Next Article in Journal
Assessing Detection Accuracy of Computerized Sonographic Features and Computer-Assisted Reading Performance in Differentiating Thyroid Cancers
Next Article in Special Issue
Local Flexibility of a New Single-Ring Chaperonin Encoded by Bacteriophage AR9 Bacillus subtilis
Previous Article in Journal
The Clinical Utility of Soluble Serum Biomarkers in Autoimmune Pancreatitis: A Systematic Review
Previous Article in Special Issue
Structural and Computational Study of the GroEL–Prion Protein Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modulating the Fibrillization of Parathyroid-Hormone (PTH) Peptides: Azo-Switches as Reversible and Catalytic Entities

1
Department of Chemistry, Faculty of Natural Sciences II, Martin-Luther University Halle-Wittenberg, 06120 Halle (Saale), Germany
2
Department of Physics, Faculty of Natural Sciences II, Martin-Luther University Halle-Wittenberg, 06120 Halle (Saale), Germany
3
Biozentrum, Martin-Luther University Halle-Wittenberg, 06120 Halle (Saale), Germany
4
Core Unit Peptide—Technologies, University Leipzig, 04103 Leipzig, Germany
*
Author to whom correspondence should be addressed.
Biomedicines 2022, 10(7), 1512; https://doi.org/10.3390/biomedicines10071512
Submission received: 23 May 2022 / Revised: 20 June 2022 / Accepted: 23 June 2022 / Published: 26 June 2022
(This article belongs to the Special Issue Protein (Re)Folding and Aggregation)

Abstract

:
We here report a novel strategy to control the bioavailability of the fibrillizing parathyroid hormone (PTH)-derived peptides, where the concentration of the bioactive form is controlled by an reversible, photoswitchable peptide. PTH1–84, a human hormone secreted by the parathyroid glands, is important for the maintenance of extracellular fluid calcium and phosphorus homeostasis. Controlling fibrillization of PTH1–84 represents an important approach for in vivo applications, in view of the pharmaceutical applications for this protein. We embed the azobenzene derivate 3-{[(4-aminomethyl)phenyl]diazenyl}benzoic acid (3,4′-AMPB) into the PTH-derived peptide PTH25–37 to generate the artificial peptide AzoPTH25–37 via solid-phase synthesis. AzoPTH25–37 shows excellent photostability (more than 20 h in the dark) and can be reversibly photoswitched between its cis/trans forms. As investigated by ThT-monitored fibrillization assays, the trans-form of AzoPTH25–37 fibrillizes similar to PTH25–37, while the cis-form of AzoPTH25–37 generates only amorphous aggregates. Additionally, cis-AzoPTH25–37 catalytically inhibits the fibrillization of PTH25–37 in ratios of up to one-fifth. The approach reported here is designed to control the concentration of PTH-peptides, where the bioactive form can be catalytically controlled by an added photoswitchable peptide.

Graphical Abstract

1. Introduction

Fibrillization of proteins and peptides is a supramolecular process [1,2] that leads to the formation of peptide aggregates, containing a cross-β-sheet motif [3]. It involves multiple steps [4] and is associated with many diseases such as Alzheimer’s disease, Parkinson’s disease or diabetes type II [5,6,7]. However, in the past decades, it has also been associated with amyloids with distinct physiological functions, so-called functional amyloids, which are found in lower organisms [8,9,10,11]. Subsequently, functional amyloids were also discovered in humans, whereby the amyloid can be the active physiological form [12,13] or the storage form of peptide hormones [14].
The parathyroid hormone, abbreviated PTH, is a human hormone secreted by the parathyroid glands [15], with PTH-like peptides also known from other animals [16,17]. It is expressed as a 115 residue pre-pro-protein, whereby the first 25 amino acids at the N-terminus (referred to PTH−31–−7) serve as a signaling peptide for the transport to the endoplasmic reticulum and are removed by a signal peptidase [18]. The formed pro-peptide is subsequently transferred to the Golgi apparatus and the N-terminal six amino acids (referred to PTH−6–−1) are proteolytically removed [19]. Before mature PTH1–84 is released into the blood, it is stored in secretory granules as amyloid fibrils [20]. The physiological role is well studied [21,22], being important in the maintenance of extracellular fluid calcium and phosphorus homeostasis. The receptor is mainly activated through the first 34 N-terminal amino acids [23], wherefore recombinant PTH1–84 and recombinant PTH1–34 are approved drugs against osteoporosis, Natpara® and Forteo®, respectively. However, its fibrillization has barely been investigated. Thus far, it is known that the amyloid fibrils of PTH1–84 are formed by the amino acid residues R25-L37, and the thermodynamic stability of the fibrils is sufficiently low to dissociate after dilution [20]. Thus, control over the fibrillization of amyloids and PTH specifically represents an important approach for controlling its factual concentration for in vivo applications, placing modulators of fibrillization and thus reversible fibrillization into the focus of pharmaceutically applicable proteins [24,25,26,27].
In the past decades, the photoinduced switching of protein functionalities has emerged as an important concept to modulate protein function, often by modulations in binding specificity between proteins and ligands. Thus, not only enzymes have been equipped with photosensitive switches, but also larger protein complexes, involved in many physiological or neurological functions [28]. To this end, artificial photoswitches are embedded into either the main chain or side chains of polypeptides, in order to change their secondary structures by photoinduced conformational changes of the photoswitches. Thereby, a plethora of different photoswitches, such as those based on cis-trans-isomerization of azo-dyes [29,30] stilbenes [31] and hemithioindigos [32,33], have been developed. Important for the proper use of a specific photoswitch inside a polypeptide chain is not only the quest to retain the initial (functional) secondary structure of the protein, but also to achieve a reasonably stable conformation after photoswitching, so as to allow for sufficient time to exert the desired effect. Many examples of such sufficiently stable and also reversible photoswitches have been reported, allowing one to modulate several expects of protein function [34,35,36,37,38,39]. Here, we report on an approach to modulate the fibrillization of PTH, equipped with a photoswitch at a specific position in the peptide sequence, in order to reversibly trigger its aggregation/disaggregation (see Figure 1).
In view of the functional design of the modified PTH25–37, we sought to embed the photoswitch into a region of the protein where aggregation is still possible, but only in a specific (untriggered) conformation of the photoswitch, whereby fibrillization should be inhibited after the conformational change. As a model system, we chose peptides derived from the PTH fibril core structure, including the amino acids 25R-37L (Figure 1a) [20], which is able to form fibrils itself. In addition, we investigated the influence of both conformations on the fibrillization of the unmodified peptide. As the photoswitch we chose a structural motif from the class of azobenzenes, as they are well known for enabling reversible control of peptide conformation [29,34,39,40,41]. Specifically we chose the azobenzene derivate 3-{[(4-aminomethyl)phenyl]diazenyl}benzoic acid (3,4′-AMPB; Figure 1b) [42], which is known to introduce a significant geometric change. 3,4′-AMPB displays both: a high photoisomerization yield and a sufficient thermodynamically stability of the cis-isomer [41]. If desired, the photoswitch can be reversed via irradiation at 405 nm, or thermally, with a half-life time of more than 20 h in the dark. We hypothesized that the incorporation of the azobenzene into the backbone would allow us to switch between the cis- and the trans-conformation, whereby one of them is able to fibrillize and the other one is not. Furthermore, azobenzenes in their cis-conformation are known to mimic β-hairpins, which allowed us to investigate the hypothesis if the PTH fibrils possess a turn region like amyloid fibrils from other peptides [43,44,45].

2. Materials and Methods

2.1. General

All technical solvents were distilled prior to use. Air- and moisture-sensitive reactions were carried out in flame-dried glassware under atmospheric pressure of nitrogen. 2-(6-Chloro-1-H-benzotriazole-1-yl)-1,1,3,3-tetramethylaminium hexafluorophosphate (HCTU), N-methyl-morpholine (NMM), N,N-dicyclohexylcarbodiimide (DIC), N-Hydroxybenzotriazole (HOBT), trifluoroacetic acid, 4-aminobenzylamine, and oxone® were purchased from Sigma Aldrich (Taufkirchen, Germany). 9-Fluorenylmethyl-N-succinimidylcarbonat (Fmoc-OSu) was received from Fluorochem. 3-Aminobenzoic acid was purchased from Merck (Darmstadt, Germany). All these chemicals were used without further purification.
NMR spectra were recorded on a Varian Gemini 400 or 500 spectrometer (400 MHz or 500 MHz; Agilent Technologies, Waldbronn, Germany) at 27 °C in DMSO--d6 (99.8 Atom%D; Chemotrade, Düsseldorf, Germany) or D2O (99.8 Atom%D; Sigma-Aldrich, Taufkirchen, Germany). Chemical shifts are given in ppm and referred to the solvent residual signal (DMSO--d6: δ = 2.50 ppm and δ = 39.5 ppm; D2O: δ = 4.79 ppm). The following abbreviations were used for 1H- and 13C-NMR peaks’ assignment: s = singlet, d = doublet, t = triplet, td = triplet of doublet, and m = multiplet. MestReNova (version 6.0.2–5475, Mestrelab Research S.L., Santiago de Compostela, Spain) was used for data interpretation.
ESI-ToF mass spectrometry was performed on a Bruker Daltonics microTOF (Bruker Corporation, Billerica, MA, USA). Samples were dissolved in HPLC-grade solvents (MeOH, THF, or mixtures; Sigma Aldrich, Taufkirchen, Germany) at concentrations of 0.1 mg/mL and measured via direct injection with a flow rate of 180 μL/h using the positive mode with a capillary voltage of 4.5 kV. The spectra were analyzed with otofControl (version 3.4, Bruker Daltonik, Bremen, Germany).

2.2. Organic Synthesis

Fmoc-protected 3,4′-AMPB was synthesized in two steps according to literature procedures [42,46].

2.3. Peptide Synthesis and Purification

Solid-phase peptide synthesis was utilized on an automated peptide synthesizer MultiPep RS (Intavis AG, Koeln, Germany) using standard Fmoc-chemistry and preloaded resins. Standard coupling of all protected natural amino acids was performed as single couplings in dimethylformamid (DMF) using 5 equivalents of amino acids, HCTU as coupling reagents, and 10 equivalents of NMM as base for 1 h at room temperature. Special building groups, such as Fmoc-3,4′-AMPB, were coupled with 3 equivalents using DIC and HOBT in DMF/N-methyl-2-pyrrolidone (NMP) at room temperature and with gentle shaking in the dark overnight.
The N-terminal Fmoc-protecting group was removed by washing the resin with 20% piperidine for 20 min. The final side chain deprotection and cleavage from the resin employed a mixture of trifluoroacetic acid and water (90:10 Vol%) with gentle agitation for 2 h at room temperature.
The crude peptides were purified to >95% purity using preparative RP-HPLC (Gilson, Limburg, Germany). For both analytical and preparative use, the mobile phase was a mixture of water (eluent A) and acetonitrile (eluent B), respectively, each containing 0.1% trifluoroacetic acid. Samples were eluted with a linear gradient from 5% B to 95% B in 15 min for analytical runs and in 90 min for preparative runs on a semipreparative PLRP-S column (300 × 25 mm, 8 μm; Agilent Technologies, Waldbronn, Germany). Finally, all peptides were characterized by analytical HPLC Dionex Ultimate 3000 (Thermo Fisher Scientific, Dreieich, Germany) using a PLRP-S column (150 × 4.6 mm, 3 μm; Agilent Technologies, Waldbronn, Germany) and MALDI-MS (Bruker Microflex LT, Bremen, Germany), which gave the expected [M+H]+ mass peaks.

2.4. Azobenzene Peptide Photoisomerization

Transcis isomerization was performed by irradiating the dissolved peptide in a 1 cm quartz cuvette for 30 min with light of 340 nm wavelength using a 50 W mercury lamp (VEB) and a 340 nm band pass filter (FB340-10, Thorlabs, Bergkirchen, Germany) under stirring. For cistrans isomerization, the dissolved peptide was irradiated with light of 405 nm wavelength using a 1.4 W LED (M405L4, Thorlabs, Bergkirchen, Germany) for 30 min under stirring.

2.5. Aggregation Kinetics

ThT-monitored fibrillization assays of artificial peptides and mixtures with PTH25–37 were investigated by fluorescence intensity measurements using thioflavin T (ThT) as fluorescent dye. Lyophilized peptides were dissolved in 50 mM Na2HPO4 buffer solution with a pH value of 7.4 in a concentration of 2 mg/mL and kept on ice for the next steps. The samples were centrifuged at 13,000× g rpm for 10 s and the concentrations were determined with a JASCO V-660 absorbance spectrometer (JASCO, Pfungstadt, Germany; PTH25–37 by absorbance at 205 nm and the molar extinction coefficient of 49,310 cm−1M−1; trans-AzoPTH25–37 by absorbance at 327 nm and the molar extinction coefficient of 13,000 cm−1M−1). Cis-AzoPTH25–37 was produced as described before. The solutions were centrifuged at 10,000 rpm for 1 h at 4 °C, the supernatant was transferred to another tube. The protein solutions were mixed in the desired ration and diluted with 50 mM Na2HPO4 buffer (pH 7.4) to obtain final concentrations of 0/100 μM PTH25–37, 50 μM ThT, and 0/10/20/50/100 μM AzoPTH25–37. For each sample, a total volume of 480 μL was prepared and 3 × 150 μL were transferred to a medium binding 96-well plate (Greiner Bio-One, Kremsmünster, Austria). The plate was sealed with a microplate cover. The fluorescence intensity was monitored at 37 °C using a BMG FLUOStar Omega multi-mode plate reader (BMG LABTECH, Ortenberg, Germany) using fluorescence excitation and emission wavelengths at 460 nm and 485 nm, respectively. One measurement cycle of 5 min consisted of double-orbital shaking for 150 s and incubating for 150 s.

2.6. Transmission Electron Microscopy (TEM)

TEM images were taken with an electron microscope (EM 900; Zeiss, Oberkochen, Germany) at 80 kV acceleration voltage. For preparation, 5 μL of the peptide solution were added on Formvar/Cu grids (mesh 200). After 3 min of incubation, the grids were gently cleaned with water for o1 min and then negatively stained using uranyl acetate (1%, w/v) for 1 min.

2.7. Seeding Assay

The seeding assay follows the same procedure as the ThT-monitored fibrillization assay for the determination of the aggregation kinetics. In addition, the final samples contained 20 μM of seeds from trans-AzoPTH25–37 fibrils. The seeds were prepared via ultrasonification of a 100 μM mature trans-AzoPTH25–37 fibrils solution (Sonifier W-250 D, Branson Ultraschall, Dietzenbach, Germany; 15 times, 1 s 10% amplitude, 1 s pause).

3. Results & Discussion

3.1. Chemistry

To investigate the fibrillization behavior of PTH25–37, the azobenzene switch was incorporated directly into the peptide backbone. We selected the 3,4′-azobenzene motif (Figure 1b) [42]. As it possesses suitable photochemical properties, e.g., an excellent half-life time with a stability larger than 20 h and switching wavelengths >300 nm. These are easily addressable by our photophysical equipment and also avoid eventual photodegradation. The synthesis was conducted in two steps (Figure 2a): in the first step, we conducted the Fmoc-protection of 2 [46], which in the second step reacts in a Mills reaction with an in situ-generated nitroso compound 3 to obtain the Fmoc-protected 3,4′-AMPB 5 in an overall yield of 68%.
The modified azobenzene switch 5, bearing the proper functionalities for Fmoc-chemistry, was incorporated into the peptide backbone of PTH25–37 via solid-phase peptide synthesis (Figure 2b). It replaces V31 in the artificial peptide AzoPTH25–37, due to its central position along the peptide, expecting the largest impact on fibrillization after photoswitching. Furthermore, we probed the replacement of D30 or the insertion between D30 and V31, which led to a greater loss of solubility in the fibrillization buffer (240 μM vs. 25 μM vs. 60 μM; Table S1). Thus, several of the generated peptides displayed strongly reduced solubility—an effect that is important for the subsequent investigations. All peptides were obtained in yields of 10–19%, and high purities as proven by both HPLC and MALDI-ToF measurements, in addition to 500 MHz NMR spectroscopy (Figures S1–S5 and S13–S15).

3.2. Photophysical Properties

We first studied the photophysical properties of the cis-trans-isomerization of AzoPTH25–37 (Figure 1b) by UV/Vis spectroscopy and HPLC analysis in pure water in order to minimize effects of a potential self-assembly and to quantify the generated amounts of the respective cis/trans-modified peptides before and after photoswitching. The UV/Vis spectra for the pure isomers (Figure S6) were separated from the spectra of trans-enriched AzoPTH25–37 in the thermodynamically stable state after synthesis and in the cis-enriched photostationary state (PSS, Figure 3) with Wolfram Mathematica 12.2. The trans-isomer displays an absorption maximum at 327 nm (ε = 13,000 cm−1M−1) and a second maximum at 427 nm, while the cis-isomer possesses maxima at 288 nm and 433 nm. Both isomers display two isobestic points at 278 nm and 388 nm. They represent in the thermodynamically stable state a cis-trans ratio of 3:97. Under irradiation with UV light (340 nm), the cis-content could be increased of up to 82% in the cis-enriched PSS. Visible light (405 nm) yields 76% of the trans-isomer in the trans-enriched PSS via the back reaction. The difference of the trans-content between the trans-enriched PSS at 405 nm and the thermodynamically stable state arises from the overlapping of the n → π* transitions of both isomers at this wavelength [47]. The rate of thermal cis-to-trans isomerization of AzoPTH25–37 follows first-order kinetics, and was determined by monitoring the increase of the π → π* absorption band at 327 nm (Figure S7) via time-dependent UV measurements. In the absence of light at 37 °C, cis-AzoPTH25–37 isomerizes thermally with a rate constant of 3.53 × 10−6 s−1, corresponding to a half-life time of 79 h.

3.3. Aggregation Kinetics and TEM-Recordings

In order to determine the kinetics of fibril formation of both modified AzoPTH25–37 isomers a thioflavin T (ThT)-monitored fibrillization assay was conducted and compared to PTH25–37. ThT is a benzothiazole compound that binds to the cross-β-sheet structure of amyloid fibrils [48]. Causing a large red shift of fluorescence excitation of ThT, which in turn enables the selective excitation of amyloid fibril-bound ThT and therefore the in situ observation of fibril formation.
In a first attempt, the fibrillization kinetics for pure trans-AzoPTH25–37, cis-AzoPTH25–37, and the PTH-derived peptide PTH25–37 were measured at 37 °C and the results are shown in Figure 4. Two characteristic times were used to characterize the fibrillization (Figure 4, Table 1): the lag time tlag corresponds to the time before an increase in the fluorescence signal occurs; the characteristic time tchar indicates at which time 50% of the maximum fluorescence was reached.
The self-assembly of the trans-AzoPTH25–37 was accelerated compared to PTH25–37, while cis-AzoPTH25–37 exhibited the opposite effect (Figure 4). The first increase of ThT fluorescence was observable after >30 h. Furthermore, cis-AzoPTH25–37 shows a biphasic fibrillization behavior, while trans-AzoPTH25–37 and PTH25–37 show monophasic fibrillization. Compared to PTH25–37, the magnitude of the ThT fluorescence of both AzoPTH25–37 isomers was significant lower (Figure S9). This effect might arise from fluorescence quenching via the azobenzene moiety. To test this hypothesis, the fluorescence lifetime of ThT was measured either alone, in the presence of PTH25–37 fibrils, or in the presence of trans-AzoPTH25–37 fibrils (Figure S8). As expected the lifetime is increased in the presence of PTH25–37 fibrils compared to the control experiment, while it is decreased significantly in the presence of trans-AzoPTH25–37, which further supports our concept. In addition, this effect could be enhanced from a reduced binding affinity of ThT through a different peptide conformation of the fibril.
The observations of the ThT-monitored fibrillization assay were supported by negative stain transmission electron microscopy (TEM) after different time points (Figure 5). After 20 h, amyloid fibrils were only observable for PTH25–37 and trans-AzoPTH25–37 (Figure 5a,b), while cis-AzoPTH25–37 formed amorphous aggregates (Figure 5e). Both peptides produced straight fibrils, whereby the single fibrils of PTH25–37 were larger (>6 μm vs. <1.5 μm) and tend to aggregate further. Interestingly, we found fibrils after 60 h for cis-AzoPTH25–37 (Figure 5g), which matched in the morphology those of trans-AzoPTH25–37 even if they were significantly shorter (<300 nm). This may result from the thermal cis-trans-isomerization, as the cis-content decreases and is reduced to 48% after 60 h.
In further experiments, we investigated the (catalytic) influence of the AzoPTH25–37 isomers on the fibrillization of PTH25–37 (Figure 6). We previously observed such catalytic effects of β-turn modified amyloids (Aβ) on the fibrillization of the Alzheimer peptide Aβ1–40 [49]. Thus 100 μM of PTH25–37 were fibrillized in the presence of various concentrations of the respective AzoPTH25–37 isomer (10/20/50/100 μM). Kinetic measurements revealed that the fibrillization behavior of PTH25–37 was affected in the same way as the pure AzoPTH25–37 isomers.
While trans-AzoPTH25–37 accelerated the fibrillization and therefore reduced tlag and tchar of the mixtures (Figure 6a), cis-AzoPTH25–37 inhibited the fibrillization and extended tlag and tchar (Figure 6b). Interestingly, the biphasic fibrillization behavior of cis-AzoPTH25–37 was also observable for the cis-AzoPTH25–37:PTH25–37 (100 μM:100 μM) mixture. These effects are reduced with decreasing concentration of the respective AzoPTH25–37 isomer. While the mixtures with trans-AzoPTH25–37 exhibited a concentration below 50 μM, trans-AzoPTH25–37 had a higher tlag than pure PTH25–37. However, tchar was still shorter, and the stationary phase of the fibrillization was reached earlier.
TEM images were recorded for the peptide mixtures after 20 h (Figure 7). In contrast to the pure peptides, we could observe fibrils for all investigated ratios. Interestingly, the fibrils formed by the mixtures exhibit a similar twisted morphology regardless of the used AzoPTH25–37 isomer. Furthermore, the formation of larger aggregates like for the pure PTH25–37 (Figure 5) were only observed for a ratio of 1:10, indicating that the AzoPTH25–37 inhibits the formation of larger fibril aggregates.

3.4. Seeding Experiments

To determine whether both isomers of AzoPTH25–37 are able to form fibrils or only the trans-isomer, we investigated, if trans-AzoPTH25–37 fibrils were able to induce seeding [50]. A 100 μM solution of each isomer was treated with 20 μM of mature trans-AzoPTH25–37 fibrils, and the kinetics of the fibril formation were investigated via a ThT-monitored fibrillization assay (Figure 8). While the fibrillization of the trans-isomer was accelerated compared to the unseeded monomer, we were not able to observe fibrillization for the cis-isomer. This indicates that the cis-isomer is unable to nucleate amyloid formation as well as elongate preformed fibrils. The observed fibrils after 60 h for the cis-isomer are presumably formed by the thermally isomerized trans-isomer.

4. Conclusions

We here report for the first time a photoswitchable fibrillizing PTH-derived peptide, which is able to modulate its fibrillization by embedding an azobenzene photoswitch in the middle of PTH25–37. PTH1–84 is a peptide hormone, which is stored as functional amyloids in secretory granules. Its physiological role is well studied, but it still lacks detailed information about its exact fibril structure. We used the 3,4′-AMPB photoswitch to investigate the fibril formation of the fibril core fragment of PTH1–84 by incorporating the azobenzene into the peptide backbone, yielding the modified PTH-derived peptide AzoPTH25–37. We could show that the trans-isomer is able to form fibrils, while the cis-isomer induces a conformational change that inhibits fibril formation. Hypothetically, we can also conclude that there might not be a β-turn in the fibril structure of PTH1–84, as the cis-conformer would be reminiscent of such a structure, whereas the trans-conformer would not. Most importantly, we were able to show that the modified peptides can catalytically inhibit fibrillization of the PTH25–37, underscoring the importance of seeding during this fibrillization process, which in the future allows for a reversible triggering of the fibrillization by light as an external stimulus. Studies are in progress to investigate if the photocontrol is also possible with the photoswitch at other positions of the backbone and if we can also control the fibrillization of full-length PTH1–84 with ours or other modified peptides. This represents a novel strategy to control bioavailability of proteins, specifically of PTH peptides and other fibrillating peptides, where not only the concentration of the bioactive form can be controlled by an added photoswitchable peptide, but also the fibrillization as such, important to guide nerve cell regeneration and other directed growth processes in euraryotic cells. For a potential clinical perspective, we want to investigate the cytotoxicity of our peptides as well as the ability to influence the fibrillization of larger PTH-derived peptides (e.g., PTH1–34 and PTH1–84) in vitro and in vivo. As known from other azobenzene containing drugs/prodrugs (e.g., Prontosil), the azobenzene moiety is metabolized in liver tissue via azoreductases, yielding two aniline moieties or through intestinal microbes [51,52]. This is potentially important for the photoswitching inside cells by light, allowing them to tune the reversible fibrillization of other amyloidogenic peptides, which important for regeneration of nerve cells, as reported earlier. Thus, peptide fibrils can seed potential harmful amyloidogenic peptides, which is known from recent work quite prominently [53]. This is a strategy to trigger fiber-formation from the outside via photochemical triggering—thus avoiding the toxic effects of the fibers outside the cells but enabling triggered fibrillization inside the cell to exert the desired effects, allowing them to promote the recovery of spinal cord injuries.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/biomedicines10071512/s1, Figure S1: (A) HPLC-trace of AzoPTH25–37 (cis-isomer at 5.353, trans-isomer at 5.530). (B) MALDI-spectrum of AzoPTH25–37; Figure S2: (A) HPLC-trace of SP1 (cis-isomer at 5.557, trans-isomer at 5.790). (B) MALDI-spectrum of SP1; Figure S3: (A) HPLC-trace of SP2 (cis-isomer at 5.223, trans-isomer at 5.460). (B) MALDI-spectrum of SP2; Figure S4: (A) HPLC-trace of SP3 (cis-isomer at 5.560, trans-isomer at 5.807). (B) MALDI-spectrum of SP3; Figure S5: (A) HPLC-trace of SP4 (cis-isomer at 5.363, trans-isomer at 5.547). (B) MALDI-spectrum of SP4; Figure S6: Separated UV/Vis-spectra of the pure isomers of AzoPTH25–37; spectra were seperated with Wolfram Mathematica 12.2; Figure S7: (A) UV/Vis-spectra of trans-isomer, cis-enriched PSS, and cis-enriched PSS sample after distinct time points in the dark. (B) logarithmic application of the absorption change over time to determine rate constant k and half-life time t1/2; Figure S8: Time-resolved fluorescence measurement (excitation wavelength = 460 nm, emission wavelength = 480 nm) of unbound ThT (black), ThT bound to PTH25–37 fibrils (dark green), ThT bound to trans-AzoPTH25–37 fibrils (light green); Figure S9: ThT monitored fibrillation assays (c = 100 μM, 37 °C, 50 mM Na2HPO4, pH 7.4). (A) PTH25–37, (B) trans-AzoPTH25–37, (C) cis-AzoPTH25–37; Figure S10: ThT monitored fibrillization assays of mixtures of PTH25–37, trans-AzoPTH25–37, and cis-AzoPTH25–37 (37 °C, 50 mM Na2HPO4, pH 7.4). (A) trans-AzoPTH25–37:PTH25–37 (100 μM:100 μM), (B) cis-AzoPTH25–37:PTH25–37 (100 μM:100 μM), (C) trans-AzoPTH25–37:PTH25–37 (50 μM:100 μM), (D) cis-AzoPTH25–37:PTH25–37 (50 μM:100 μM), (E) trans-AzoPTH25–37:PTH25–37 (20 μM:100 μM), (F) cis-AzoPTH25–37:PTH25–37 (20 μM:100 μM), (G) trans-AzoPTH25–37:PTH25–37 (10 μM:100 μM), (H) cis-AzoPTH25–37:PTH25–37 (10 μM:100 μM); Figure S11: 1H-NMR spectrum (top; 400 MHz, DMSO-d6) and 13C-NMR spectrum (bottom; 100 MHz, DMSO-d6) of (9H-Fluoren-9-yl)methyl (4-aminobenzyl)carbamate; Figure S12: 1H-NMR spectrum (top; 400 MHz, DMSO-d6) and 13C-NMR spectrum (bottom; 100 MHz, DMSO-d6) of Fmoc-3,4′-AMPB (mixture of isomers); Figure S13: 1H-NMR spectra (500 MHz, D2O) of AzoPTH25–37 (top, trans-isomer) and SP1 (bottom, trans-isomer); Figure S14: 1H-NMR spectra (500 MHz, D2O) of SP2 (top, trans-isomer) and SP3 (bottom, trans-isomer); Figure S15: 1H-NMR spectrum (500 MHz, D2O) of SP4 (trans-isomer); Scheme S1: Synthesis of Fmoc-protected 3,4′-AMPB 7. (a) Fmoc-ONSu, triethylamin, DMF/MeCN, 16 h, room temperature. (b) Oxone®, DCM, water, 3 h, room temperature. (c) AcOH, DMSO, N2, 72 h, room temperature; Table S1: Primary sequence and solubility in 50 mM Na2HPO4 buffer (pH 7.4) of peptides AzoPTH25–37 and SP1–SP4.

Author Contributions

Conceptualization, A.P. and W.H.B.; methodology, A.P., B.V., G.H., T.K. and S.R.; validation, A.P. and B.V.; investigation, A.P.; writing—original draft preparation, A.P.; writing—review and editing, A.P. and W.H.B.; supervision, W.H.B.; project administration, W.H.B.; funding acquisition, W.H.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by DFG—Deutsche Forschungsgemeinschaft, Project ID 189853844—TRR 102, TP A12.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors thank Jochen Balbach (Martin Luther University Halle-Wittenberg, Department of Physics) and Maria Ott (Martin Luther University Halle-Wittenberg, Department of Biochemistry and Biotechnology) for the use of equipment, their advice, and for discussions.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Hamley, I.W. Peptide fibrillization. Angew. Chem. Int. Ed. 2007, 46, 8128–8147. [Google Scholar] [CrossRef] [PubMed]
  2. Binder, W.H.; Smrzka, O.W. Self-Assembly of Fibers and Fibrils. Angew. Chem. Int. Ed. 2006, 45, 7324–7328. [Google Scholar] [CrossRef] [PubMed]
  3. Sunde, M.; Serpell, L.C.; Bartlam, M.; Fraser, P.E.; Pepys, M.B.; Blake, C.C. Common core structure of amyloid fibrils by synchrotron X-ray diffraction. J. Mol. Biol. 1997, 273, 729–739. [Google Scholar] [CrossRef] [Green Version]
  4. Karamanos, T.K.; Kalverda, A.P.; Thompson, G.S.; Radford, S.E. Mechanisms of amyloid formation revealed by solution NMR. Prog. Nucl. Magn. Reson. Spectrosc. 2015, 88, 86–104. [Google Scholar] [CrossRef] [Green Version]
  5. Ferreira, S.T.; Vieira, M.N.; De Felice, F.G. Soluble protein oligomers as emerging toxins in Alzheimer’s and other amyloid diseases. IUBMB Life 2007, 59, 332–345. [Google Scholar] [CrossRef] [PubMed]
  6. Irvine, G.B.; El-Agnaf, O.M.; Shankar, G.M.; Walsh, D.M. Protein aggregation in the brain: The molecular basis for Alzheimer’s and Parkinson’s diseases. Mol. Med. 2008, 14, 451–464. [Google Scholar] [CrossRef] [PubMed]
  7. Haataja, L.; Gurlo, T.; Huang, C.J.; Butler, P.C. Islet amyloid in type 2 diabetes, and the toxic oligomer hypothesis. Endocr. Rev. 2008, 29, 303–316. [Google Scholar] [CrossRef] [Green Version]
  8. Chapman, M.R.; Robinson, L.S.; Pinkner, J.S.; Roth, R.; Heuser, J.; Hammar, M.; Normark, S.; Hultgren, S.J. Role of Escherichia coli curli operons in directing amyloid fiber formation. Science 2002, 295, 851–855. [Google Scholar] [CrossRef] [Green Version]
  9. Bayro, M.J.; Daviso, E.; Belenky, M.; Griffin, R.G.; Herzfeld, J. An amyloid organelle, solid-state NMR evidence for cross-β assembly of gas vesicles. J. Biol. Chem. 2012, 287, 3479–3484. [Google Scholar] [CrossRef] [Green Version]
  10. Mackay, J.P.; Matthews, J.M.; Winefield, R.D.; Mackay, L.G.; Haverkamp, R.G.; Templeton, M.D. The hydrophobin EAS is largely unstructured in solution and functions by forming amyloid-like structures. Structure 2001, 9, 83–91. [Google Scholar] [CrossRef] [Green Version]
  11. Kenney, J.M.; Knight, D.; Wise, M.J.; Vollrath, F. Amyloidogenic nature of spider silk. Eur. J. Biochem. 2002, 269, 4159–4163. [Google Scholar] [CrossRef] [PubMed]
  12. Fowler, D.M.; Koulov, A.V.; Alory-Jost, C.; Marks, M.S.; Balch, W.E.; Kelly, J.W. Functional amyloid formation within mammalian tissue. PLoS Biol. 2006, 4, e6. [Google Scholar] [CrossRef]
  13. Li, J.; McQuade, T.; Siemer, A.B.; Napetschnig, J.; Moriwaki, K.; Hsiao, Y.-S.; Damko, E.; Moquin, D.; Walz, T.; McDermott, A. The RIP1/RIP3 necrosome forms a functional amyloid signaling complex required for programmed necrosis. Cell 2012, 150, 339–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Maji, S.K.; Perrin, M.H.; Sawaya, M.R.; Jessberger, S.; Vadodaria, K.; Rissman, R.A.; Singru, P.S.; Nilsson, K.R.; Simon, R.; Schubert, D.; et al. Functional Amyloids As Natural Storage of Peptide Hormones in Pituitary Secretory Granules. Science 2009, 325, 328–332. [Google Scholar] [CrossRef] [Green Version]
  15. Cohn, D.V.; Elting, J. Biosynthesis, processing, and secretion of parathormone and secretory protein-I. Recent Prog. Horm. Res. 1983, 39, 181–209. [Google Scholar]
  16. Usdin, T.B. The PTH2 receptor and TIP39: A new peptide–receptor system. Trends Pharmacol. Sci. 2000, 21, 128–130. [Google Scholar] [CrossRef]
  17. Guerreiro, P.M.; Renfro, J.L.; Power, D.M.; Canario, A.V. The parathyroid hormone family of peptides: Structure, tissue distribution, regulation, and potential functional roles in calcium and phosphate balance in fish. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2007, 292, R679–R696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Freeman, M.W.; Wiren, K.M.; Rapoport, A.; Lazar, M.; Potts, J.T., Jr.; Kronenberg, H.M. Consequences of Amino-Terminal Deletions of Preproparathyroid Hormone Signal Sequence. Mol. Endocrinol. 1987, 1, 628–638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Wiren, K.M.; Potts, J.T., Jr.; Kronenberg, H.M. Importance of the propeptide sequence of human preproparathyroid hormone for signal sequence function. J. Biol. Chem. 1988, 263, 19771–19777. [Google Scholar] [CrossRef]
  20. Gopalswamy, M.; Kumar, A.; Adler, J.; Baumann, M.; Henze, M.; Kumar, S.T.; Fändrich, M.; Scheidt, H.A.; Huster, D.; Balbach, J. Structural characterization of amyloid fibrils from the human parathyroid hormone. Biochim. Biophys. Acta (BBA)-Proteins Proteom. 2015, 1854, 249–257. [Google Scholar] [CrossRef]
  21. Friedlander, G.; Amiel, C. Cellular mode of action of parathyroid hormone. Adv. Nephrol. Necker Hosp. 1994, 23, 265–279. [Google Scholar] [PubMed]
  22. Martin, T.J.; Sims, N.A.; Seeman, E. Physiological and Pharmacological Roles of PTH and PTHrP in Bone using their Shared Receptor, PTH1R. Endocr. Rev. 2021, 42, 383–406. [Google Scholar] [CrossRef] [PubMed]
  23. Mosekilde, L.; Søgaard, C.; Danielsen, C.; Tørring, O.; Nilsson, M. The anabolic effects of human parathyroid hormone (hPTH) on rat vertebral body mass are also reflected in the quality of bone, assessed by biomechanical testing: A comparison study between hPTH-(1–34) and hPTH-(1–84). Endocrinology 1991, 129, 421–428. [Google Scholar] [CrossRef]
  24. Evgrafova, Z.; Voigt, B.; Roos, A.H.; Hause, G.; Hinderberger, D.; Balbach, J.; Binder, W.H. Modulation of amyloid β peptide aggregation by hydrophilic polymers. Phys. Chem. Chem. Phys. 2019, 21, 20999–21006. [Google Scholar] [CrossRef] [PubMed]
  25. Evgrafova, Z.; Voigt, B.; Baumann, M.; Stephani, M.; Binder, W.H.; Balbach, J. Probing Polymer Chain Conformation and Fibril Formation of Peptide Conjugates. ChemPhysChem 2019, 20, 236–240. [Google Scholar] [CrossRef] [PubMed]
  26. Evgrafova, Z.; Rothemund, S.; Voigt, B.; Hause, G.; Balbach, J.; Binder, W.H. Synthesis and Aggregation of Polymer-Amyloid β Conjugates. Macromol. Rapid Commun. 2020, 41, 1900378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Funtan, S.; Evgrafova, Z.; Adler, J.; Huster, D.; Binder, W.H. Amyloid Beta Aggregation in the Presence of Temperature-Sensitive Polymers. Polymers 2016, 8, 178. [Google Scholar] [CrossRef] [Green Version]
  28. Hull, K.; Morstein, J.; Trauner, D. In Vivo Photopharmacology. Chem. Rev. 2018, 118, 10710–10747. [Google Scholar] [CrossRef]
  29. Volgraf, M.; Gorostiza, P.; Numano, R.; Kramer, R.H.; Isacoff, E.Y.; Trauner, D. Allosteric control of an ionotropic glutamate receptor with an optical switch. Nat. Chem. Biol. 2006, 2, 47–52. [Google Scholar] [CrossRef]
  30. Samanta, S.; Qin, C.; Lough, A.J.; Woolley, G.A. Bidirectional photocontrol of peptide conformation with a bridged azobenzene derivative. Angew. Chem. Int. Ed. 2012, 51, 6452–6455. [Google Scholar] [CrossRef]
  31. Lindgren, N.J.V.; Varedian, M.; Gogoll, A. Photochemical Regulation of an Artificial Hydrolase by a Backbone Incorporated Tertiary Structure Switch. Chem. Eur. J. 2009, 15, 501–505. [Google Scholar] [CrossRef] [PubMed]
  32. Schadendorf, T.; Hoppmann, C.; Rück-Braun, K. Synthesis of rigid photoswitchable hemithioindigo ω-amino acids. Tetrahedron Lett. 2007, 48, 9044–9047. [Google Scholar] [CrossRef]
  33. Lougheed, T.; Borisenko, V.; Hennig, T.; Ruck-Braun, K.; Woolley, G.A. Photomodulation of ionic current through hemithioindigo-modified gramicidin channels. Org. Biomol. Chem. 2004, 2, 2798–2801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Albert, L.; Peñalver, A.; Djokovic, N.; Werel, L.; Hoffarth, M.; Ruzic, D.; Xu, J.; Essen, L.O.; Nikolic, K.; Dou, Y. Modulating Protein–Protein Interactions with Visible-Light-Responsive Peptide Backbone Photoswitches. ChemBioChem 2019, 20, 1417–1429. [Google Scholar] [CrossRef] [Green Version]
  35. Pozhidaeva, N.; Cormier, M.-E.; Chaudhari, A.; Woolley, G.A. Reversible photocontrol of peptide helix content: Adjusting thermal stability of the cis state. Bioconjug. Chem. 2004, 15, 1297–1303. [Google Scholar] [CrossRef]
  36. Broichhagen, J.; Podewin, T.; Meyer-Berg, H.; Von Ohlen, Y.; Johnston, N.R.; Jones, B.J.; Bloom, S.R.; Rutter, G.A.; Hoffmann-Röder, A.; Hodson, D.J. Optical control of insulin secretion using an incretin switch. Angew. Chem. Int. Ed. 2015, 54, 15565–15569. [Google Scholar] [CrossRef] [Green Version]
  37. Zhang, Y.; Erdmann, F.; Fischer, G. Augmented photoswitching modulates immune signaling. Nat. Chem. Biol. 2009, 5, 724–726. [Google Scholar] [CrossRef]
  38. Hoppmann, C.; Schmieder, P.; Domaing, P.; Vogelreiter, G.; Eichhorst, J.; Wiesner, B.; Morano, I.; Rück-Braun, K.; Beyermann, M. Photocontrol of contracting muscle fibers. Angew. Chem. Int. Ed. 2011, 50, 7699–7702. [Google Scholar] [CrossRef]
  39. Zhang, F.; Timm, K.A.; Arndt, K.M.; Woolley, G.A. Photocontrol of coiled-coil proteins in living cells. Angew. Chem. Int. Ed. 2010, 49, 3943–3946. [Google Scholar] [CrossRef]
  40. Aemissegger, A.; Kräutler, V.; van Gunsteren, W.F.; Hilvert, D. A photoinducible β-hairpin. J. Am. Chem. Soc. 2005, 127, 2929–2936. [Google Scholar] [CrossRef]
  41. Yeoh, Y.Q.; Yu, J.; Polyak, S.W.; Horsley, J.R.; Abell, A.D. Photopharmacological Control of Cyclic Antimicrobial Peptides. ChemBioChem 2018, 19, 2591–2597. [Google Scholar] [CrossRef]
  42. Rück-Braun, K.; Kempa, S.; Priewisch, B.; Richter, A.; Seedorff, S.; Wallach, L. Azobenzene-Based ω-Amino Acids and Related Building Blocks: Synthesis, Properties, and Application in Peptide Chemistry. Synthesis 2009, 24, 4256–4267. [Google Scholar] [CrossRef] [Green Version]
  43. Ahmed, M.; Davis, J.; Aucoin, D.; Sato, T.; Ahuja, S.; Aimoto, S.; Elliott, J.I.; Van Nostrand, W.E.; Smith, S.O. Structural conversion of neurotoxic amyloid-β 1–42 oligomers to fibrils. Nat. Struct. Mol. Biol. 2010, 17, 561–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Der-Sarkissian, A.; Jao, C.C.; Chen, J.; Langen, R. Structural organization of α-synuclein fibrils studied by site-directed spin labeling. J. Biol. Chem. 2003, 278, 37530–37535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Makin, O.S.; Serpell, L.C. Structural characterisation of islet amyloid polypeptide fibrils. J. Mol. Biol. 2004, 335, 1279–1288. [Google Scholar] [CrossRef]
  46. Murawska, G.M.; Poloni, C.; Simeth, N.A.; Szymanski, W.; Feringa, B.L. Comparative Study of Photoswitchable Zinc-Finger Domain and AT-Hook Motif for Light-Controlled Peptide–DNA Binding. Chem. Eur. J. 2019, 25, 4965–4973. [Google Scholar] [CrossRef] [Green Version]
  47. Bandara, H.M.D.; Burdette, S.C. Photoisomerization in different classes of azobenzene. Chem. Soc. Rev. 2012, 41, 1809–1825. [Google Scholar] [CrossRef]
  48. LeVine, H., III. [18] Quantification of β-sheet amyloid fibril structures with thioflavin T. In Methods in Enzymology; Elsevier: Amsterdam, The Netherlands, 1999; Volume 309, pp. 274–284. [Google Scholar]
  49. Deike, S.; Rothemund, S.; Voigt, B.; Samantray, S.; Strodel, B.; Binder, W.H. β-Turn mimetic synthetic peptides as amyloid-β aggregation inhibitors. Bioorg. Chem. 2020, 101, 104012. [Google Scholar] [CrossRef]
  50. Ren, B.; Zhang, Y.; Zhang, M.; Liu, Y.; Zhang, D.; Gong, X.; Feng, Z.; Tang, J.; Chang, Y.; Zheng, J. Fundamentals of cross-seeding of amyloid proteins: An introduction. J. Mater. Chem. B 2019, 7, 7267–7282. [Google Scholar] [CrossRef]
  51. Mulatihan, D.; Guo, T.; Zhao, Y. Azobenzene photoswitch for isomerization-dependent cancer therapy via azo-combretastatin A4 and phototrexate. Photochem. Photobiol. 2020, 96, 1163–1168. [Google Scholar] [CrossRef]
  52. Williams, R. Hepatic metabolism of drugs. Gut 1972, 13, 579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Álvarez, Z.; Kolberg-Edelbrock, A.N.; Sasselli, I.R.; Ortega, J.A.; Qiu, R.; Syrgiannis, Z.; Mirau, P.A.; Chen, F.; Chin, S.M.; Weigand, S.; et al. Bioactive scaffolds with enhanced supramolecular motion promote recovery from spinal cord injury. Science 2021, 374, 848–856. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Primary sequence of PTH25–37 and the azobenzene-modified PTH25–37 (AzoPTH25–37, azobenzene-moiety highlighted in red). (b) Cis-trans-isomerization of the incorporated 3,4′-AMPB switch. (c) Equilibrium of the monomeric peptides PTH25–37 and AzoPTH25–37 in both forms and their aggregates.
Figure 1. (a) Primary sequence of PTH25–37 and the azobenzene-modified PTH25–37 (AzoPTH25–37, azobenzene-moiety highlighted in red). (b) Cis-trans-isomerization of the incorporated 3,4′-AMPB switch. (c) Equilibrium of the monomeric peptides PTH25–37 and AzoPTH25–37 in both forms and their aggregates.
Biomedicines 10 01512 g001
Figure 2. (a) Synthesis of Fmoc-protected trans-3,4′-AMPB 5. (b) Solid-phase peptide synthesis strategy towards the peptide AzoPTH25–37.
Figure 2. (a) Synthesis of Fmoc-protected trans-3,4′-AMPB 5. (b) Solid-phase peptide synthesis strategy towards the peptide AzoPTH25–37.
Biomedicines 10 01512 g002
Figure 3. UV/Vis absorption spectra for trans-AzoPTH25–37 after synthesis and for the cis-enriched photo-stationary state after irradiation at 340 nm, which almost corresponds to cis-AzoPTH25–37.
Figure 3. UV/Vis absorption spectra for trans-AzoPTH25–37 after synthesis and for the cis-enriched photo-stationary state after irradiation at 340 nm, which almost corresponds to cis-AzoPTH25–37.
Biomedicines 10 01512 g003
Figure 4. ThT-monitored fibrillization assay of PTH25–37, cis-AzoPTH25–37, and trans-AzoPTH25–37 (average of triplets; T = 37 °C, buffer = 50 mM Na2HPO4, pH = 7.4): (black) PTH25–37 (100 μM), (red) trans-AzoPTH25–37 (100 μM), and (blue) cis-AzoPTH25–37 (100 μM).
Figure 4. ThT-monitored fibrillization assay of PTH25–37, cis-AzoPTH25–37, and trans-AzoPTH25–37 (average of triplets; T = 37 °C, buffer = 50 mM Na2HPO4, pH = 7.4): (black) PTH25–37 (100 μM), (red) trans-AzoPTH25–37 (100 μM), and (blue) cis-AzoPTH25–37 (100 μM).
Biomedicines 10 01512 g004
Figure 5. TEM recordings of fibrils obtained from PTH25–37, cis-AzoPTH25–37, and trans-AzoPTH25–37 at (T = 37 °C, buffer = 50 mM Na2HPO4, pH = 7.4) after different time points (all scale bars corresponds to 500 nm). (a). PTH25–37 after 20 h, (b) trans-AzoPTH25–37 after 20 h (100 μM), (c) trans-AzoPTH25–37 after 40 h (100 μM), (d) trans-AzoPTH25–37 after 60 h (100 μM), (e) cis-AzoPTH25–37 after 20 h (100 μM), (f) cis-AzoPTH25–37 after 40 h (100 μM), and (g) cis-AzoPTH25–37 after 60 h (100 μM).
Figure 5. TEM recordings of fibrils obtained from PTH25–37, cis-AzoPTH25–37, and trans-AzoPTH25–37 at (T = 37 °C, buffer = 50 mM Na2HPO4, pH = 7.4) after different time points (all scale bars corresponds to 500 nm). (a). PTH25–37 after 20 h, (b) trans-AzoPTH25–37 after 20 h (100 μM), (c) trans-AzoPTH25–37 after 40 h (100 μM), (d) trans-AzoPTH25–37 after 60 h (100 μM), (e) cis-AzoPTH25–37 after 20 h (100 μM), (f) cis-AzoPTH25–37 after 40 h (100 μM), and (g) cis-AzoPTH25–37 after 60 h (100 μM).
Biomedicines 10 01512 g005
Figure 6. (a) ThT-monitored fibrillization assay of PTH25–37 and mixtures with trans-AzoPTH25–37 (average of triplets; T = 37 °C, buffer = 50 mM Na2HPO4, pH = 7.4): (black) PTH25–37 (100 μM), (red) trans-AzoPTH25–37 (100 μM), (green) trans-AzoPTH25–37:PTH25–37 (100 μM:100 μM), (dark yellow) trans-AzoPTH25–37:PTH25–37 (50 μM:100 μM), (brown) trans-AzoPTH25–37:PTH25–37 (20 μM:100 μM), and (orange) trans-AzoPTH25–37:PTH25–37 (10 μM:100 μM) (b) ThT-monitored fibrillization assay of PTH25–37 and mixtures with cis-AzoPTH25–37 at 37 °C: (black) PTH25–37 (100 μM), (blue) cis-AzoPTH25–37 (100 μM), (purple) cis-AzoPTH25–37:PTH25–37 (100 μM:100 μM), (cyan) cis-AzoPTH25–37:PTH25–37 (50 μM:100 μM), (olive) cis-AzoPTH25–37:PTH25–37 (20 μM:100 μM), and (light blue) cis-AzoPTH25–37:PTH25–37 (10 μM:100 μM).
Figure 6. (a) ThT-monitored fibrillization assay of PTH25–37 and mixtures with trans-AzoPTH25–37 (average of triplets; T = 37 °C, buffer = 50 mM Na2HPO4, pH = 7.4): (black) PTH25–37 (100 μM), (red) trans-AzoPTH25–37 (100 μM), (green) trans-AzoPTH25–37:PTH25–37 (100 μM:100 μM), (dark yellow) trans-AzoPTH25–37:PTH25–37 (50 μM:100 μM), (brown) trans-AzoPTH25–37:PTH25–37 (20 μM:100 μM), and (orange) trans-AzoPTH25–37:PTH25–37 (10 μM:100 μM) (b) ThT-monitored fibrillization assay of PTH25–37 and mixtures with cis-AzoPTH25–37 at 37 °C: (black) PTH25–37 (100 μM), (blue) cis-AzoPTH25–37 (100 μM), (purple) cis-AzoPTH25–37:PTH25–37 (100 μM:100 μM), (cyan) cis-AzoPTH25–37:PTH25–37 (50 μM:100 μM), (olive) cis-AzoPTH25–37:PTH25–37 (20 μM:100 μM), and (light blue) cis-AzoPTH25–37:PTH25–37 (10 μM:100 μM).
Biomedicines 10 01512 g006
Figure 7. TEM recordings of fibrils obtained from PTH25–37, cis-AzoPTH25–37, and trans-AzoPTH25–37 at 37 °C after 20 h (scale bar = 500 nm): (a) trans-AzoPTH25–37 (100 μM), (b) trans-AzoPTH25–37:PTH25–37 (100 μM:100 μM), (c) trans-AzoPTH25–37:PTH25–37 (10 μM:100 μM), (d) cis-AzoPTH25–37 (100 μM), (e) cis-AzoPTH25–37:PTH25–37 (100 μM:100 μM), and (f) cis-AzoPTH25–37:PTH25–37 (10 μM:100 μM).
Figure 7. TEM recordings of fibrils obtained from PTH25–37, cis-AzoPTH25–37, and trans-AzoPTH25–37 at 37 °C after 20 h (scale bar = 500 nm): (a) trans-AzoPTH25–37 (100 μM), (b) trans-AzoPTH25–37:PTH25–37 (100 μM:100 μM), (c) trans-AzoPTH25–37:PTH25–37 (10 μM:100 μM), (d) cis-AzoPTH25–37 (100 μM), (e) cis-AzoPTH25–37:PTH25–37 (100 μM:100 μM), and (f) cis-AzoPTH25–37:PTH25–37 (10 μM:100 μM).
Biomedicines 10 01512 g007
Figure 8. ThT-monitored fibrillization assay of cross-seeding studies with cis-AzoPTH25–37 and trans-AzoPTH25–37 monomeric peptides and mature trans-AzoPTH25–37 fibrils as seeds (average of triplets; cmonomer = 100 μM, cseed = 20 μM, T = 37 °C, buffer 50 mM Na2HPO4, pH = 7.4): (black) trans-AzoPTH25–37 with seeds, (red) trans-AzoPTH25–37 without seeds, and (blue) cis-AzoPTH25–37 with seeds.
Figure 8. ThT-monitored fibrillization assay of cross-seeding studies with cis-AzoPTH25–37 and trans-AzoPTH25–37 monomeric peptides and mature trans-AzoPTH25–37 fibrils as seeds (average of triplets; cmonomer = 100 μM, cseed = 20 μM, T = 37 °C, buffer 50 mM Na2HPO4, pH = 7.4): (black) trans-AzoPTH25–37 with seeds, (red) trans-AzoPTH25–37 without seeds, and (blue) cis-AzoPTH25–37 with seeds.
Biomedicines 10 01512 g008
Table 1. Fibrillization parameters (tlag, tchar) of PTH25–37, cis-AzoPTH25–37, trans-AzoPTH25–37, and mixtures thereof (T = 37 °C, buffer = 50 mM Na2HPO4, pH = 7.4).
Table 1. Fibrillization parameters (tlag, tchar) of PTH25–37, cis-AzoPTH25–37, trans-AzoPTH25–37, and mixtures thereof (T = 37 °C, buffer = 50 mM Na2HPO4, pH = 7.4).
Sampletlag [h]tchar [h]
PTH25–37 (100 μM)7.210.9
cis-AzoPTH25–37 (100 μM)34.442.4
cis-AzoPTH25–37:PTH25–37 (100 μM:100 μM)27.935.7
cis-AzoPTH25–37:PTH25–37 (50 μM:100 μM)16.321.2
cis-AzoPTH25–37:PTH25–37 (20 μM:100 μM)10.114.5
cis-AzoPTH25–37:PTH25–37 (10 μM:100 μM)6.97.9
trans-AzoPTH25–37 (100 μM)1.62.1
trans-AzoPTH25–37:PTH25–37 (100 μM:100 μM)3.04.8
trans-AzoPTH25–37:PTH25–37 (50 μM:100 μM)8.08.6
trans-AzoPTH25–37:PTH25–37 (20 μM:100 μM)8.79.7
trans-AzoPTH25–37:PTH25–37 (10 μM:100 μM)7.58.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Paschold, A.; Voigt, B.; Hause, G.; Kohlmann, T.; Rothemund, S.; Binder, W.H. Modulating the Fibrillization of Parathyroid-Hormone (PTH) Peptides: Azo-Switches as Reversible and Catalytic Entities. Biomedicines 2022, 10, 1512. https://doi.org/10.3390/biomedicines10071512

AMA Style

Paschold A, Voigt B, Hause G, Kohlmann T, Rothemund S, Binder WH. Modulating the Fibrillization of Parathyroid-Hormone (PTH) Peptides: Azo-Switches as Reversible and Catalytic Entities. Biomedicines. 2022; 10(7):1512. https://doi.org/10.3390/biomedicines10071512

Chicago/Turabian Style

Paschold, André, Bruno Voigt, Gerd Hause, Tim Kohlmann, Sven Rothemund, and Wolfgang H. Binder. 2022. "Modulating the Fibrillization of Parathyroid-Hormone (PTH) Peptides: Azo-Switches as Reversible and Catalytic Entities" Biomedicines 10, no. 7: 1512. https://doi.org/10.3390/biomedicines10071512

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop