1 v/v to the cell suspension, followed by incubation and measurement.
For association experiments, polypropylene tubes (usable as injection tubes for the FACSContoII flow cytometer) were prefilled with 2000 μL of cell suspension and 20 μL of 10 mM aq. HCl (samples for total binding) or 20 μL of a 100 μM solution of hPP (samples for unspecific binding) were added. The association was started by the addition of 20 μL of a 30 nM (buffer I) or 70 nM (DPBS) solution of 16 (final concentrations: 0.3 and 0.7 nM, respectively), 20 μL of a 100 nM solution of 17 in DPBS (final concentration: 1.0 nM), or 20 μL of a 50 nM solution of 18 in DPBS (final concentration: 0.5 nM) to the cell suspension, followed by short mixing and gentle shaking under protection from light at 22 ± 2 °C. After different periods of time (0.5–120 min for all studied fluorescent ligands), sample aliquots were measured by placing the tube in the injection port of the cytometer.
For dissociation experiments, polypropylene tubes (usable as injection tubes for the FACSContoII flow cytometer) were prefilled with 2000 μL of cell suspension followed by the addition of 20 μL of 10 mM aq. HCl (samples for total binding) or 20 μL of a 100 μM solution of hPP (samples for unspecific binding). Preincubation was started by the addition of 20 μL of a 150 nM (buffer I) or 350 nM (DPBS) solution of 16 (final concentrations: 1.5 and 3.5 nM, respectively), 20 μL of a 500 nM solution of 17 in DPBS (final concentration: 5 nM) or a 250 nM solution of 18 in DPBS (final concentration: 2.5 nM). The samples were gently shaken in the dark at 22 ± 2 °C for 2 h. The dissociation was initiated by the addition of 20 μL of a solution of hPP and 5 (final concentrations: 1000-fold (hPP) and 100-fold (5) over the final concentration of the respective fluorescent ligand) prepared in the buffer used for the respective experiment (note: a combination of hPP and 5 was used to prevent rebinding of the labeled ligand more effectively). After different periods of time (16 in buffer I: 1–360 min, 16 in DPBS: 1–300 min, 17 in DPBS: 1–200 min, 18 in DPBS: 1–300 min), sample aliquots were measured by placing the tube in the injection port of the cytometer.
For competition binding experiments with 16, the wells of a polypropylene round-bottom 96-well plate were prefilled with cell suspension (200 μL) followed by the addition of 2 μL of a 100-fold concentrated solution of the competing (nonlabeled) ligand and gentle shaking in the dark for 5 min. 2 μL of the fluorescent ligand solution (100-fold concentrated) were added and shaking in the dark at 22 ± 2 °C was continued for 2 h. The final concentrations of 16 were 0.5 nM (buffer I) and 1 nM (DPBS). For the determination of unspecific binding and total binding in the absence of the studied competitor, 2 μL of a 100 μM solution of hPP (final concentration: 1 μM) and 2 μL of DMSO/water 1:1 v/v, respectively, were initially added instead of the competitor solution, followed by the addition of the fluorescent ligand after the preincubation period and further processing as described above.
Specific binding data were obtained by subtracting unspecific binding from total binding. Unspecific binding data shown in Figure 4 are autofluorescence corrected. Specific binding data from saturation binding experiments were analyzed by an equation describing a hyperbolic isotherm (binding-saturation: one site-specific binding, GraphPad Prism 5) yielding Kd values, which were converted to pKd values. Specific binding data from association experiments were analyzed by a three-parameter equation describing a monophasic exponential rise to a maximum (Y0 constrained to zero), yielding kobs(mono), and by a five-parameter equation describing a biphasic exponential rise to a maximum (Y0 constrained to zero) yielding kobs(bi,fast) and kobs(bi,slow) (GraphPad Prism 5). Specific binding data from dissociation experiments were analyzed by a two-parameter equation describing a monophasic exponential decline (GraphPad Prism 5) yielding koff. Association rate constants kon(mono), kon(bi,fast), and kon(bi,slow) were calculated from the observed association rate constants kobs(mono), kobs(bi,fast), and kobs(bi,slow) (mean values), the mean values of the corresponding dissociation rate constants koff (cf. Table 4), and the fluorescent ligand concentrations used for the association experiments (see above) according to the equation kon = (kobskoff)/[fluorescent ligand]. The kinetically derived dissociation constants Kd(kin) were calculated from kon(mono) and the mean value of the dissociation rate constants koff (cf. Table 4) according to the equation Kd(kin) = koff/kon.
Total binding data from competition binding experiments were plotted as fluorescence intensity over log(competitor concentration, M) and analyzed by a four-parameter logistic equation [log(inhibitor) vs response-variable slope, GraphPad Prism 5] to obtain the TOP value (upper curve plateau), which was used for data normalization (100% = TOP value, 0% = average signal obtained from unspecific binding). The normalized data (%) were plotted over log(competitor concentration, M) and analyzed by a four-parameter logistic equation to obtain pIC50 values, which were converted to pKi values according to the Cheng–Prusoff equation (logarithmic form) using the following Kd values of 16: buffer I, Kd = 0.30 nM; DPBS, Kd = 0.70 nM (mean Kd values from three individual saturation binding experiments).
For cell viability studies, cell suspensions were prepared as for fluorescent ligand binding experiments, but with a higher cell density (3 × 105 cells/mL for buffer I and DPBS). The cell suspensions were shaken in 50 mL polypropylene vessels (VWR International, Radnor, PA) at 22 °C. For measurements, 250 μL aliquots were transferred into polystyrene measurement tubes. 2.5 μL of an aqueous solution of propidium iodide (200 μg/mL; final concentration: 2 μg/mL) were added followed by a short incubation period (1 min) and measurement. Measurements were stopped after counting of 20,000 gated events.

Fluorescence Anisotropy Y4R Binding Assays

Fluorescence anisotropy-based Y4R binding studies with fluorescent ligand 16 were performed at hY4R displaying budded baculovirus particles (termed BBVs below), which were obtained from Sf9 insect cells following the procedure described for the preparation of Y1R displaying BBVs. (40) Two types of MultiBac baculoviruses were used: one type encodes only for hY4R resulting in BBVs displaying Y4R devoid of a mammalian-like glycosylation (herein termed Y4Rnonglyco) and the second type additionally encodes for enzymes providing a mammalian-like glycosylation in Sf9 cells (herein termed Y4RSwBac; see also Section Molecular Cloning). The obtained viruses were collected by centrifugation at 1600g for 10 min, and the virus titer was determined with an image-based cell-size estimation assay as described elsewhere. (64) The viruses were amplified using a multiplicity of infection (MOI) between 0.01 and 0.1 until a high-titer baculovirus stock (virus concentration of at least 1.0 × 108 infectious viral particles/mL) was acquired. To produce BBV preparations with a high receptor density, Sf9 cells with a density of 1.9 × 106 cells/mL were infected with an MOI of 3 and the BBVs were collected by centrifugation 63 h after the infection, when the cell viability was <30%. The BBVs were concentrated 40-fold by centrifugation at 48,000g at 4 °C for 40 min, washed with ice-cold DPBS, resuspended, and homogenized using a 0.3 mm diameter needle (Sterican, B. Braun, Melsungen, Germany). Aliquots of the BBV preparations were stored at −90 °C until use.
FA measurements were carried out using a Synergy NEO multimode plate reader (BioTek, Winooski, VT) and black, half-area, flat-bottom polystyrene NBS (nonbinding surface) 96-well plates (Corning, Corning, NY, product no. 3993). Polarizing excitation (530 ± 15 nm) and dual emission (590 ± 17.5 nm) filters with a dichroic mirror were used, allowing the simultaneous detection of parallelly and perpendicularly polarized fluorescence emission. All measurements were performed at 27 °C. Saturation binding experiments, association and dissociation experiments, and competition binding experiments were performed in duplicate, following a previously reported protocol with minor modifications. (65) All ligand dilutions were prepared in DPBS at 4- or 5-fold concentrations compared to the final concentration in the assay. The final total volume per well was 100 μL. Frozen preparations of hY4R displaying BBVs were thawed and thoroughly resuspended using a 1 mL syringe (Norm-Ject-F, Braun Melsungen) with a 0.3 mm diameter needle (Sterican, B. Braun).
For equilibrium binding studies (saturation binding experiments), used for the determination of equilibrium Kd values and the Y4R concentration in the BBV preparations, 25 μL of a 4-fold concentrated solution of 16 (used at final concentrations of 0.5 and 2 nM) and 25 μL of DPBS were premixed in the 96-well plate followed by the addition of 50 μL of the BBV suspension (used in 6 or 12 different dilutions) to initiate total binding. To determine unspecific binding, the procedure was the same, but instead of neat binding buffer, 25 μL of a 4 μM solution of hPP (final concentration: 1 μM) were added. For blank measurements, 50 μL of the BBV suspension were added to 50 μL of DPBS. Experiments were performed with 6 or 12 different BBV dilutions. Measurements were started immediately after the addition of BBV suspension and were stopped after 90 min.
For association experiments, 20 μL of a 5-fold concentrated solution of 16 (used at final concentrations of 0.5 and 2 nM) and 20 μL of DPBS were premixed in the 96-well plate followed by the addition of 60 μL of the BBV suspension (used in six different dilutions) to initiate total binding. To determine unspecific binding, the procedure was the same, but instead of neat binding buffer, 20 μL of a 5 μM solution of hPP (final concentration: 1 μM) were added. For blank measurements, 60 μL of the BBV suspension were added to 40 μL of binding buffer. Note that due to the very fast association of 16 with the Y4R, measurements were not performed simultaneously for all BBV dilutions. Instead, to enable short time intervals between the readout cycles, individual measurements (for 10 min) were performed for all BBV concentrations. For each combination of receptor and ligand concentration, binding equilibrium was reached within this time period.
For dissociation experiments, the preincubation was set up as the association experiments (final concentrations of 16 of 0.5 or 2 nM, same BBV dilutions). Samples were incubated for 1 h, followed by starting the dissociation by the addition of 2 μL of a solution containing 50 μM hPP and 5 μM 5 (final concentrations: 1 μM and 100 nM, respectively) prepared in DPBS, and the FA signal was recorded for 2 h.
Note that in principle, association, equilibrium binding, and dissociation experiments can be performed simultaneously in one plate. However, as mentioned above, association experiments, requiring a higher data sampling rate than equilibrium binding and dissociation experiments, had to be performed consecutively for the various combinations of BBV and ligand concentrations. In contrast, in the case of equilibrium binding and dissociation experiments, all combinations were measured simultaneously.
For competition binding experiments (only performed with Y4RSwBac), 20 μL of a 5-fold concentrated solution of the competitive ligand (used at varying concentrations) and 20 μL of a 5-fold concentrated solution of 16 (final concentration: 0.5 nM) were premixed in the 96-well plate followed by the addition of 60 μL of the BBV suspension (final concentration of Y4RSwBac: 0.13 nM) to initiate total binding. To determine total binding in the absence of competitor and unspecific binding, the same procedure was used, but instead of the addition of competitive ligand, 20 μL of neat binding buffer and 20 μL of a 5 μM solution of hPP (final concentration: 1 μM), respectively, were added. FA signals were measured for 1.5 h.
Data processing for FA binding assays was assisted by the software Aparecium 2.0.20 developed in house (http://www.gpcr.ut.ee/software.html). Prior to the calculation of FA using eq 1, (19) the measured parallel [I(t)] and perpendicular [I(t)] fluorescence intensities were blank-corrected on the basis of the values obtained from wells containing BBVs but no receptor ligand.
FA(t)=I(t)I(t)I(t)+2I(t)
(1)
where I(t) and I(t) are the blank-corrected parallel and perpendicular fluorescence intensities, respectively, measured at time t, and FA(t) is the calculated fluorescence anisotropy at time t. FA equilibrium binding data were globally fitted using a modified version of a previously described user-defined GraphPad Prism-compatible binding model (19) (for the detailed equation, see the Supporting Information), which takes ligand depletion into account and yields the dissociation constant Kd and simultaneously the Y4R concentration of the BBV stock. FA association binding data were globally fitted to obtain association rate constants kon and Kd(kin) values also using a previously presented user-defined GraphPad Prism-compatible binding model, which takes ligand depletion into account and requires the specification of the koff value (the mean koff value obtained from the dissociation experiments was used). (19) Individual association rate constants kon and individual pKd values, obtained from at least three independent FA association binding experiments using six different BBV concentrations and two different concentrations of 16, were averaged to obtain mean kon and mean pKd values ± SEM (cf. Table 5). Kinetically derived dissociation constants Kd(kin) were calculated from individual (n = 3) kon values and the mean koff value (cf. Table 5) according to the equation Kd(kin) = koff/kon, and averaged. The mean value ± SEM of the dissociation rate constant koff was calculated from individual koff values obtained from three independent FA dissociation experiments each using six different BBV concentrations and two different concentrations of 16 (GraphPad Prism 5). Data from FA competition binding experiments using 16 as the fluorescent probe (Kd = 0.14 nM, mean value from four individual FA equilibrium binding experiments) were analyzed by a four-parameter logistic equation (log(inhibitor) vs response–variable slope, GraphPad Prism 5) to obtain pIC50 values, which were converted to pKi values according to the Cheng–Prusoff equation (63) (logarithmic form) using the Kd value for 16 that is mentioned above.

NanoBRET Y4R Binding Assay

BRET-based Y4R binding studies with 16 (saturation binding, association, dissociation, displacement by Y4R reference ligands) were performed following a protocol reported for a NanoBRET Y4R binding assays using HEK293T cells stably expressing an N-terminally Nluc-tagged human Y4R. (51) One the day prior to the experiment, HEK293T-hY4R-Nluc(intraECL2) cells were detached from the culture flask by trypsinization, centrifuged (500g, 5 min), and resuspended in FBS-free L15-HEPES. The cell density was adjusted to 1.14 × 106 cells/mL, and the cells were seeded into white 96-well plates (Brand, Wertheim, Germany) (70 μL, 80,000 cells per well) and incubated overnight in a water-saturated atmosphere (only atmospheric CO2). On the day of the experiment, dilutions of 16, 5, and Y4R ligands studied in competition binding assays were prepared as 10-fold concentrated (compared to the final concentration) solutions in L15-HEPES supplemented with 5% BSA (in the following referred to as L15-HEPES-BSA). The addition of these feed solutions to the wells resulted in a 1:10 dilution and the final compound concentrations. The final BSA concentration in the assay was 1.5% (30 μL of a solution containing 5% BSA added to 70 μL of BSA-free medium, as outlined below).
All BRET measurements were performed at a temperature of 25 °C using a TECAN Infinite Lumi plate reader (Tecan, Crailsheim, Germany) equipped with an injection module. Bioluminescence of the nanoluciferase was detected using a 460/35 nm band-pass (460/35 BP) filter. The emission of the fluorescent ligands was detected through a 610 nm long-pass (610 LP) filter. For all experiments integration times were set as follows: 100 ms for the 460/35 BP channel and 1000 ms for the 610 LP filter to increase S/N.
For saturation binding experiments, 10 μL of L15-HEPES-BSA and 10 μL of the 10-fold concentrated solution of 16 (determination of total binding) or 10 μL of a 10 μM solution of 5 and 10 μL of the 10-fold concentrated solution of 16 (determination of unspecific binding) were added to the cells. After incubation under shaking in the dark at 22 ± 1 °C for 1.5 h, 10 μL of a 10 μM solution of the Nluc substrate coelenterazine h in L15-HEPES-BSA (obtained by a 1:200 dilution of a 2 mM stock solution in methanol) were added to each well, the plate was transferred to the preheated plate reader and luminescence signals were recorded for 15 min (5 cycles).
For association experiments, 10 μL of L15-HEPES-BSA (determination of total binding) or a 10 μM solution of 5 in L15-HEPES-BSA (determination of unspecific binding) were added to the cells followed by the addition of 10 μL of the feed solution of the Nluc substrate furimazine (obtained by diluting the furimazine stock 1:1000 with L15-HEPES-BSA). The plate was immediately transferred to the plate reader, luminescence was measured for one cycle (ca. 15 s) followed by the addition of 10 μL of a 10 nM solution of 16 (final concentration: 1 nM) to each well via the injector module and continued measurement for 60 min (300 cycles).
For dissociation experiments, 10 μL of L15-HEPES-BSA (determination of total binding) or 10 μL of a 10 μM solution of 5 in L15-HEPES-BSA (determination of unspecific binding) were added to the cells followed by the addition of 10 μL of a 35 nM solution of 16 in L15-HEPES-BSA and incubation under shaking in the dark at 22 ± 1 °C for 1.5 h. 10 μL of the feed solution of the Nluc substrate furimazine (obtained by diluting the furimazine stock 1:600 with L15-HEPES-BSA) were added to each well and the plate was immediately transferred to the plate reader. Luminescence was measured for five cycles (1.5 min) followed by the addition of 10 μL of a 10 μM solution of 5 in L15-HEPES-BSA to each well with a pipette and continued measurement of luminescence for 150 min (750 cycles). In order to take the time span between the addition of the 10 μM solution of 5 (initiation of the dissociation) and the start of the measurement into account for data analysis, this delay was determined with a stopwatch.
For competition binding experiments, 10 μL of a 10-fold concentrated solution of the competing ligand (determination of total binding in the presence of a competitor), 10 μL of L15-HEPES-BSA (determination of total binding in the absence of competitor), or 10 μL of a 10 μM solution of 5 in L15-HEPES-BSA (determination of unspecific binding) were added to the cells. After a short period of preincubation (5 min), 10 μL of a 15 nM solution of 16 in L15-HEPES-BSA were added followed by incubation under gentle shaking in the dark at 22 ± 2 °C for 1.5 h. 10 μL of the 10 μM feed solution of the Nluc substrate coelenterazine h in L15-HEPES-BSA (obtained as described above) were added, the plate was immediately transferred to the plate reader and luminescence was measured for 15 min (5 cycles).
BRET ratios were calculated by dividing the signal measured in the 610LP channel through the signal obtained from the 460/35 BP channel. Corrected BRET ratios were calculated by subtracting the BRET ratio obtained from a vehicle control sample (no receptor ligand added) from the BRET ratios obtained from samples containing fluorescent ligand. Specific binding data were obtained by subtracting unspecific binding from total binding. Specific binding data from saturation binding experiments were analyzed by an equation describing a hyperbolic isotherm (binding-saturation: one site-specific binding, GraphPad Prism 5) yielding Kd values, which were converted to pKd values. Specific binding data from association experiments were analyzed by a three-parameter equation describing a monophasic exponential rise to a maximum (Y0 constrained to zero) yielding kobs(mono) values, and by a five-parameter equation describing a biphasic exponential rise to a maximum (Y0 constrained to zero) affording kobs(bi,slow) and kobs(bi,fast) values (GraphPad Prism 5). Specific binding data from dissociation experiments were analyzed by a three-parameter equation describing a monophasic exponential decline (Y0 constrained to 100%) yielding koff(mono) values, and by a five-parameter equation describing a biphasic exponential decline (Y0 constrained to 100%, plateau constrained to zero) yielding koff(bi,fast) and koff(bi,slow) values (GraphPad Prism 5.0). Note: Dissociation data were initially analyzed with a variable plateau. As the plateau mean value was not significantly different from zero in the case of the biphasic analysis (P > 0.05, t test), data were reanalyzed with the plateau value constrained to zero.
The association rate constants kon(mono), kon(bi,fast), and kon(bi,slow) were calculated from the observed association rate constants kobs(mono), kobs(bi,fast), and kobs(bi,slow), respectively (mean values), the mean values of the dissociation rate constants koff(mono), koff(bi,fast), or koff(bi,slow) (cf. Table 6), and the fluorescent ligand concentration used for the association experiments (see above) according to the equation kon = (kobskoff)/[fluorescent ligand]. The kinetically derived dissociation constants Kd(kin) were calculated from kon(mono), kon(bi,fast), or kon(bi,slow) and the mean values of the dissociation rate constants koff(mono), koff(bi,fast), or koff(bi,slow) (cf. Table 6) according to the equation Kd(kin) = koff/kon.
Total binding data from competition binding experiments were plotted as corrected BRET ratio over log(competitor concentration, M) and analyzed by a four-parameter logistic equation (log(inhibitor) vs response–variable slope, GraphPad Prism 5) to obtain the TOP value (upper curve plateau), which was used for data normalization (100% = TOP value, 0% = average of the corrected BRET ratio obtained from unspecific binding). The normalized data (%) were plotted over log(competitor concentration, M) and analyzed by a four-parameter logistic equation to obtain pIC50 values, which were converted to pKi values according to the Cheng–Prusoff equation (63) (logarithmic form) using the following Kd value for 16: Kd = 0.94 nM (mean value from three individual saturation binding experiments).

Confocal Microscopy

Confocal microscopy was performed with a Zeiss LSM 710 confocal laser scanning microscope (Zeiss). The objective was 63× magnification with oil (1.4 NA). One day prior to the experiment CHO-hY4R-Gqi5-mtAEQ cells were seeded in Nunc LabTekTM II cover glasses with 8 chambers (Thermo Fisher Scientific). On the day of the experiment, the confluency of the cells was 60–80%. The culture medium was removed, cells were washed with L15 medium (200 μL) and covered with L15 medium (100 μL) containing H33342 (2 μg/mL). L15 medium (100 μL) containing H33342 (2 μg/mL) and either compound 16 or 17 (final concentration: 20 nM) were added for total binding. For the determination of unspecific binding, L15 medium (100 μL) containing H33342 (2 μg/mL) and hPP (final concentration: 1 μM), and L15 medium (100 μL) containing H33342 (2 μg/mL) and either 16 or 17 (final concentration: 20 nM) were added. Images were acquired after an incubation period of 30 min (37 °C). The settings for compound 16 were: laser power/pinhole: 405 nm: 1.8%/1.97 airy units, 561 nm: 2%/1.32 airy units. The settings for compound 17 were: laser power/pinhole: 405 nm: 2.0%/1.90 airy units, 633 nm: 28%/1.35 airy units. Filter settings for fluorescence detection: 410–549 nm (H33342) (studies with 16), 410–585 nm (H33342) (studies with 17), 562–649 nm (16), and 638–759 nm (17).

Wide-Field Epifluorescence and TIRF Microscopy

For fluorescent ligand binding imaging experiments, human SK-OV-3 ovarian adenocarcinoma cells (ATCC HTB-77) were seeded at a density of 20,000 cells per well into eight-well CG imaging chambers (Zell Kontakt GmbH, Nörten-Hardenberg, Germany). After incubation for 24 h in McCoy’s 5A Medium (supplemented with 10% FBS, 2 mM l-glutamine, 100 U/mL penicillin, and 100 μg/mL streptomycin) at 37 °C in a humidified atmosphere (95% air, 5% CO2), recombinant MultiBacMam baculoviruses of Y4RVSVG were added (cell confluency: approximately 70%) at an MOI of 3 and incubated for additional 24 h (medium was supplemented with 1 mM sodium butyrate). Native cells without the addition of Y4R baculovirus were used as a negative control. Prior to imaging, nuclei were stained with 2 mM (final concentration) Hoechst 34580 (Chemodex Ltd., St. Gallen, Switzerland) for 5 min. The culture medium was replaced by LiveLight MEMO imaging medium (Cell Guidance Systems, Cambridge, U.K.) (200 μL, supplemented with supplement A according to the manufacturer’s protocol). After incubation at 37 °C in a humidified atmosphere containing 5% CO2 for 1 h, compound 16 (final concentration 0.1 or 1 nM) was added. In the case of unspecific binding, hPP was additionally added at a final concentration of 10 μM. Cells were imaged after 30 min of incubation (37 °C, atmosphere with 5% CO2). Imaging was performed using a microscope setup as described previously. (66) Briefly, wide-field epifluorescence and TIRF imaging were performed using an inverted microscope based on a Till iMIC body (Till Photonics/FEI, Munich, Germany), equipped with a UPLSAPO 20× oil (NA 0.85) and TIRF APON 60× oil (NA 1.49) objectives (Olympus Co., Tokyo, Japan). The samples were alternately excited using a 405 nm (120 mW) PhoxX laser diode (Omicron-Laserage, Rodgau, Germany) for Hoechst 34580 or at 561 nm (106 mW) for compound 16. Lasers were combined in a SOLE-6 light engine (Omicron-Laserage) and their emission was coupled into a Yanus scan head, which, along with a Polytrope galvanometric mirror (Till Photonics/FEI), was used to position the laser beam for wide-field epifluorescence or TIRF (approximately 100 nm depth) illumination. Excitation and emission light were spectrally separated using an imaging filter cube containing a flat 2 mm beamsplitter zt 405/488/531/640rpc (Chroma Technology Co.) and a TIRF emission filter ZET 405/488/561/640 (Chroma Technology Co.). In addition, a bright-field channel was used to define cell boundaries. The electron-multiplying charge-coupled device Ultra 897 camera (Andor Technology, Belfast, U.K.) was mounted to the microscope via a TuCam adapter with 2× magnification (Andor Technology). The camera was cooled to −100 °C using an Oasis 160 liquid recirculating chiller (Solid State Cooling Systems, Wappingers Falls, NY). All measurements were performed in the eight-well CG imaging chambers, and at least 10 selected areas per well were recorded as 16-bit multicolor OME-TIFF Z-stacks (100 frames, with 200 nm piezo-focusing increment in the case of the 60× objective or 500 nm in case of 20× objective), with an exposure time of 100 ms and an EM gain of 300 in the case of wide-field epifluorescence imaging or one-color time stacks (1000 frames at 30 Hz) in the case of single-particle tracking TIRF imaging. For different conditions, the laser powers were varied to enable optimal signal levels (the absence of a detectable crosstalk signal in the channels was ensured). Wide-field epifluorescence Z-stacks were deconvolved using the EpiDEMIC plugin. (67) Single-particle time lapse tracking was performed in the Icy image analysis environment using the Spot Tracking (68) plugin and classified using Track Manager. Tracks longer than 150 frames were filtered out for video presentation.

Statistical Significance

For the applied statistical tests (F-test, t test), the significance level was set to P = 0.05.

Calculation of Propagated Errors

Propagated errors (applying to specifically bound fluorescent ligand (saturation binding), association rate constants kon, and kinetically derived dissociation constants Kd(kin) obtained from flow cytometric and NanoBRET assays) were calculated as described elsewhere. (31)

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsptsci.4c00013.

  • Preparation of the azido-functionalized Py-5 derivative 15; chromatograms of the investigation of the chemical stability of 1618 in PBS pH 7.4 (Figure S1); radioligand displacement curves from competition binding experiments with [3H]UR-KK200 (Figure S2); concentration–response curves of hPP, 11 and 1619 obtained from a hY4R Ca2+ aequorin assay (Figure S3); concentration–response curves of hPP, 5, 11, and 1619 obtained from a hY4R mini-Gsi protein recruitment assay (Figure S4); concentration–response curves of hPP, 16 and 17 obtained from a hY4R CAMYEN cAMP assay (Figure S5); study of dummy fluorescent ligands in functional assays (Figure S6); excitation and emission spectra (Figure S7); viability of CHO-hY4R-Gqi5-mtAEQ cells (Figure S8); analyses of flow cytometric saturation binding data of 16 based on nonviable and viable cell populations (Figure S9); syntax of the equation used to fit FA equilibrium binding data (GraphPad Prism 5); RP-HPLC chromatograms of compounds 11 and 1619; and 1H NMR spectra of compound 11 (PDF)

  • TIRF video sequence (recorded and shown at a 30 Hz resolution, shown at double speed) showing the interactions of 16 with the basal plasma membrane of adherent SK-OV-3-Y4R cells including single-particle tracking (magnified region) (AVI)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
    • Jakob Gleixner - Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, GermanyOrcidhttps://orcid.org/0000-0002-8488-870X
    • Sergei Kopanchuk - Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, EstoniaOrcidhttps://orcid.org/0000-0003-1756-9327
    • Lukas Grätz - Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, GermanyPresent Address: Section of Receptor Biology & Signaling, Department of Physiology & Pharmacology, Karolinska Institutet, S-17165 Stockholm, SwedenOrcidhttps://orcid.org/0000-0001-6755-0742
    • Maris-Johanna Tahk - Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, EstoniaOrcidhttps://orcid.org/0000-0002-4566-9192
    • Tõnis Laasfeld - Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, Estonia
    • Santa Veikšina - Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, Estonia
    • Carina Höring - Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, GermanyPresent Address: Biontech SE, An der Goldgrube 12, 55131 Mainz, Germany
    • Albert O. Gattor - Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, GermanyOrcidhttps://orcid.org/0000-0001-5166-2168
    • Laura J. Humphrys - Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, GermanyPresent Address: Monash Institute of Pharmaceutical Sciences, 399 Royal Parade, Parkville VIC 3052, AustraliaOrcidhttps://orcid.org/0000-0003-4019-5538
    • Christoph Müller - Institute of Pharmacy, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, GermanyPresent Address: Axolabs GmbH, Fritz-Hornschuch-Straße 9, 95326 Kulmbach, GermanyOrcidhttps://orcid.org/0000-0001-6121-6292
    • Nataliya Archipowa - Institute of Biophysics and Physical Biochemistry, Faculty of Biology and Preclinical Medicine, University of Regensburg, Universitätsstraße 31, D-93040 Regensburg, Germany
    • Johannes Köckenberger - Department of Chemistry and Pharmacy, Molecular and Clinical Pharmacy, Friedrich-Alexander-University Erlangen-Nürnberg, Nikolaus-Fiebiger-Straße 10, D-91058 Erlangen, GermanyPresent Address: Department of Chemistry and Biochemistry, University of California, La Jolla, San Diego, California 92093, USA
    • Markus R. Heinrich - Department of Chemistry and Pharmacy, Molecular and Clinical Pharmacy, Friedrich-Alexander-University Erlangen-Nürnberg, Nikolaus-Fiebiger-Straße 10, D-91058 Erlangen, GermanyOrcidhttps://orcid.org/0000-0001-7113-2025
    • Roger Jan Kutta - Institute of Physical and Theoretical Chemistry, Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstraße 31, D-93053 Regensburg, Germany
  • Author Contributions

    J.G., C.M., J.K., and L.G. performed the syntheses. J.G. performed stability studies and radiochemical binding assays. J.G., A.O.G., C.H., and L.J.H. conducted functional Y4R assays. C.M., N.A., and R.J.K. characterized the fluorescent ligands with respect to excitation and emission maxima and quantum yields. J.G., M-J.T., S.K., and T.L. performed fluorescent anisotropy assays. S.K. performed TIRF microscopy. S.V. prepared the BBVs. L.G. and S.V. performed molecular cloning. C.H. and J.G. prepared the stably transfected HEK293T cell line. M.K. initiated and planned the project. M.K., M.R.H., A.R., and R.J.K. supervised the research. J.G. and M.K. with impact from all coauthors wrote the manuscript. All authors have given approval to the final version of the manuscript.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

The authors thank Susanne Bollwein, Maria Beer-Krön, Brigitte Wenzl, Carmen Piirsalu, and Merve Gökalp for excellent technical assistance, and Sabrina Diwisch, Luise Liebig, and Tobias Spitzl for support with the flow cytometric assays. This work was funded by the Deutsche Forschungsgemeinschaft (DFG) (research grant KE 1857/1-3) and the Estonian Research Council grant PSG230 and was additionally supported by the Graduate Training Program GRK 1910 of the DFG.

Abbreviations Used

Click to copy section linkSection link copied!

Arg(carb)

Nω-carbamoylated arginine containing a tetramethylene spacer or a hexyne structure with a terminal alkyne group in the carbamoyl substituent (cf. Figure 3)

BSA

bovine serum albumin

CAMYEN

cAMP sensor using YFP-Epac-nanoluciferase

CHO

Chinese hamster ovary

Cy3B-SE

Cy3B succinimidyl ester

DIPEA

N,N-diisopropylethylamine

Emax

efficacy (maximum effect) of a receptor agonist determined in a functional assay

Epac

exchange protein activated by cAMP

EtOH

ethanol

FA

fluorescence anisotropy

FBS

fetal bovine serum

FL

fluorescent ligand

Fmoc amino acid

Nα-Fmoc-protected and side-chain-protected (if required) amino acid

HBTU

3-[bis(dimethylamino)methyliumyl]-3H-benzotriazol-1-oxide hexafluorophosphate

HEC

human endometrial cancer

HEPES

4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid

HOBt

hydroxybenzotriazole

hPP

human pancreatic polypeptide

hY4R

human Y4 receptor

k

retention (or capacity) factor (HPLC)

MeOH

methanol

Nluc

nanoluciferase

Pbf

2,2,4,6,7-pentamethyl-dihydrobenzofuran-5-sulfonyl

PBS

phosphate-buffered saline

pEC50

negative decadic logarithm of the half-maximal effective concentration (functional assays)

pKi

negative decadic logarithm of the dissociation constant Ki (in M) obtained from a competition binding experiment

pNPY

porcine neuropeptide Y

PS

polystyrene

PyBOP

benzotriazol-1-yl-oxytripyrrolidinophosphonium-hexafluorophosphate

PYY

peptide YY

RP

reversed phase

SEM

standard error of the mean

SPPS

solid-phase peptide synthesis

TFA

trifluoroacetic acid

TIRF

total internal reflection fluorescence

tR

retention time

YFP

yellow fluorescent protein

YR

NPY receptor

References

Click to copy section linkSection link copied!

This article references 68 other publications.

  1. 1
    Michel, M. C.; Beck-Sickinger, A.; Cox, H.; Doods, H. N.; Herzog, H.; Larhammar, D.; Quirion, R.; Schwartz, T.; Westfall, T. XVI. International Union of Pharmacology recommendations for the nomenclature of neuropeptide Y, peptide YY, and pancreatic polypeptide receptors. Pharmacol. Rev. 1998, 50, 143150
  2. 2
    Pedragosa-Badia, X.; Stichel, J.; Beck-Sickinger, A. G. Neuropeptide Y receptors: how to get subtype selectivity. Front. Endocrinol. 2013, 4, 5  DOI: 10.3389/fendo.2013.00005
  3. 3
    Ueno, N.; Inui, A.; Iwamoto, M.; Kaga, T.; Asakawa, A.; Okita, M.; Fujimiya, M.; Nakajima, Y.; Ohmoto, Y.; Ohnaka, M.; Nakaya, Y.; Miyazaki, J.-I.; Kasuga, M. Decreased food intake and body weight in pancreatic polypeptide-overexpressing mice. Gastroenterology 1999, 117, 14271432,  DOI: 10.1016/S0016-5085(99)70293-3
  4. 4
    Batterham, R. L.; Le Roux, C. W.; Cohen, M. A.; Park, A. J.; Ellis, S. M.; Patterson, M.; Frost, G. S.; Ghatei, M. A.; Bloom, S. R. Pancreatic polypeptide reduces appetite and food intake in humans. J. Clin. Endocrinol. Metab. 2003, 88, 39893992,  DOI: 10.1210/jc.2003-030630
  5. 5
    Schmidt, P. T.; Näslund, E.; Grybäck, P.; Jacobsson, H.; Holst, J. J.; Hilsted, L.; Hellström, P. M. A role for pancreatic polypeptide in the regulation of gastric emptying and short-term metabolic control. J. Clin. Endocrinol. Metab. 2005, 90, 52415246,  DOI: 10.1210/jc.2004-2089
  6. 6
    Painsipp, E.; Wultsch, T.; Edelsbrunner, M. E.; Tasan, R. O.; Singewald, N.; Herzog, H.; Holzer, P. Reduced anxiety-like and depression-related behavior in neuropeptide Y Y4 receptor knockout mice. Genes, Brain Behav. 2008, 7, 532542,  DOI: 10.1111/j.1601-183X.2008.00389.x
  7. 7
    Lin, S.; Shi, Y.-C.; Yulyaningsih, E.; Aljanova, A.; Zhang, L.; Macia, L.; Nguyen, A. D.; Lin, E.-J. D.; During, M. J.; Herzog, H.; Sainsbury, A. Critical role of arcuate Y4 receptors and the melanocortin system in pancreatic polypeptide-induced reduction in food intake in mice. PLoS One 2009, 4, e8488  DOI: 10.1371/journal.pone.0008488
  8. 8
    Tasan, R. O.; Lin, S.; Hetzenauer, A.; Singewald, N.; Herzog, H.; Sperk, G. Increased novelty-induced motor activity and reduced depression-like behavior in neuropeptide Y (NPY)-Y4 receptor knockout mice. Neuroscience 2009, 158, 17171730,  DOI: 10.1016/j.neuroscience.2008.11.048
  9. 9
    Sainsbury, A.; Shi, Y. C.; Zhang, L.; Aljanova, A.; Lin, Z.; Nguyen, A. D.; Herzog, H.; Lin, S. Y4 receptors and pancreatic polypeptide regulate food intake via hypothalamic orexin and brain-derived neurotropic factor dependent pathways. Neuropeptides 2010, 44, 261268,  DOI: 10.1016/j.npep.2010.01.001
  10. 10
    Li, J. B.; Asakawa, A.; Terashi, M.; Cheng, K.; Chaolu, H.; Zoshiki, T.; Ushikai, M.; Sheriff, S.; Balasubramaniam, A.; Inui, A. Regulatory effects of Y4 receptor agonist (BVD-74D) on food intake. Peptides 2010, 31, 17061710,  DOI: 10.1016/j.peptides.2010.06.011
  11. 11
    Moriya, R.; Fujikawa, T.; Ito, J.; Shirakura, T.; Hirose, H.; Suzuki, J.; Fukuroda, T.; MacNeil, D. J.; Kanatani, A. Pancreatic polypeptide enhances colonic muscle contraction and fecal output through neuropeptide Y Y4 receptor in mice. Eur. J. Pharmacol. 2010, 627, 258264,  DOI: 10.1016/j.ejphar.2009.09.057
  12. 12
    Brothers, S. P.; Wahlestedt, C. Therapeutic potential of neuropeptide Y (NPY) receptor ligands. EMBO Mol. Med. 2010, 2, 429439,  DOI: 10.1002/emmm.201000100
  13. 13
    Verma, D.; Hörmer, B.; Bellmann-Sickert, K.; Thieme, V.; Beck-Sickinger, A. G.; Herzog, H.; Sperk, G.; Tasan, R. O. Pancreatic polypeptide and its central Y4 receptors are essential for cued fear extinction and permanent suppression of fear. Br. J. Pharmacol. 2016, 173, 19251938,  DOI: 10.1111/bph.13456
  14. 14
    Zhang, L.; Bijker, M. S.; Herzog, H. The neuropeptide Y system: Pathophysiological and therapeutic implications in obesity and cancer. Pharmacol. Ther. 2011, 131, 91113,  DOI: 10.1016/j.pharmthera.2011.03.011
  15. 15
    Yi, M.; Li, H.; Wu, Z.; Yan, J.; Liu, Q.; Ou, C.; Chen, M. A promising therapeutic target for metabolic diseases: Neuropeptide Y receptors in humans. Cell. Physiol. Biochem. 2018, 45, 88107,  DOI: 10.1159/000486225
  16. 16
    Allen, M.; Reeves, J.; Mellor, G. High throughput fluorescence polarization: a homogeneous alternative to radioligand binding for cell surface receptors. J. Biomol. Screening 2000, 5, 6369,  DOI: 10.1177/108705710000500202
  17. 17
    Leyris, J.-P.; Roux, T.; Trinquet, E.; Verdié, P.; Fehrentz, J.-A.; Oueslati, N.; Douzon, S.; Bourrier, E.; Lamarque, L.; Gagne, D. Homogeneous time-resolved fluorescence-based assay to screen for ligands targeting the growth hormone secretagogue receptor type 1a. Anal. Biochem. 2011, 408, 253262,  DOI: 10.1016/j.ab.2010.09.030
  18. 18
    Emami-Nemini, A.; Roux, T.; Leblay, M.; Bourrier, E.; Lamarque, L.; Trinquet, E.; Lohse, M. J. Time-resolved fluorescence ligand binding for G protein–coupled receptors. Nat. Protoc. 2013, 8, 13071320,  DOI: 10.1038/nprot.2013.073
  19. 19
    Veiksina, S.; Kopanchuk, S.; Rinken, A. Budded baculoviruses as a tool for a homogeneous fluorescence anisotropy-based assay of ligand binding to G protein-coupled receptors: The case of melanocortin 4 receptors. Biochim. Biophys. Acta, Biomembr. 2014, 1838, 372381,  DOI: 10.1016/j.bbamem.2013.09.015
  20. 20
    Schiele, F.; Ayaz, P.; Fernández-Montalván, A. A universal homogeneous assay for high-throughput determination of binding kinetics. Anal. Biochem. 2015, 468, 4249,  DOI: 10.1016/j.ab.2014.09.007
  21. 21
    Antoine, T.; Ott, D.; Ebell, K.; Hansen, K.; Henry, L.; Becker, F.; Hannus, S. Homogeneous time-resolved G protein-coupled receptor-ligand binding assay based on fluorescence cross-correlation spectroscopy. Anal. Biochem. 2016, 502, 2435,  DOI: 10.1016/j.ab.2016.02.017
  22. 22
    Stoddart, L. A.; White, C. W.; Nguyen, K.; Hill, S. J.; Pfleger, K. D. Fluorescence- and bioluminescence-based approaches to study GPCR ligand binding. Br. J. Pharmacol. 2016, 173, 30283037,  DOI: 10.1111/bph.13316
  23. 23
    Stoddart, L. A.; Kilpatrick, L. E.; Hill, S. J. NanoBRET approaches to study ligand binding to GPCRs and RTKs. Trends Pharmacol. Sci. 2018, 39, 136147,  DOI: 10.1016/j.tips.2017.10.006
  24. 24
    Iliopoulos-Tsoutsouvas, C.; Kulkarni, R. N.; Makriyannis, A.; Nikas, S. P. Fluorescent probes for G-protein-coupled receptor drug discovery. Expert Opin. Drug Discovery 2018, 13, 933947,  DOI: 10.1080/17460441.2018.1518975
  25. 25
    Ziemek, R.; Schneider, E.; Kraus, A.; Cabrele, C.; Beck-Sickinger, A. G.; Bernhardt, G.; Buschauer, A. Determination of affinity and activity of ligands at the human neuropeptide Y Y4 receptor by flow cytometry and aequorin luminescence. J. Recept. Signal Transduction 2007, 27, 217233,  DOI: 10.1080/10799890701505206
  26. 26
    Dukorn, S.; Littmann, T.; Keller, M.; Kuhn, K.; Cabrele, C.; Baumeister, P.; Bernhardt, G.; Buschauer, A. Fluorescence- and radiolabeling of [Lys4,Nle17,30]hPP yields molecular tools for the NPY Y4 receptor. Bioconjugate Chem. 2017, 28, 12911304,  DOI: 10.1021/acs.bioconjchem.7b00103
  27. 27
    Tang, T.; Tan, Q.; Han, S.; Diemar, A.; Löbner, K.; Wang, H.; Schüß, C.; Behr, V.; Mörl, K.; Wang, M.; Chu, X.; Yi, C.; Keller, M.; Kofoed, J.; Reedtz-Runge, S.; Kaiser, A.; Beck-Sickinger, A. G.; Zhao, Q.; Wu, B. Receptor-specific recognition of NPY peptides revealed by structures of NPY receptors. Sci. Adv. 2022, 8, eabm1232  DOI: 10.1126/sciadv.abm1232
  28. 28
    Kuhn, K. K.; Littmann, T.; Dukorn, S.; Tanaka, M.; Keller, M.; Ozawa, T.; Bernhardt, G.; Buschauer, A. In search of NPY Y4R antagonists: Incorporation of carbamoylated arginine, aza-amino acids, or D-amino acids into oligopeptides derived from the C-termini of the endogenous agonists. ACS Omega 2017, 2, 36163631,  DOI: 10.1021/acsomega.7b00451
  29. 29
    Spinnler, K.; von Krüchten, L.; Konieczny, A.; Schindler, L.; Bernhardt, G.; Keller, M. An alkyne-functionalized arginine for solid-phase synthesis enabling “bioorthogonal” peptide conjugation. ACS Med. Chem. Lett. 2020, 11, 334339,  DOI: 10.1021/acsmedchemlett.9b00388
  30. 30
    Konieczny, A.; Conrad, M.; Ertl, F. J.; Gleixner, J.; Gattor, A. O.; Grätz, L.; Schmidt, M. F.; Neu, E.; Horn, A. H. C.; Wifling, D.; Gmeiner, P.; Clark, T.; Sticht, H.; Keller, M. N-terminus to arginine side-chain cyclization of linear peptidic neuropeptide Y Y4 receptor ligands results in picomolar binding constants. J. Med. Chem. 2021, 64, 1674616769,  DOI: 10.1021/acs.jmedchem.1c01574
  31. 31
    Gleixner, J.; Gattor, A. O.; Humphrys, L. J.; Brunner, T.; Keller, M. 3H]UR-JG102 - a radiolabeled cyclic peptide with high affinity and excellent selectivity for the neuropeptide Y Y4 receptor. J. Med. Chem. 2023, 66, 1378813808,  DOI: 10.1021/acs.jmedchem.3c01224
  32. 32
    Keller, M.; Kuhn, K. K.; Einsiedel, J.; Hübner, H.; Biselli, S.; Mollereau, C.; Wifling, D.; Svobodová, J.; Bernhardt, G.; Cabrele, C.; Vanderheyden, P. M. L.; Gmeiner, P.; Buschauer, A. Mimicking of arginine by functionalized Nω-carbamoylated arginine as a new broadly applicable approach to labeled bioactive peptides: high affinity angiotensin, neuropeptide Y, neuropeptide FF, and neurotensin receptor ligands as examples. J. Med. Chem. 2016, 59, 19251945,  DOI: 10.1021/acs.jmedchem.5b01495
  33. 33
    Keller, M.; Weiss, S.; Hutzler, C.; Kuhn, K. K.; M?llereau, C.; Dukorn, S.; Schindler, L.; Bernhardt, G.; König, B.; Buschauer, A. Nω-carbamoylation of the argininamide moiety: An avenue to insurmountable NPY Y1 receptor antagonists and a radiolabeled selective high-affinity molecular tool ([3H]UR-MK299) with extended residence time. J. Med. Chem. 2015, 58, 88348849,  DOI: 10.1021/acs.jmedchem.5b00925
  34. 34
    Konieczny, A.; Braun, D.; Wifling, D.; Bernhardt, G.; Keller, M. Oligopeptides as neuropeptide Y Y4 receptor ligands: identification of a high-affinity tetrapeptide agonist and a hexapeptide antagonist. J. Med. Chem. 2020, 63, 81988215,  DOI: 10.1021/acs.jmedchem.0c00426
  35. 35
    Kuhn, K. K.; Ertl, T.; Dukorn, S.; Keller, M.; Bernhardt, G.; Reiser, O.; Buschauer, A. High affinity agonists of the neuropeptide Y (NPY) Y4 receptor derived from the C-terminal pentapeptide of human pancreatic polypeptide (hPP): Synthesis, stereochemical discrimination, and radiolabeling. J. Med. Chem. 2016, 59, 60456058,  DOI: 10.1021/acs.jmedchem.6b00309
  36. 36
    Berlicki, Ł.; Kaske, M.; Gutiérrez-Abad, R.; Bernhardt, G.; Illa, O.; Ortuño, R. M.; Cabrele, C.; Buschauer, A.; Reiser, O. Replacement of Thr32 and Gln34 in the C-terminal neuropeptide Y fragment 25–36 by cis-cyclobutane and cis-cyclopentane β-amino acids shifts selectivity toward the Y4 receptor. J. Med. Chem. 2013, 56, 84228431,  DOI: 10.1021/jm4008505
  37. 37
    Wirth, U.; Erl, J.; Azzam, S.; Höring, C.; Skiba, M.; Singh, R.; Hochmuth, K.; Keller, M.; Wegener, J.; König, B. Monitoring the reversibility of GPCR signaling by combining photochromic ligands with label-free impedance analysis. Angew. Chem., Int. Ed. 2023, 62, e202215547  DOI: 10.1002/anie.202215547
  38. 38
    Archipowa, N.; Wittmann, L.; Köckenberger, J.; Ertl, F. J.; Gleixner, J.; Keller, M.; Heinrich, M. R.; Kutta, R. J. Characterization of fluorescent dyes frequently used for bioimaging: Photophysics and photocatalytical reactions with proteins. J. Phys. Chem. B 2023, 127, 95329542,  DOI: 10.1021/acs.jpcb.3c04484
  39. 39
    Keller, M.; Mahuroof, S. A.; Yee, V. H.; Carpenter, J.; Schindler, L.; Littmann, T.; Pegoli, A.; Hübner, H.; Bernhardt, G.; Gmeiner, P.; Holliday, N. D. Fluorescence labeling of neurotensin(8–13) via arginine residues gives molecular tools with high receptor affinity. ACS Med. Chem. Lett. 2020, 11, 1622,  DOI: 10.1021/acsmedchemlett.9b00462
  40. 40
    Müller, C.; Gleixner, J.; Tahk, M.-J.; Kopanchuk, S.; Laasfeld, T.; Weinhart, M.; Schollmeyer, D.; Betschart, M. U.; Lüdeke, S.; Koch, P.; Rinken, A.; Keller, M. Structure-based design of high-affinity fluorescent probes for the neuropeptide Y Y1 receptor. J. Med. Chem. 2022, 65, 48324853,  DOI: 10.1021/acs.jmedchem.1c02033
  41. 41
    Keller, M.; Erdmann, D.; Pop, N.; Pluym, N.; Teng, S.; Bernhardt, G.; Buschauer, A. Red-fluorescent argininamide-type NPY Y1 receptor antagonists as pharmacological tools. Biorg. Med. Chem. 2011, 19, 28592878,  DOI: 10.1016/j.bmc.2011.03.045
  42. 42
    She, X.; Pegoli, A.; Gruber, C. G.; Wifling, D.; Carpenter, J.; Hübner, H.; Chen, M.; Wan, J.; Bernhardt, G.; Gmeiner, P.; Holliday, N. D.; Keller, M. Red-emitting dibenzodiazepinone derivatives as fluorescent dualsteric probes for the muscarinic acetylcholine M2 receptor. J. Med. Chem. 2020, 63, 41334154,  DOI: 10.1021/acs.jmedchem.9b02172
  43. 43
    Katritch, V.; Fenalti, G.; Abola, E. E.; Roth, B. L.; Cherezov, V.; Stevens, R. C. Allosteric sodium in class A GPCR signaling. Trends Biochem. Sci. 2014, 39, 233244,  DOI: 10.1016/j.tibs.2014.03.002
  44. 44
    White, K. L.; Eddy, M. T.; Gao, Z. G.; Han, G. W.; Lian, T.; Deary, A.; Patel, N.; Jacobson, K. A.; Katritch, V.; Stevens, R. C. Structural connection between activation microswitch and allosteric sodium site in GPCR Signaling. Structure 2018, 26, 259269,  DOI: 10.1016/j.str.2017.12.013
  45. 45
    DeVree, B. T.; Mahoney, J. P.; Velez-Ruiz, G. A.; Rasmussen, S. G.; Kuszak, A. J.; Edwald, E.; Fung, J. J.; Manglik, A.; Masureel, M.; Du, Y.; Matt, R. A.; Pardon, E.; Steyaert, J.; Kobilka, B. K.; Sunahara, R. K. Allosteric coupling from G protein to the agonist-binding pocket in GPCRs. Nature 2016, 535, 182186,  DOI: 10.1038/nature18324
  46. 46
    Wingler, L. M.; Lefkowitz, R. J. Conformational basis of G protein-coupled receptor signaling versatility. Trends Cell Biol. 2020, 30, 736747,  DOI: 10.1016/j.tcb.2020.06.002
  47. 47
    Copeland, R. A. Conformational adaptation in drug–target interactions and residence time. Future Med. Chem. 2011, 3, 14911501,  DOI: 10.4155/fmc.11.112
  48. 48
    Vauquelin, G. Simplified models for heterobivalent ligand binding: when are they applicable and which are the factors that affect their target residence time. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2013, 386, 949962,  DOI: 10.1007/s00210-013-0881-0
  49. 49
    Palmberger, D.; Wilson, I. B.; Berger, I.; Grabherr, R.; Rendic, D. SweetBac: a new approach for the production of mammalianised glycoproteins in insect cells. PLoS One 2012, 7, e34226  DOI: 10.1371/journal.pone.0034226
  50. 50
    Schneider, E. H.; Seifert, R. Sf9 cells: A versatile model system to investigate the pharmacological properties of G protein-coupled receptors. Pharmacol. Ther. 2010, 128, 387418,  DOI: 10.1016/j.pharmthera.2010.07.005
  51. 51
    Grätz, L.; Müller, C.; Pegoli, A.; Schindler, L.; Bernhardt, G.; Littmann, T. Insertion of Nanoluc into the extracellular loops as a complementary method to establish BRET-based binding assays for GPCRs. ACS Pharmacol. Transl. Sci. 2022, 5, 11421155,  DOI: 10.1021/acsptsci.2c00162
  52. 52
    Calderón, J. C.; Plut, E.; Keller, M.; Cabrele, C.; Reiser, O.; Gervasio, F. L.; Clark, T. Extended metadynamics protocol for binding/unbinding free energies of peptide ligands to class A G-protein-coupled receptors. J. Chem. Inf. Model. 2024, 64, 205218,  DOI: 10.1021/acs.jcim.3c01574
  53. 53
    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.; Wood, K. V. Engineered luciferase reporter from a deep sea shrimp utilizing a novel imidazopyrazinone substrate. ACS Chem. Biol. 2012, 7, 18481857,  DOI: 10.1021/cb3002478
  54. 54
    Keller, M.; Kaske, M.; Holzammer, T.; Bernhardt, G.; Buschauer, A. Dimeric argininamide-type neuropeptide Y receptor antagonists: Chiral discrimination between Y1 and Y4 receptors. Biorg. Med. Chem. 2013, 21, 63036322,  DOI: 10.1016/j.bmc.2013.08.065
  55. 55
    Pegoli, A.; She, X.; Wifling, D.; Hubner, H.; Bernhardt, G.; Gmeiner, P.; Keller, M. Radiolabeled dibenzodiazepinone-type antagonists give evidence of dualsteric binding at the M(2) muscarinic acetylcholine receptor. J. Med. Chem. 2017, 60, 33143334,  DOI: 10.1021/acs.jmedchem.6b01892
  56. 56
    Schindler, L.; Moosbauer, J.; Schmidt, D.; Spruss, T.; Grätz, L.; Lüdeke, S.; Hofheinz, F.; Meister, S.; Echtenacher, B.; Bernhardt, G.; Pietzsch, J.; Hellwig, D.; Keller, M. Development of a neurotensin-derived 68Ga-labeled PET ligand with high in vivo stability for imaging of NTS1 receptor-expressing tumors. Cancers 2022, 14, 4922  DOI: 10.3390/cancers14194922
  57. 57
    Schindler, L.; Wohlfahrt, K.; von Krüchten, L. G.; Prante, O.; Keller, M.; Maschauer, S. Neurotensin analogs by fluoroglycosylation at Nω-carbamoylated arginines for PET imaging of NTS1-positive tumors. Sci. Rep. 2022, 12, 15028  DOI: 10.1038/s41598-022-19296-0
  58. 58
    Höfelschweiger, B. K. The pyrylium dyes: A new class of biolabels. Synthesis, spectroscopy, and application as labels and in general protein assay. Doctoral thesis, University of Regensburg, Regensburg2005.
  59. 59
    Moser, C.; Bernhardt, G.; Michel, J.; Schwarz, H.; Buschauer, A. Cloning and functional expression of the hNPY Y5 receptor in human endometrial cancer (HEC-1B) cells. Can. J. Physiol. Pharmacol. 2000, 78, 134142,  DOI: 10.1139/y99-125
  60. 60
    Bartole, E.; Grätz, L.; Littmann, T.; Wifling, D.; Seibel, U.; Buschauer, A.; Bernhardt, G. UR-DEBa242: A Py-5-labeled fluorescent multipurpose probe for investigations on the histamine H3 and H4 receptors. J. Med. Chem. 2020, 63, 52975311,  DOI: 10.1021/acs.jmedchem.0c00160
  61. 61
    Mansouri, M.; Bellon-Echeverria, I.; Rizk, A.; Ehsaei, Z.; Cosentino, C. C.; Silva, C. S.; Xie, Y.; Boyce, F. M.; Davis, M. W.; Neuhauss, S. C. F.; Taylor, V.; Ballmer-Hofer, K.; Berger, I.; Berger, P. Highly efficient baculovirus-mediated multigene delivery in primary cells. Nat. Commun. 2016, 7, 11529  DOI: 10.1038/ncomms11529
  62. 62
    Nie, Y.; Chaillet, M.; Becke, C.; Haffke, M.; Pelosse, M.; Fitzgerald, D.; Collinson, I.; Schaffitzel, C.; Berger, I. ACEMBL tool-kits for high-throughput multigene delivery and expression in prokaryotic and eukaryotic hosts. Adv. Exp. Med. Biol. 2016, 896, 2742,  DOI: 10.1007/978-3-319-27216-0_3
  63. 63
    Cheng, Y.-C.; Prusoff, W. H. Relationship between the inhibition constant (KI) and the concentration of inhibitor which causes 50% inhibition (IC50) of an enzymatic reaction. Biochem. Pharmacol. 1973, 22, 30993108,  DOI: 10.1016/0006-2952(73)90196-2
  64. 64
    Laasfeld, T.; Kopanchuk, S.; Rinken, A. Image-based cell-size estimation for baculovirus quantification. BioTechniques 2017, 63, 161168,  DOI: 10.2144/000114595
  65. 65
    Veiksina, S.; Tahk, M.-J.; Laasfeld, T.; Link, R.; Kopanchuk, S.; Rinken, A. Fluorescence Anisotropy-Based Assay for Characterization of Ligand Binding Dynamics to GPCRs: The Case of Cy3B-Labeled Ligands Binding to MC4 Receptors in Budded Baculoviruses. In G Protein-Coupled Receptor Screening Assays: Methods and Protocols; Martins, S. A. M.; Prazeres, D. M. F., Eds.; Springer: US, New York, NY, 2021; pp 119136.
  66. 66
    Kopanchuk, S.; Vavers, E.; Veiksina, S.; Ligi, K.; Zvejniece, L.; Dambrova, M.; Rinken, A. Intracellular dynamics of the Sigma-1 receptor observed with super-resolution imaging microscopy. PLoS One 2022, 17, e0268563  DOI: 10.1371/journal.pone.0268563
  67. 67
    Soulez, F. In A “Learn 2D, Apply 3D” Method for 3D Deconvolution Microscopy, 2014 IEEE 11th International Symposium on Biomedical Imaging (ISBI); IEEE, 2014; pp 10751078.
  68. 68
    Chenouard, N.; Bloch, I.; Olivo-Marin, J. C. Multiple hypothesis tracking for cluttered biological image sequences. IEEE Trans. Pattern Anal. Mach. Intell. 2013, 35, 27363750,  DOI: 10.1109/TPAMI.2013.97

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 3 publications.

  1. Fabian J. Ertl, Sergei Kopanchuk, Nicola C. Dijon, Santa Veikšina, Maris-Johanna Tahk, Tõnis Laasfeld, Franziska Schettler, Albert O. Gattor, Harald Hübner, Nataliya Archipowa, Johannes Köckenberger, Markus R. Heinrich, Peter Gmeiner, Roger J. Kutta, Nicholas D. Holliday, Ago Rinken, Max Keller. Dually Labeled Neurotensin NTS1R Ligands for Probing Radiochemical and Fluorescence-Based Binding Assays. Journal of Medicinal Chemistry 2024, 67 (18) , 16664-16691. https://doi.org/10.1021/acs.jmedchem.4c01470
  2. Renáta Szabó, Ágnes Hornyánszky, Dóra Judit Kiss, György Miklós Keserű. Fluorescent tools for imaging class A G-protein coupled receptors. European Journal of Pharmaceutical Sciences 2025, 209 , 107074. https://doi.org/10.1016/j.ejps.2025.107074
  3. Max E. Huber, Silas L. Wurnig, Aurélien F. A. Moumbock, Lara Toy, Evi Kostenis, Ana Alonso Bartolomé, Martyna Szpakowska, Andy Chevigné, Stefan Günther, Finn K. Hansen, Matthias Schiedel. Development of a NanoBRET Assay Platform to Detect Intracellular Ligands for the Chemokine Receptors CCR6 and CXCR1. ChemMedChem 2024, 19 (20) https://doi.org/10.1002/cmdc.202400284

ACS Pharmacology & Translational Science

Cite this: ACS Pharmacol. Transl. Sci. 2024, 7, 4, 1142–1168
Click to copy citationCitation copied!
https://doi.org/10.1021/acsptsci.4c00013
Published March 20, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

1426

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. (A) Structures and Y4R affinities of reported Y4R fluorescent ligands. 1 (26) and 2 (27) represent derivatives of the endogenous Y4R ligand hPP labeled with sulfo-Cy5 (S0223) and 5-TAMRA, respectively. Compound 3 (29) represents a derivative of the hexapeptide UR-KK236 (28) labeled with Sulfo-Cy5.5 (lumiprobe, ref no. 7330). (B) Structures of the reported cyclic Y4R ligands 47 showing high binding affinity to Y4R (pKi > 10). (30,31)

    Figure 2

    Figure 2. Study design of the present work: synthesis and characterization of fluorescent probes with high Y4R binding affinity derived from recently reported cyclic hexapeptides (6, (31) 4 (30)).

    Figure 3

    Figure 3. Structures of the reported Nω-carbamoylated arginines 8 (32) and 9 (29) used in SPPS for the preparation of precursor peptide 11.

    Scheme 1

    Scheme 1. Synthesis of the Cyclic Peptide 11 via SPPSa

    aReagents and conditions: (a) Fmoc amino acid/HOBt/HBTU/DIPEA (5/5/4.9/10 equiv), DMF/NMP 8:2 v/v, 38 °C, 2 × 45 min (“double” coupling); Fmoc deprotection (following amino acid coupling): 20% piperidine in DMF/NMP 8:2 v/v, rt, 2 × 10 min; (b) building block 8 or 9/HOBt/HBTU/DIPEA (3/3/2.9/6 equiv), DMF/NMP 8:2 v/v, 38 °C, 16 h (single coupling); Fmoc deprotection (following amino acid coupling): as under (a); (c) succinic anhydride/DIPEA (10/10 equiv), DMF/NMP 8:2 v/v, 38 °C, 45 min; (d) (1) TFA/CH2Cl2 1:3 v/v, rt, 2 × 20 min; (2) TFA/H2O 95:5 v/v, rt, 5 h; overall yield: 37%; (e) peptide cyclization: HOBt/PyBOP/DIPEA (3/2.9/6 equiv), DMF/NMP 8:2 v/v, rt, 16 h, 21%.

    Scheme 2

    Scheme 2. Synthesis of the Fluorescent Y4R Ligands 1619a

    aReagents and conditions: (a) DIPEA, DMF, rt, 2 h, 34–69%; (b) CuSO4, sodium ascorbate, H2O/NMP 1:1 v/v, rt, 2 h, 27%.

    Figure 4

    Figure 4. Flow cytometric saturation binding of 16 (A), 17 (B), and 18 (C) at whole CHO-hY4R-Gqi5-mtAEQ cells at 22 ± 2 °C. (A) Representative saturation isotherms (red circle) of 16 from experiments performed in sodium-free buffer (buffer I) and sodium-containing buffer (DPBS, 137 mM Na+). (B, C) Representative saturation isotherms (red circle) of 17 (B) and 18 (C) from experiments performed in sodium-containing buffer (DPBS). Unspecific binding (blue squares) was determined in the presence of 1 μM hPP (A–C). Total and unspecific binding data represent mean values ± SEM. Specific binding, representing calculated values ± propagated error, were fitted according to an equation describing a hyperbolic isotherm (binding-saturation: one site-specific binding, GraphPad Prism 5).

    Figure 5

    Figure 5. Binding kinetics of 1618 determined by flow cytometry at whole CHO-hY4R-Gqi5-mtAEQ cells at 22 ± 2 °C. (A) Association of 16 to the hY4R under sodium-free conditions (buffer I) and in sodium-containing buffer (DPBS, 137 mM Na+). Concentration of 16: 0.3 and 0.7 nM, respectively. Proportion of the initial fast association (two-phase association fit, GraphPad Prism 5): 38 ± 4% (mean ± SEM). (B) Dissociation of 16 from the hY4R determined in buffer I and DPBS. The dissociation was initiated after 1.5 h of preincubation with 16 (c = 1.5 nM (Na+-free) and 3.5 nM (137 mM Na+)) by the addition of an excess of hPP (1000-fold) and 5 (100-fold). Plateau values of the three-parameter fits (monophasic exponential decline): 13% (Na+-free), 22% (137 mM Na+). (C) Association of 17 (c = 1 nM) and 18 (c = 0.5 nM) to the hY4R determined in DPBS (137 mM Na+). Proportion of the initial fast association (two-phase association fit, GraphPad Prism 5): 74 ± 1% (17), 29 ± 3% (18) (mean values ± SEM). (D) Dissociation of 17 and 18 from the hY4R in DPBS. The dissociation was initiated after 1.5 h of preincubation with 17 (c = 5 nM) or 18 (c = 2.5 nM) by the addition of an excess of hPP (1000-fold) and 5 (100-fold). Plateau values of the three-parameter fits (monophasic exponential decline): 13% (17), 22% (18). Data (A–D) represent mean values ± SEM from three independent experiments performed in duplicate.

    Figure 6

    Figure 6. Binding of 16 to hY4R displaying BBVs studied by FA measurement at 27 °C. (A) Binding isotherms of 16 obtained from experiments using fixed concentrations of 16 (0.5 or 2 nM) and increasing amounts of Y4R. Total binding is represented by filled symbols and unspecific binding (determined in the presence of 1 μM hPP) is represented by open symbols. Depicted data (mean values ± SEM from a representative experiment performed in duplicate) represent snapshots at 90 min incubation. Y4R concentrations displayed on the abscissa were calculated after global analysis of the data from three or four individual experiments by a modified version of a model described by Veiksina et al., (19) affording the estimated binding site (Y4R) concentration of the applied BBV stock which amounted to 6 ± 1 nM (Y4Rnonglyco, mean value ± SEM, n = 3) or 2.1 ± 0.1 nM (Y4RSwBac, mean value ± SEM, n = 4). (B) Association and dissociation of 16 (0.5 nM) determined in real time for three different Y4R concentrations (green, blue, and red symbols). Total binding is represented by filled symbols and unspecific binding (determined in the presence of 1 μM hPP) is represented by open symbols. Data represent mean values ± SEM from a representative experiment performed in duplicate.

    Figure 7

    Figure 7. Characterization of Y4R binding of fluorescent ligand 16 in a NanoBRET-based binding assay at 25 °C using intact HEK293T-hY4R-NLuc(intraECL2) cells. (A) Representative saturation isotherm (specific binding) from saturation binding experiments. Unspecific binding was determined in the presence of 1 μM 5. Total and unspecific binding data represent mean values ± SEM. Specific binding data represent calculated values ± propagated error. (B) Association of 16 (c = 1 nM) to Y4R. Mean values ± SEM from three independent experiments performed in triplicate. (C) Dissociation of 16 from Y4R. The dissociation was initiated after 1.5 h of preincubation with 16 (c = 3.5 nM) by the addition of a 1000-fold excess of 5. Mean values ± SEM from three independent experiments performed in duplicate. Plateau value of the three-parameter fit describing a monophasic exponential decline: 6% (note: for the biphasic fit the plateau value was not different from zero, see discussion).

    Figure 8

    Figure 8. Determination of Y4R affinities of Y4R reference ligands (hPP, 5, 7, UR-MK188, UR-MEK388, UR-KK200) in different fluorescence-based assays by competition binding with 16. (A) Displacement curves based on data from flow cytometric competition binding experiments performed with intact CHO-hY4R-Gqi5-mtAEQ cells in sodium-free buffer (buffer I) and sodium-containing buffer (DPBS, 137 mM Na+). (B) Displacement curves from fluorescence anisotropy-based competition binding experiments performed with Y4RSwBac-displaying BBVs in DPBS. (C) Displacement curves from NanoBRET-based competition binding experiments performed with intact HEK293T-hY4R-NLuc(intraECL2) cells in L15-HEPES (140 mM Na+). Data (A–C), representing mean values ± SEM from three to five independent experiments performed in duplicate (B) or triplicate (A, C), were fitted according to a four-parameter logistic model.

    Figure 9

    Figure 9. Visualization of fluorescent ligand (16, 17) binding to CHO-hY4R-Gqi5-mtAEQ cells by confocal microscopy. Shown are representative images acquired after incubation of the cells with 16 or 17 (each 20 nM) at 22 °C for 10 and 30 min. Unspecific binding was determined in the presence of 1 μM hPP. Nuclei were stained with H33342 (2 μM). Fluorescence of 16 and 17 is shown in green and red, respectively. Fluorescence of H33342 is shown in blue. Scale bar: 10 μm.

    Figure 10

    Figure 10. Visualization of binding of 16 to the hY4R transiently expressed by SK-OV-3 cells using wide-field and TIRF microscopy. (A) Wide-field fluorescence images acquired after incubation of the cells with 16 at 37 °C for 30 min. The two-color composite of individual focal planes after Z-stack deconvolution is shown with the green pseudocolor for 16 (561 nm excitation) and the blue pseudocolor for the nuclear stain channel (Hoechst 34580, 405 nm excitation). (B) Wide-field fluorescence and TIRF images of the same cells obtained after incubation of the cells with 16 (1 nM) at 37 °C for 30 min. Wide-field images were processed as under (A). In TIRF images, fluorescence of 16 is shown in white pseudocolor. Scale bar: 10 μm.

  • References


    This article references 68 other publications.

    1. 1
      Michel, M. C.; Beck-Sickinger, A.; Cox, H.; Doods, H. N.; Herzog, H.; Larhammar, D.; Quirion, R.; Schwartz, T.; Westfall, T. XVI. International Union of Pharmacology recommendations for the nomenclature of neuropeptide Y, peptide YY, and pancreatic polypeptide receptors. Pharmacol. Rev. 1998, 50, 143150
    2. 2
      Pedragosa-Badia, X.; Stichel, J.; Beck-Sickinger, A. G. Neuropeptide Y receptors: how to get subtype selectivity. Front. Endocrinol. 2013, 4, 5  DOI: 10.3389/fendo.2013.00005
    3. 3
      Ueno, N.; Inui, A.; Iwamoto, M.; Kaga, T.; Asakawa, A.; Okita, M.; Fujimiya, M.; Nakajima, Y.; Ohmoto, Y.; Ohnaka, M.; Nakaya, Y.; Miyazaki, J.-I.; Kasuga, M. Decreased food intake and body weight in pancreatic polypeptide-overexpressing mice. Gastroenterology 1999, 117, 14271432,  DOI: 10.1016/S0016-5085(99)70293-3
    4. 4
      Batterham, R. L.; Le Roux, C. W.; Cohen, M. A.; Park, A. J.; Ellis, S. M.; Patterson, M.; Frost, G. S.; Ghatei, M. A.; Bloom, S. R. Pancreatic polypeptide reduces appetite and food intake in humans. J. Clin. Endocrinol. Metab. 2003, 88, 39893992,  DOI: 10.1210/jc.2003-030630
    5. 5
      Schmidt, P. T.; Näslund, E.; Grybäck, P.; Jacobsson, H.; Holst, J. J.; Hilsted, L.; Hellström, P. M. A role for pancreatic polypeptide in the regulation of gastric emptying and short-term metabolic control. J. Clin. Endocrinol. Metab. 2005, 90, 52415246,  DOI: 10.1210/jc.2004-2089
    6. 6
      Painsipp, E.; Wultsch, T.; Edelsbrunner, M. E.; Tasan, R. O.; Singewald, N.; Herzog, H.; Holzer, P. Reduced anxiety-like and depression-related behavior in neuropeptide Y Y4 receptor knockout mice. Genes, Brain Behav. 2008, 7, 532542,  DOI: 10.1111/j.1601-183X.2008.00389.x
    7. 7
      Lin, S.; Shi, Y.-C.; Yulyaningsih, E.; Aljanova, A.; Zhang, L.; Macia, L.; Nguyen, A. D.; Lin, E.-J. D.; During, M. J.; Herzog, H.; Sainsbury, A. Critical role of arcuate Y4 receptors and the melanocortin system in pancreatic polypeptide-induced reduction in food intake in mice. PLoS One 2009, 4, e8488  DOI: 10.1371/journal.pone.0008488
    8. 8
      Tasan, R. O.; Lin, S.; Hetzenauer, A.; Singewald, N.; Herzog, H.; Sperk, G. Increased novelty-induced motor activity and reduced depression-like behavior in neuropeptide Y (NPY)-Y4 receptor knockout mice. Neuroscience 2009, 158, 17171730,  DOI: 10.1016/j.neuroscience.2008.11.048
    9. 9
      Sainsbury, A.; Shi, Y. C.; Zhang, L.; Aljanova, A.; Lin, Z.; Nguyen, A. D.; Herzog, H.; Lin, S. Y4 receptors and pancreatic polypeptide regulate food intake via hypothalamic orexin and brain-derived neurotropic factor dependent pathways. Neuropeptides 2010, 44, 261268,  DOI: 10.1016/j.npep.2010.01.001
    10. 10
      Li, J. B.; Asakawa, A.; Terashi, M.; Cheng, K.; Chaolu, H.; Zoshiki, T.; Ushikai, M.; Sheriff, S.; Balasubramaniam, A.; Inui, A. Regulatory effects of Y4 receptor agonist (BVD-74D) on food intake. Peptides 2010, 31, 17061710,  DOI: 10.1016/j.peptides.2010.06.011
    11. 11
      Moriya, R.; Fujikawa, T.; Ito, J.; Shirakura, T.; Hirose, H.; Suzuki, J.; Fukuroda, T.; MacNeil, D. J.; Kanatani, A. Pancreatic polypeptide enhances colonic muscle contraction and fecal output through neuropeptide Y Y4 receptor in mice. Eur. J. Pharmacol. 2010, 627, 258264,  DOI: 10.1016/j.ejphar.2009.09.057
    12. 12
      Brothers, S. P.; Wahlestedt, C. Therapeutic potential of neuropeptide Y (NPY) receptor ligands. EMBO Mol. Med. 2010, 2, 429439,  DOI: 10.1002/emmm.201000100
    13. 13
      Verma, D.; Hörmer, B.; Bellmann-Sickert, K.; Thieme, V.; Beck-Sickinger, A. G.; Herzog, H.; Sperk, G.; Tasan, R. O. Pancreatic polypeptide and its central Y4 receptors are essential for cued fear extinction and permanent suppression of fear. Br. J. Pharmacol. 2016, 173, 19251938,  DOI: 10.1111/bph.13456
    14. 14
      Zhang, L.; Bijker, M. S.; Herzog, H. The neuropeptide Y system: Pathophysiological and therapeutic implications in obesity and cancer. Pharmacol. Ther. 2011, 131, 91113,  DOI: 10.1016/j.pharmthera.2011.03.011
    15. 15
      Yi, M.; Li, H.; Wu, Z.; Yan, J.; Liu, Q.; Ou, C.; Chen, M. A promising therapeutic target for metabolic diseases: Neuropeptide Y receptors in humans. Cell. Physiol. Biochem. 2018, 45, 88107,  DOI: 10.1159/000486225
    16. 16
      Allen, M.; Reeves, J.; Mellor, G. High throughput fluorescence polarization: a homogeneous alternative to radioligand binding for cell surface receptors. J. Biomol. Screening 2000, 5, 6369,  DOI: 10.1177/108705710000500202
    17. 17
      Leyris, J.-P.; Roux, T.; Trinquet, E.; Verdié, P.; Fehrentz, J.-A.; Oueslati, N.; Douzon, S.; Bourrier, E.; Lamarque, L.; Gagne, D. Homogeneous time-resolved fluorescence-based assay to screen for ligands targeting the growth hormone secretagogue receptor type 1a. Anal. Biochem. 2011, 408, 253262,  DOI: 10.1016/j.ab.2010.09.030
    18. 18
      Emami-Nemini, A.; Roux, T.; Leblay, M.; Bourrier, E.; Lamarque, L.; Trinquet, E.; Lohse, M. J. Time-resolved fluorescence ligand binding for G protein–coupled receptors. Nat. Protoc. 2013, 8, 13071320,  DOI: 10.1038/nprot.2013.073
    19. 19
      Veiksina, S.; Kopanchuk, S.; Rinken, A. Budded baculoviruses as a tool for a homogeneous fluorescence anisotropy-based assay of ligand binding to G protein-coupled receptors: The case of melanocortin 4 receptors. Biochim. Biophys. Acta, Biomembr. 2014, 1838, 372381,  DOI: 10.1016/j.bbamem.2013.09.015
    20. 20
      Schiele, F.; Ayaz, P.; Fernández-Montalván, A. A universal homogeneous assay for high-throughput determination of binding kinetics. Anal. Biochem. 2015, 468, 4249,  DOI: 10.1016/j.ab.2014.09.007
    21. 21
      Antoine, T.; Ott, D.; Ebell, K.; Hansen, K.; Henry, L.; Becker, F.; Hannus, S. Homogeneous time-resolved G protein-coupled receptor-ligand binding assay based on fluorescence cross-correlation spectroscopy. Anal. Biochem. 2016, 502, 2435,  DOI: 10.1016/j.ab.2016.02.017
    22. 22
      Stoddart, L. A.; White, C. W.; Nguyen, K.; Hill, S. J.; Pfleger, K. D. Fluorescence- and bioluminescence-based approaches to study GPCR ligand binding. Br. J. Pharmacol. 2016, 173, 30283037,  DOI: 10.1111/bph.13316
    23. 23
      Stoddart, L. A.; Kilpatrick, L. E.; Hill, S. J. NanoBRET approaches to study ligand binding to GPCRs and RTKs. Trends Pharmacol. Sci. 2018, 39, 136147,  DOI: 10.1016/j.tips.2017.10.006
    24. 24
      Iliopoulos-Tsoutsouvas, C.; Kulkarni, R. N.; Makriyannis, A.; Nikas, S. P. Fluorescent probes for G-protein-coupled receptor drug discovery. Expert Opin. Drug Discovery 2018, 13, 933947,  DOI: 10.1080/17460441.2018.1518975
    25. 25
      Ziemek, R.; Schneider, E.; Kraus, A.; Cabrele, C.; Beck-Sickinger, A. G.; Bernhardt, G.; Buschauer, A. Determination of affinity and activity of ligands at the human neuropeptide Y Y4 receptor by flow cytometry and aequorin luminescence. J. Recept. Signal Transduction 2007, 27, 217233,  DOI: 10.1080/10799890701505206
    26. 26
      Dukorn, S.; Littmann, T.; Keller, M.; Kuhn, K.; Cabrele, C.; Baumeister, P.; Bernhardt, G.; Buschauer, A. Fluorescence- and radiolabeling of [Lys4,Nle17,30]hPP yields molecular tools for the NPY Y4 receptor. Bioconjugate Chem. 2017, 28, 12911304,  DOI: 10.1021/acs.bioconjchem.7b00103
    27. 27
      Tang, T.; Tan, Q.; Han, S.; Diemar, A.; Löbner, K.; Wang, H.; Schüß, C.; Behr, V.; Mörl, K.; Wang, M.; Chu, X.; Yi, C.; Keller, M.; Kofoed, J.; Reedtz-Runge, S.; Kaiser, A.; Beck-Sickinger, A. G.; Zhao, Q.; Wu, B. Receptor-specific recognition of NPY peptides revealed by structures of NPY receptors. Sci. Adv. 2022, 8, eabm1232  DOI: 10.1126/sciadv.abm1232
    28. 28
      Kuhn, K. K.; Littmann, T.; Dukorn, S.; Tanaka, M.; Keller, M.; Ozawa, T.; Bernhardt, G.; Buschauer, A. In search of NPY Y4R antagonists: Incorporation of carbamoylated arginine, aza-amino acids, or D-amino acids into oligopeptides derived from the C-termini of the endogenous agonists. ACS Omega 2017, 2, 36163631,  DOI: 10.1021/acsomega.7b00451
    29. 29
      Spinnler, K.; von Krüchten, L.; Konieczny, A.; Schindler, L.; Bernhardt, G.; Keller, M. An alkyne-functionalized arginine for solid-phase synthesis enabling “bioorthogonal” peptide conjugation. ACS Med. Chem. Lett. 2020, 11, 334339,  DOI: 10.1021/acsmedchemlett.9b00388
    30. 30
      Konieczny, A.; Conrad, M.; Ertl, F. J.; Gleixner, J.; Gattor, A. O.; Grätz, L.; Schmidt, M. F.; Neu, E.; Horn, A. H. C.; Wifling, D.; Gmeiner, P.; Clark, T.; Sticht, H.; Keller, M. N-terminus to arginine side-chain cyclization of linear peptidic neuropeptide Y Y4 receptor ligands results in picomolar binding constants. J. Med. Chem. 2021, 64, 1674616769,  DOI: 10.1021/acs.jmedchem.1c01574
    31. 31
      Gleixner, J.; Gattor, A. O.; Humphrys, L. J.; Brunner, T.; Keller, M. 3H]UR-JG102 - a radiolabeled cyclic peptide with high affinity and excellent selectivity for the neuropeptide Y Y4 receptor. J. Med. Chem. 2023, 66, 1378813808,  DOI: 10.1021/acs.jmedchem.3c01224
    32. 32
      Keller, M.; Kuhn, K. K.; Einsiedel, J.; Hübner, H.; Biselli, S.; Mollereau, C.; Wifling, D.; Svobodová, J.; Bernhardt, G.; Cabrele, C.; Vanderheyden, P. M. L.; Gmeiner, P.; Buschauer, A. Mimicking of arginine by functionalized Nω-carbamoylated arginine as a new broadly applicable approach to labeled bioactive peptides: high affinity angiotensin, neuropeptide Y, neuropeptide FF, and neurotensin receptor ligands as examples. J. Med. Chem. 2016, 59, 19251945,  DOI: 10.1021/acs.jmedchem.5b01495
    33. 33
      Keller, M.; Weiss, S.; Hutzler, C.; Kuhn, K. K.; M?llereau, C.; Dukorn, S.; Schindler, L.; Bernhardt, G.; König, B.; Buschauer, A. Nω-carbamoylation of the argininamide moiety: An avenue to insurmountable NPY Y1 receptor antagonists and a radiolabeled selective high-affinity molecular tool ([3H]UR-MK299) with extended residence time. J. Med. Chem. 2015, 58, 88348849,  DOI: 10.1021/acs.jmedchem.5b00925
    34. 34
      Konieczny, A.; Braun, D.; Wifling, D.; Bernhardt, G.; Keller, M. Oligopeptides as neuropeptide Y Y4 receptor ligands: identification of a high-affinity tetrapeptide agonist and a hexapeptide antagonist. J. Med. Chem. 2020, 63, 81988215,  DOI: 10.1021/acs.jmedchem.0c00426
    35. 35
      Kuhn, K. K.; Ertl, T.; Dukorn, S.; Keller, M.; Bernhardt, G.; Reiser, O.; Buschauer, A. High affinity agonists of the neuropeptide Y (NPY) Y4 receptor derived from the C-terminal pentapeptide of human pancreatic polypeptide (hPP): Synthesis, stereochemical discrimination, and radiolabeling. J. Med. Chem. 2016, 59, 60456058,  DOI: 10.1021/acs.jmedchem.6b00309
    36. 36
      Berlicki, Ł.; Kaske, M.; Gutiérrez-Abad, R.; Bernhardt, G.; Illa, O.; Ortuño, R. M.; Cabrele, C.; Buschauer, A.; Reiser, O. Replacement of Thr32 and Gln34 in the C-terminal neuropeptide Y fragment 25–36 by cis-cyclobutane and cis-cyclopentane β-amino acids shifts selectivity toward the Y4 receptor. J. Med. Chem. 2013, 56, 84228431,  DOI: 10.1021/jm4008505
    37. 37
      Wirth, U.; Erl, J.; Azzam, S.; Höring, C.; Skiba, M.; Singh, R.; Hochmuth, K.; Keller, M.; Wegener, J.; König, B. Monitoring the reversibility of GPCR signaling by combining photochromic ligands with label-free impedance analysis. Angew. Chem., Int. Ed. 2023, 62, e202215547  DOI: 10.1002/anie.202215547
    38. 38
      Archipowa, N.; Wittmann, L.; Köckenberger, J.; Ertl, F. J.; Gleixner, J.; Keller, M.; Heinrich, M. R.; Kutta, R. J. Characterization of fluorescent dyes frequently used for bioimaging: Photophysics and photocatalytical reactions with proteins. J. Phys. Chem. B 2023, 127, 95329542,  DOI: 10.1021/acs.jpcb.3c04484
    39. 39
      Keller, M.; Mahuroof, S. A.; Yee, V. H.; Carpenter, J.; Schindler, L.; Littmann, T.; Pegoli, A.; Hübner, H.; Bernhardt, G.; Gmeiner, P.; Holliday, N. D. Fluorescence labeling of neurotensin(8–13) via arginine residues gives molecular tools with high receptor affinity. ACS Med. Chem. Lett. 2020, 11, 1622,  DOI: 10.1021/acsmedchemlett.9b00462
    40. 40
      Müller, C.; Gleixner, J.; Tahk, M.-J.; Kopanchuk, S.; Laasfeld, T.; Weinhart, M.; Schollmeyer, D.; Betschart, M. U.; Lüdeke, S.; Koch, P.; Rinken, A.; Keller, M. Structure-based design of high-affinity fluorescent probes for the neuropeptide Y Y1 receptor. J. Med. Chem. 2022, 65, 48324853,  DOI: 10.1021/acs.jmedchem.1c02033
    41. 41
      Keller, M.; Erdmann, D.; Pop, N.; Pluym, N.; Teng, S.; Bernhardt, G.; Buschauer, A. Red-fluorescent argininamide-type NPY Y1 receptor antagonists as pharmacological tools. Biorg. Med. Chem. 2011, 19, 28592878,  DOI: 10.1016/j.bmc.2011.03.045
    42. 42
      She, X.; Pegoli, A.; Gruber, C. G.; Wifling, D.; Carpenter, J.; Hübner, H.; Chen, M.; Wan, J.; Bernhardt, G.; Gmeiner, P.; Holliday, N. D.; Keller, M. Red-emitting dibenzodiazepinone derivatives as fluorescent dualsteric probes for the muscarinic acetylcholine M2 receptor. J. Med. Chem. 2020, 63, 41334154,  DOI: 10.1021/acs.jmedchem.9b02172
    43. 43
      Katritch, V.; Fenalti, G.; Abola, E. E.; Roth, B. L.; Cherezov, V.; Stevens, R. C. Allosteric sodium in class A GPCR signaling. Trends Biochem. Sci. 2014, 39, 233244,  DOI: 10.1016/j.tibs.2014.03.002
    44. 44
      White, K. L.; Eddy, M. T.; Gao, Z. G.; Han, G. W.; Lian, T.; Deary, A.; Patel, N.; Jacobson, K. A.; Katritch, V.; Stevens, R. C. Structural connection between activation microswitch and allosteric sodium site in GPCR Signaling. Structure 2018, 26, 259269,  DOI: 10.1016/j.str.2017.12.013
    45. 45
      DeVree, B. T.; Mahoney, J. P.; Velez-Ruiz, G. A.; Rasmussen, S. G.; Kuszak, A. J.; Edwald, E.; Fung, J. J.; Manglik, A.; Masureel, M.; Du, Y.; Matt, R. A.; Pardon, E.; Steyaert, J.; Kobilka, B. K.; Sunahara, R. K. Allosteric coupling from G protein to the agonist-binding pocket in GPCRs. Nature 2016, 535, 182186,  DOI: 10.1038/nature18324
    46. 46
      Wingler, L. M.; Lefkowitz, R. J. Conformational basis of G protein-coupled receptor signaling versatility. Trends Cell Biol. 2020, 30, 736747,  DOI: 10.1016/j.tcb.2020.06.002
    47. 47
      Copeland, R. A. Conformational adaptation in drug–target interactions and residence time. Future Med. Chem. 2011, 3, 14911501,  DOI: 10.4155/fmc.11.112
    48. 48
      Vauquelin, G. Simplified models for heterobivalent ligand binding: when are they applicable and which are the factors that affect their target residence time. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2013, 386, 949962,  DOI: 10.1007/s00210-013-0881-0
    49. 49
      Palmberger, D.; Wilson, I. B.; Berger, I.; Grabherr, R.; Rendic, D. SweetBac: a new approach for the production of mammalianised glycoproteins in insect cells. PLoS One 2012, 7, e34226  DOI: 10.1371/journal.pone.0034226
    50. 50
      Schneider, E. H.; Seifert, R. Sf9 cells: A versatile model system to investigate the pharmacological properties of G protein-coupled receptors. Pharmacol. Ther. 2010, 128, 387418,  DOI: 10.1016/j.pharmthera.2010.07.005
    51. 51
      Grätz, L.; Müller, C.; Pegoli, A.; Schindler, L.; Bernhardt, G.; Littmann, T. Insertion of Nanoluc into the extracellular loops as a complementary method to establish BRET-based binding assays for GPCRs. ACS Pharmacol. Transl. Sci. 2022, 5, 11421155,  DOI: 10.1021/acsptsci.2c00162
    52. 52
      Calderón, J. C.; Plut, E.; Keller, M.; Cabrele, C.; Reiser, O.; Gervasio, F. L.; Clark, T. Extended metadynamics protocol for binding/unbinding free energies of peptide ligands to class A G-protein-coupled receptors. J. Chem. Inf. Model. 2024, 64, 205218,  DOI: 10.1021/acs.jcim.3c01574
    53. 53
      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.; Wood, K. V. Engineered luciferase reporter from a deep sea shrimp utilizing a novel imidazopyrazinone substrate. ACS Chem. Biol. 2012, 7, 18481857,  DOI: 10.1021/cb3002478
    54. 54
      Keller, M.; Kaske, M.; Holzammer, T.; Bernhardt, G.; Buschauer, A. Dimeric argininamide-type neuropeptide Y receptor antagonists: Chiral discrimination between Y1 and Y4 receptors. Biorg. Med. Chem. 2013, 21, 63036322,  DOI: 10.1016/j.bmc.2013.08.065
    55. 55
      Pegoli, A.; She, X.; Wifling, D.; Hubner, H.; Bernhardt, G.; Gmeiner, P.; Keller, M. Radiolabeled dibenzodiazepinone-type antagonists give evidence of dualsteric binding at the M(2) muscarinic acetylcholine receptor. J. Med. Chem. 2017, 60, 33143334,  DOI: 10.1021/acs.jmedchem.6b01892
    56. 56
      Schindler, L.; Moosbauer, J.; Schmidt, D.; Spruss, T.; Grätz, L.; Lüdeke, S.; Hofheinz, F.; Meister, S.; Echtenacher, B.; Bernhardt, G.; Pietzsch, J.; Hellwig, D.; Keller, M. Development of a neurotensin-derived 68Ga-labeled PET ligand with high in vivo stability for imaging of NTS1 receptor-expressing tumors. Cancers 2022, 14, 4922  DOI: 10.3390/cancers14194922
    57. 57
      Schindler, L.; Wohlfahrt, K.; von Krüchten, L. G.; Prante, O.; Keller, M.; Maschauer, S. Neurotensin analogs by fluoroglycosylation at Nω-carbamoylated arginines for PET imaging of NTS1-positive tumors. Sci. Rep. 2022, 12, 15028  DOI: 10.1038/s41598-022-19296-0
    58. 58
      Höfelschweiger, B. K. The pyrylium dyes: A new class of biolabels. Synthesis, spectroscopy, and application as labels and in general protein assay. Doctoral thesis, University of Regensburg, Regensburg2005.
    59. 59
      Moser, C.; Bernhardt, G.; Michel, J.; Schwarz, H.; Buschauer, A. Cloning and functional expression of the hNPY Y5 receptor in human endometrial cancer (HEC-1B) cells. Can. J. Physiol. Pharmacol. 2000, 78, 134142,  DOI: 10.1139/y99-125
    60. 60
      Bartole, E.; Grätz, L.; Littmann, T.; Wifling, D.; Seibel, U.; Buschauer, A.; Bernhardt, G. UR-DEBa242: A Py-5-labeled fluorescent multipurpose probe for investigations on the histamine H3 and H4 receptors. J. Med. Chem. 2020, 63, 52975311,  DOI: 10.1021/acs.jmedchem.0c00160
    61. 61
      Mansouri, M.; Bellon-Echeverria, I.; Rizk, A.; Ehsaei, Z.; Cosentino, C. C.; Silva, C. S.; Xie, Y.; Boyce, F. M.; Davis, M. W.; Neuhauss, S. C. F.; Taylor, V.; Ballmer-Hofer, K.; Berger, I.; Berger, P. Highly efficient baculovirus-mediated multigene delivery in primary cells. Nat. Commun. 2016, 7, 11529  DOI: 10.1038/ncomms11529
    62. 62
      Nie, Y.; Chaillet, M.; Becke, C.; Haffke, M.; Pelosse, M.; Fitzgerald, D.; Collinson, I.; Schaffitzel, C.; Berger, I. ACEMBL tool-kits for high-throughput multigene delivery and expression in prokaryotic and eukaryotic hosts. Adv. Exp. Med. Biol. 2016, 896, 2742,  DOI: 10.1007/978-3-319-27216-0_3
    63. 63
      Cheng, Y.-C.; Prusoff, W. H. Relationship between the inhibition constant (KI) and the concentration of inhibitor which causes 50% inhibition (IC50) of an enzymatic reaction. Biochem. Pharmacol. 1973, 22, 30993108,  DOI: 10.1016/0006-2952(73)90196-2
    64. 64
      Laasfeld, T.; Kopanchuk, S.; Rinken, A. Image-based cell-size estimation for baculovirus quantification. BioTechniques 2017, 63, 161168,  DOI: 10.2144/000114595
    65. 65
      Veiksina, S.; Tahk, M.-J.; Laasfeld, T.; Link, R.; Kopanchuk, S.; Rinken, A. Fluorescence Anisotropy-Based Assay for Characterization of Ligand Binding Dynamics to GPCRs: The Case of Cy3B-Labeled Ligands Binding to MC4 Receptors in Budded Baculoviruses. In G Protein-Coupled Receptor Screening Assays: Methods and Protocols; Martins, S. A. M.; Prazeres, D. M. F., Eds.; Springer: US, New York, NY, 2021; pp 119136.
    66. 66
      Kopanchuk, S.; Vavers, E.; Veiksina, S.; Ligi, K.; Zvejniece, L.; Dambrova, M.; Rinken, A. Intracellular dynamics of the Sigma-1 receptor observed with super-resolution imaging microscopy. PLoS One 2022, 17, e0268563  DOI: 10.1371/journal.pone.0268563
    67. 67
      Soulez, F. In A “Learn 2D, Apply 3D” Method for 3D Deconvolution Microscopy, 2014 IEEE 11th International Symposium on Biomedical Imaging (ISBI); IEEE, 2014; pp 10751078.
    68. 68
      Chenouard, N.; Bloch, I.; Olivo-Marin, J. C. Multiple hypothesis tracking for cluttered biological image sequences. IEEE Trans. Pattern Anal. Mach. Intell. 2013, 35, 27363750,  DOI: 10.1109/TPAMI.2013.97
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsptsci.4c00013.

    • Preparation of the azido-functionalized Py-5 derivative 15; chromatograms of the investigation of the chemical stability of 1618 in PBS pH 7.4 (Figure S1); radioligand displacement curves from competition binding experiments with [3H]UR-KK200 (Figure S2); concentration–response curves of hPP, 11 and 1619 obtained from a hY4R Ca2+ aequorin assay (Figure S3); concentration–response curves of hPP, 5, 11, and 1619 obtained from a hY4R mini-Gsi protein recruitment assay (Figure S4); concentration–response curves of hPP, 16 and 17 obtained from a hY4R CAMYEN cAMP assay (Figure S5); study of dummy fluorescent ligands in functional assays (Figure S6); excitation and emission spectra (Figure S7); viability of CHO-hY4R-Gqi5-mtAEQ cells (Figure S8); analyses of flow cytometric saturation binding data of 16 based on nonviable and viable cell populations (Figure S9); syntax of the equation used to fit FA equilibrium binding data (GraphPad Prism 5); RP-HPLC chromatograms of compounds 11 and 1619; and 1H NMR spectra of compound 11 (PDF)

    • TIRF video sequence (recorded and shown at a 30 Hz resolution, shown at double speed) showing the interactions of 16 with the basal plasma membrane of adherent SK-OV-3-Y4R cells including single-particle tracking (magnified region) (AVI)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.