Next Article in Journal
Multiple Roles of SMC5/6 Complex during Plant Sexual Reproduction
Next Article in Special Issue
Sea Anemones, Actinoporins, and Cholesterol
Previous Article in Journal
Resolution of Inflammation in Retinal Disorders: Briefly the State
Previous Article in Special Issue
Erythrocyte Membrane Nanomechanical Rigidity Is Decreased in Obese Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Alchemical Design of Pharmacological Chaperones with Higher Affinity for Phenylalanine Hydroxylase

by
María Conde-Giménez
1,2,
Juan José Galano-Frutos
1,2,
María Galiana-Cameo
1,2,†,
Alejandro Mahía
1,2,
Bruno L. Victor
3,‡,
Sandra Salillas
1,2,
Adrián Velázquez-Campoy
1,2,4,5,
Rui M. M. Brito
3,
José Antonio Gálvez
6,
María D. Díaz-de-Villegas
6,* and
Javier Sancho
1,2,4,*
1
Departamento de Bioquímica y Biología Molecular y Celular, Facultad de Ciencias, Universidad de Zaragoza, 50009 Zaragoza, Spain
2
Biocomputation and Complex Systems Physics Institute (BIFI)-GBsC-CSIC Joint Unit, Universidad de Zaragoza, 50018 Zaragoza, Spain
3
Coimbra Chemistry Center-Institute of Molecular Sciences (CQC-IMS), Department of Chemistry, University of Coimbra, 3004-535 Coimbra, Portugal
4
Aragon Health Research Institute (IIS Aragón), 50009 Zaragoza, Spain
5
CIBER de Enfermedades Hepáticas y Digestivas CIBERehd, Instituto de Salud Carlos III, 28029 Madrid, Spain
6
Instituto de Síntesis Química y Catálisis Homogénea (ISQCH), Departamento de Química Orgánica, Facultad de Ciencias, Universidad de Zaragoza, 50009 Zaragoza, Spain
*
Authors to whom correspondence should be addressed.
Present Address: Instituto de Síntesis Química y Catálisis Homogénea (ISQCH), Departamento de Química Inorgánica, Facultad de Ciencias, Universidad de Zaragoza, 50009 Zaragoza, Spain.
Present Address: BioISI, Biosystems and Integrative Sciences Institute, Faculty of Sciences, Universidade de Lisboa, 1749-016 Lisboa, Portugal.
Int. J. Mol. Sci. 2022, 23(9), 4502; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23094502
Submission received: 24 March 2022 / Revised: 14 April 2022 / Accepted: 16 April 2022 / Published: 19 April 2022
(This article belongs to the Special Issue State-of-the-Art Biochemistry in Spain)

Abstract

:
Phenylketonuria (PKU) is a rare metabolic disease caused by variations in a human gene, PAH, encoding phenylalanine hydroxylase (PAH), and the enzyme converting the essential amino acid phenylalanine into tyrosine. Many PKU-causing variations compromise the conformational stability of the encoded enzyme, decreasing or abolishing its catalytic activity, and leading to an elevated concentration of phenylalanine in the blood, which is neurotoxic. Several therapeutic approaches have been developed to treat the more severe manifestations of the disorder, but they are either not entirely effective or difficult to adhere to throughout life. In a search for novel pharmacological chaperones to treat PKU, a lead compound was discovered (compound IV) that exhibited promising in vitro and in vivo chaperoning activity on PAH. The structure of the PAH-IV complex has been reported. Here, using alchemical free energy calculations (AFEC) on the structure of the PAH-IV complex, we design a new generation of compound IV-analogues with a higher affinity for the enzyme. Seventeen novel analogues were synthesized, and thermal shift and isothermal titration calorimetry (ITC) assays were performed to experimentally evaluate their stabilizing effect and their affinity for the enzyme. Most of the new derivatives bind to PAH tighter than lead compound IV and induce a greater thermostabilization of the enzyme upon binding. Importantly, the correspondence between the calculated alchemical binding free energies and the experimentally determined ΔΔGb values is excellent, which supports the use of AFEC to design pharmacological chaperones to treat PKU using the X-ray structure of their complexes with the target PAH enzyme.

1. Introduction

Phenylketonuria (PKU) is an inborn error of metabolism caused by more than 1200 different variants of the PAH gene, many of them leading to a reduced enzymatic activity of the encoded phenylalanine hydroxylase (PAH) enzyme (PAHvdb, http://www.biopku.org, last accessed on 7 February 2022). As a consequence, high phenylalanine blood levels build up, which are toxic for the brain [1]. PKU patients are classified according to their blood phenylalanine levels in three groups exhibiting phenotypes of increasing severity, namely, mild hyperphenylalaninemia, mild PKU and classic PKU [2,3]. The neurological symptoms associated with untreated PKU are mental retardation and developmental problems. Fortunately, an early diagnosis from generalized neonatal screening followed by prompt dietary intervention effectively avoids the most severe disease outcomes [4]. The mainstay of PKU treatment is a low phenylalanine diet [5] and, for milder PKU patients, supplementation with BH4 [6], the cofactor of the enzyme, as this compound acts as a pharmacological chaperone (PC) partly recovering the lost enzymatic activity [7]. Despite these treatments, new therapies for different PKU phenotypes are needed [8]. Enzyme replacement therapy based on a PEG-coated phenylalanine metabolizing enzyme of bacterial origin is a novel approach, but substantial side effects have already been reported upon subcutaneous administration [9,10].
In a previous work, a compound (named IV) was identified as promising PC alternative to BH4 in the treatment of PKU [11]. The compound showed a stabilizing effect on tetrameric wild-type PAH and several PKU variants, and gave rise to increased PAH activity in cells transiently transfected, and in mouse liver after oral administration. Compound IV, which will be referred to from here on as IVPC, was described to stabilize the tetrameric functional form of the enzyme in vitro and appeared to act as a canonical PC, binding to the PAH tetrameric folded state, displacing the folding equilibrium towards the native form and rescuing its physiological function [11]. The crystallographic structure of human PAH in complex with IVPC [12] indicates this compound binds to the active site of the enzyme, participating in the catalytic metal coordination sphere (Figure 1), which explains its behavior as a weak competitive PAH inhibitor [11]. On the other hand, PAH refolding kinetics demonstrated an additional chaperoning role for IVPC: it accelerates the folding reaction by stabilizing partly folded species transiently accumulating along the PAH-folding pathway [12]. Interestingly, the reported X-ray structure of the PAH-IVPC complex opens a way to the rational design of improved PCs for PKU treatment.
Aiming at this, we use here the X-ray structure of the complex to design reasonable IVPC-variants compatible with tight binding and to perform alchemical free-energy calculations [13,14,15] to compute the change in the affinity of their complexes, relative to that of IVPC. We have synthesized seventeen such derivatives, spanning a calculated ΔΔGb of 2 kcal mol−1 (unsigned). Their actual ΔΔGb and mid-denaturation temperature change (ΔTm) relative to the reference PAH-IVPC complex were determined using ITC experiments and thermal unfolding, respectively. The feasibility of this approach to obtain PCs with improved binding affinity for the target enzyme was assessed.

2. Results and Discussion

2.1. Design of Novel IVPC Analogues as Potential Pharmacological Chaperones with New Chemical Properties

IVPC, i.e., 5,6-dimethyl-3-(4-methyl-2-pyridinyl)-2-thioxo-2,3-dihydrothieno[2,3-d] pyrimidin-4(1H)-one, is the lead PC of a second generation of designed analogues (compounds IVx). The X-ray structure of the complex between PAH and IVPC (Figure 1) [12] was used to rationally design analogues with novel substituents, aiming at increasing the affinity of the complex.
From a thermodynamic point of view, the affinity of a compound for a protein can be increased either by improving its interactions with the protein or by decreasing its interactions with the solvent. Based on the crystallographic structure of the PAH-IVPC complex, we designed chemical modifications in the three heterocycles (thiophene, pyridine and pyrimidine) of IVPC that introduce new substituents pointing toward either small cavities neighboring bound compound IVPC or to the solvent. The proposed modifications sought to enhance ligand/protein van der Waals interactions and the entropic stabilization of the complex through increased hydrophobic effect. In total, we designed and synthesized 17 IVPC analogues (Table 1) with different physicochemical and lipophilic properties (Table S4). The chemical modifications in IVPC analogues include addition of polar groups to establish new interactions with PAH residues and addition of apolar groups (or substitution of the pyridine ring by other apolar ring systems) to fill parts of the large groove conforming the catalytic site, which remains accessible to the solvent after compound binding. One of these fillable spaces appears just below the pyridine ring (as per the spatial orientation given in Figure S1b) packing onto the α-helix depicted on the right side wall of the catalytic site. Another one appears behind the coordinated center, in the direction pointed by the sulfur atom of the 2-thioxo group (in the pyrimidine central ring), and a third one appears in the direction pointed by the most buried methyl group out of the two in the thiophene ring.
Overall, the modifications lead to analogues (Table S4) more hydrophilic than IVPC (e.g., IVa, IVb, IVd or IVg) and others more hydrophobic (e.g., IVh, IVj, IVm, IVn or IVo). In general, the IVPC analogues maintain the adequate drug likeness of the lead for oral administration (Lipinski’s [16], Veber’s [17] and Egan’s [18] rules), and, according to the BOILED-Egg predictive model (Figure S2), they retain the low probability of crossing the blood–brain barrier exhibited by the lead, which is convenient as they are wished to act in the liver and kidney. A few analogues, though, (IVc, IVd, IVh, IVi, IVj and IVq) may show suboptimal intestinal absorption.

2.2. Synthesis of IVPC Analogues IVa–IVq

Analogues IVaIVg in which substituents on the thiophene or/and the pyridine ring of the lead compound have been varied were prepared from cyclization of the corresponding mixed thioureas using conditions for tert-butoxide-assisted amidation of esters [19] and subsequent side chain transformation when necessary, as detailed below.
Reaction of 2-aminothiophene 1 or 2 with 1,1′-thiocarbonylbis(pyridin-2(1H)-one) [20,21] in dry dichloromethane at room temperature provided thiophene-2-isothiocyanate 3 or 4 with a 69 and 85% yield, respectively. Mixed thioureas 7 and 8 were obtained with excellent yield (98–92%) by reaction of thiophene-2-isothiocyanate 3 with 4-methyl-pyridin-2-amine 5 or 4-(((tert-butyldimethylsilyl)oxy)methyl)pyridin-2-amine 6. The reaction of thiophene-2-isothiocyanate 4 with 4-(((tert-butyldimethylsilyl)oxy)methyl)pyridin-2-amine 6 cleanly afforded mixed thiourea 9 with a 90% yield (Scheme 1).
Mixed thioureas 1315 were synthesized through a one pot procedure from 4-substituted pyridin-2-amines (Scheme 2). Coupling was initialized by reaction of 1,1′-thiocarbonylbis(pyridin-2(1H)-one) with the corresponding pyridin-2-amine. After the time required for in situ generation of the pyridin-2-isothiocyanate, 2-amino-thiophene 2 was added to the resulting reaction mixture. In this way, mixed thioureas 1315 were obtained starting from 4-chloropyridin-2-amine 10, 4-methoxypyridin-2-amine 11 and 4-cyanopyridin-2-amine 12 with a 92, 85 and 88% two-step yield, respectively. This one pot procedure did not lead to the expected results when pyridin-2-amine 5 or 6 with an inductive electron-donating group at C4 was used as starting material.
Cyclization of the mixed thioureas 79, 1315 using potassium tert-butoxide in tert-butanol under reflux conditions provided the corresponding 3-(pyridin-2-yl)-2-thioxo-2,3-dihydrothieno[2,3-d]pyrimidin-4(1H)-ones (Scheme 3). The yield obtained depended on the substitution pattern: IVc (63%), 16 (95%), 17 (51%), IVe (61%), IVf (74%) and IVg (74%).
Subsequent transformation of thiophene and/or pyridine side chain in compounds IVc, 16 and 17 provided analogues IVa, IVd and IVb, respectively.
Analogue IVa was obtained in 31% yield by reduction with LiAlH4 of the ethoxycarbonyl group at C6 in analogue IVc (Scheme 4).
Removal of tert-butyldimethylsilyl group in compound 17 using tetrabutylammonium fluoride in THF cleanly afforded analogue IVb in 79% yield (Scheme 5).
Reduction of the ethoxycarbonyl group at C6 in compound 16 with LiAlH4 followed by removal of tert-butyldimethylsilyl group in the obtained compound 18 using tetrabutylammonium fluoride in THF provided analogue IVd with a 49% overall yield over two steps (Scheme 6).
Analogues IVhIVk modified in the pyrimidine ring were prepared by S-Alkylation of analogue IVc with the corresponding alkyl halide in the presence of potassium hydroxide [22] (Scheme 7). In this way, S-benzylation of IVc provided IVh with 92% yield, reaction of IVc with iodoacetonitrile led to IVi in 91% yield and S-allylation reaction of IVc with 3-chloro-2-methyl-1-propene and allyl bromide provided IVj and IVk with 84 and 91% yield, respectively.
In addition, analogue IVl also modified in the pyrimidine ring was prepared in 95% yield by reduction of the carboxyethyl group at C6 in chaperone IVk with LiAlH4 (Scheme 8).
Sodium hydride promoted reaction of 2-aminothiophene 2 with aryl isothiocyanates led to analogues IVmIVo, in which pyridine ring was replaced by another aromatic ring in a single-step process (Scheme 9). The reaction with 1-isothiocyanato-3,5-dimethylbenzene, 1-isothiocyanatonaphthalene and 5-isothiocyanato-1,2,3,4-tetrahydronaphthalene cleanly afforded compounds IVm, IVn and IVo in 79, 80 and 74% yield, respectively.
Reaction of analogue IVn with iodoacetonitrile in the presence of potassium hydroxide provided analogue IVp with 67% yield (Scheme 10).
The synthesis of analogue IVq started with the preparation of thiophene 19 in 31% yield by cyclization of ethyl-2-oxobutyrate and ethyl cyanoacetate with elemental sulfur according to Gewald procedure [23,24] using diethylamine as base and ethanol as solvent to avoid side chain reactions. The thiophene 19 was converted into isothiocyanate 20 with a 67% yield by reaction with 1,1′-thiocarbonylbis(pyridin-2(1H)-one) as described above. Isothiocyanate 20 reacted with 4-methylpyridin-2-amine 5 to afford mixed thiourea 21 in 95% yield. Finally, analogue IVq was obtained in 85% by treatment of mixed thiourea 21 with potassium tert-butoxide in tert-butanol (Scheme 11).

2.3. In Silico Calculation of the Affinity of the PAH-IVPC and PAH-IVx Complexes

The difference in binding free energy between the PAH-IVPC complex and any of those formed by the enzyme with the IVPC analogues was determined by alchemical free-energy calculations relying on short 5 ns H-REMD simulations, as described in Methods. Two binding scenarios were considered where either a ferrous or a ferric cation (previously parameterized as described in Methods) appears coordinated to the PAH catalytic triad (residues H285, H290 and E330). Using the calculated free energies obtained from the alchemical transformations simulated on the compounds (IVPCIVx) when bound to PAH (∆∆GboundIVPC→IVx) and when solvated alone (∆GsolvIVPC→IVx), the relative binding free energy of each PAH-IVx complex relative to the PAH-IVPC one (∆∆GbIVPC→IVx = ∆GbIVx–∆GbIVPC) was calculated (Table 2). The short simulation time (5 ns) setup used here for the simulations proved to be enough to obtain reproducible results with the number of replicas run (from 3 to 6). In longer simulations, a higher number of replicas led to separation of the compound from the metal center (not shown), which had to be discarded. Clearly, the differential binding energies pertaining to transformations of complexes bearing FeII show much better correspondence with the experimentally determined ones (Table 2) and are subsequently presented. Irrespective of the modeled iron redox state, complexes involving IVPC analogues carrying modifications in the iron coordinating 2-thioxo group of the pyrimidine ring (analogues IVh, IVi, IVj, IVk, IVl and IVp) resulted in poorly reproducible trajectories, with the ligand often being displaced from the enzyme binding site and losing its coordination with the metal center. For some of these complexes we could still determine ∆∆GbIVPCIVx values by increasing the number of simulation replicas and selecting those where the ligand remained bound at the original site at the end of the alchemical transformation. This compound disconnection from the metal center was not observed in the other simulated complexes.
For the complexes bearing FeII, negative ∆∆GbIVPCIVx values (meaning increased affinity) were calculated for analogues IVa, IVc, IVe, IVf, IVm, IVn, IVo and IVq, while analogues IVb, IVd and IVg were calculated to form just slightly less tight complexes than IVPC (Table 2). Thus, unlike the substitutions at the 2-thioxo group in the central pyrimidine ring, which are calculated to be destabilizing, the single substitutions introduced at the thiophene ring and the nonpolar substitutions done at the pyridine ring are calculated to either significantly increase the affinity of the analogue for the protein or, in a few cases, to only mildly decrease it.

2.4. Actual Affinity of the PAH-IVx Complexes and its Effect on PAH Thermostability

The thermostabilizing effect of IVPC analogues on PAH was determined by monitoring PAH unfolding curves using Trp emission fluorescence. In the initial screening work where IVPC was discovered, its PAH stabilizing effect was assessed by fitting the fluorescence curves to a two-state unfolding model [11]. In a recent work describing PAH thermal unfolding in more detail, two spectroscopic thermal unfolding transitions were noticed, and the unfolding curves were fitted to a three-state model [25]. As IVPC primarily stabilizes the second unfolding transition of PAH, the one that takes place at a higher temperature and gives rise to the larger emission intensity, and as IVPC analogues appear to do the same, we evaluated, for simplicity, their thermostabilizing effects by fitting the corresponding unfolding curves to the simpler two-state model (Figure 2). Thus, ΔTm values (TmIVxTm) are reported, which essentially coincide with ΔTm2 values.
Only three analogues carrying substitutions in the thioxo group (IVh, IVj and IVk) show a thermodestabilizing effect, suggesting they could preferentially bind to the PAH unfolded state rather than to the folded one (Figure 2a). Analogues IVb, IVf and IVl exert a moderate stabilizing effect (0 < ΔTm < 5 °C), which is below that of the lead IVpc (Table 2 and Figure 2a). All the other analogues stabilize the enzyme more than IVpc (Figure 2b). Compounds IVi, IVm, IVn and IVp show an impressive effect as they increase the Tm by more than 10 °C.
Thermostabilization of a protein because of ligand binding tends to correlate with the affinity of the complex the ligand forms with the native state of the protein [26]. However, the affinity of the complex is also influenced by other thermodynamic parameters, such as ΔHb, or ΔCpb. Therefore, a perfect correlation between ΔTm and ΔΔGb is not expected. To determine ΔΔGb experimentally, ITC assays were performed with the synthesized analogues and PAH (Table 2 and Figure 3). The affinity of all the complexes is in the micromolar range, and nine of them exhibit a higher affinity (lower Kd) than IVPC (Figure 3). As anticipated, qualitative agreement is observed (Figure S3) between the ΔTm associated to each ligand and the ΔΔGb determined in the IVx complexes relative to the complex formed by IVPC. Clearly, the more thermostabilizing analogues are among those with higher ΔΔGb values (Figure S3).
The effect of introducing substituents in the 2-thioxo group of the central pyrimidine ring on the affinity of the complexes can be assessed by comparing the affinity of the complexes formed by analogues IVh, IVi, IVj and IVk with that of IVc, the affinity of complex formed by IVl with that of IVa, and by comparing the affinity of the complex formed by IVp with that of IVn. The small and polar –CH2–CN substituent (IVi and IVp) leaves the affinity of the complexes close to that of their references (IVc and IVn, respectively, see Table 2). The two other substituents tested, –CH2–C(CH3)CH2 (IVj) and –CH2–CHCH2 (IVk and IVl) destabilize the complex, as shown by comparison with their references (IVc, IVc and IVa, respectively, Table 2), in agreement with the anomalous behavior observed in the MD simulations used to calculate the alchemical ΔΔGb values for those derivatives.
In contrast, the IVa/IVPC, IVc/IVPC and IVd/IVb pairs indicate that replacing the apolar 6-methyl group at the thiophene ring either with –CH2OH or COOEt groups increase the affinity of the complex. The affinity is also increased by replacing the pyridine ring with bulkier chemical groups (see IVm,n,o/IVPC pairs) or by substitution of the methyl group by –OCH3 (see IVf/IVPC pair) or –Cl (see pair IVe/IVPC), but not by -CN (pair IVg/IVPC), which hardly changes the affinity, or –CH2OH (see IVb/IVPC and IVd/IVa pairs) which is destabilizing. The described stabilizing substitutions indicate that both the surface-exposed thiophene ring and the deeply buried pyridine one can be substantially modified resulting in an increase of the affinity of the complex with PAH. Instead, modifications of the central pyrimidine ring may not be equally promising.
The thermodynamic profile of ligand binding gives clues on the interactions and effects contributing to the observed stability of a protein/ligand complex, such as the dominance of direct protein/ligand interactions (e.g., hydrogen bonds or van der Waals) or of nonspecific protein or ligand desolvation (e.g., hydrophobic effect). In general, the entropic optimization of ligands has proved easier to achieve than the enthalpic one due to the difficulty of adding polar groups in the ligand structure at the appropriate distance and orientation to establish strong interactions with the target [27]. The binding thermodynamic profiles of IVPC and all thermostabilizing analogues (Figure 3) were obtained from isothermal calorimetric titrations of PAH (Figure S4). The thermodynamically favorable binding of IVPC to PAH, reflected in its negative ΔGb value, arises from favorable enthalpic and entropic contributions. Although the PAH-IVPC complex involves the formation of two hydrogen bonds between chaperone atoms and protein residues [12], most of the complex stability results from the favorable entropic contribution. Thus, desolvation entropy seems to drive IVPC binding. In the case of the analogues, the same binding scenario is observed: both the entropy and the enthalpy components are stabilizing, but the entropic component dominates, evidencing that complexes exhibiting a higher affinity that IVPC tend to benefit from a larger entropic component than that of the IVPC complex.

2.5. Usefulness of AFEC for the Rational Design of Better PAH Binders

To validate the AFEC methodology implemented here, we compared the alchemically calculated Δ Δ G b values for the IVx analogues (Table 2 and Figure 4) with those experimentally determined by ITC. As discussed above, S-alkylated analogues in the pyrimidine ring gave rise to poorly reproducible MD trajectories and were excluded from the comparison. The fitting in Figure 4, corresponding to the calculations done with the FeII-bearing enzyme, are in fine agreement with the experimental data. This indicates that, having the X-ray structure of a PC bound to the PAH active site, analogues can be designed, the affinity of which can be accurately calculated using AFEC prior to chemical synthesis.
This approach can save much synthetic effort by focusing on the synthesis and testing of the more promising analogues. As the catalytic reaction mechanism of nonheme iron pterin-dependent aromatic amino acid hydroxylases is not totally clear and the redox state of the iron atom during the enzyme catalytic cycle changes [28], we also parameterized the ion as FeIII coordinated to the catalytic triad. The AFEC relative binding energies obtained for this alternative parameterization are compared to the experimental energies in Figure S5. The agreement is clearly worse than that obtained with the FeII parameterization (see also Figure S6). Thus, of the two binding models implemented in this work, the model with the coordinated ferrous cation shows the best correlation with the experimental affinity data obtained by the ITC measurement (Figure 4).
The consistency of this AFEC methodology, as implemented here on the PAH-IVx complexes carrying FeII, was further tested by taking advantage of the energy relationships that are implicit in the thermodynamic cycles shown in Figure 5. As the cycles show, the transformation of IVPC into IVd can be split into that of IVPC into IVa plus that of IVa into IVd (left branch) or, alternatively, into that of IVPC into IVb plus that of IVb into IVd (right branch).
Thus, by using the previously calculated Δ Δ G b   I V P C I V d , Δ Δ G b   I V P C I V a and Δ Δ G b   I V P C I V b values for the corresponding alchemical transformations of IVPC into IVa,b,d (see Table 2), one can arithmetically anticipate that Δ Δ G b   I V a I V d and Δ Δ G b   I V b I V d should have values of +2.25 ± 0.58 kJ/mol and −1.49 ± 1.16 kJ/mol, respectively. To check for consistency of the overall methodology, we calculated those values from the corresponding alchemical transformations using the same parameterization and AFEC protocol. The calculated values of +2.36 ± 0.69 kJ/mol and −1.38 ± 0.21 kJ/mol, respectively, show excellent agreement with the values anticipated from the cycles (Figure 5). The good correlations observed between the computational and experimental relative binding free-energy values (Figure 4) and between the calculated values from the thermodynamic cycles expressions (Figure 5) make it possible to propose this AFEC approach as a valuable tool to ease the design of second generation PCs with improved affinity (better binders) for PAH. Concerning derivatives of IVPC, the approach can be used to anticipate the change in affinity upon introducing modifications in either the thiophene or the pyridine rings, but not in the thioxo group of the pyrimidine ring. The approach is also expected to be useful to design better binders to newly identified PAH ligands for which the X-ray structure in complex with PAH becomes available.
PCs may help recover the lost activity of a protein carrying a deleterious variation (e.g., a single amino acid replacement) by different mechanisms. The classic chaperoning mechanism consists in the PC binding to the native fraction of defective protein molecules, making some of the unfolded ones to fold, as governed by the folding equilibrium constant. To exert this effect efficiently, the higher the affinity of the PC for the native enzyme, the better. This is particularly true of allosteric PCs [29], the activity of which is not expected to be complicated by concomitant inhibition of enzyme activity but to be directly related to binding affinity. In this respect, our approach demonstrates the great potential of AFEC as a reliable design tool for raising the binding affinity of a PC for a protein when the structure of its complex with the target protein is known. By calculating whether an analogue will be either a better or a worse binder to the target protein, e.g., PAH, the collection of analogues that have to be synthesized and then tested in cell or animal models can be significantly narrowed, which is essential in order to reduce costs and time in the drug development process, particularly when it is carried out in an academic setting. However, it should be recalled that the chaperoning effect may be exerted through alternative mechanisms such as modifying folding kinetics [12] or protecting the enzyme against inactivation [30]. It should be also remembered that PCs binding at active sites may behave as enzyme inhibitors. This detrimental effect can be minimized through judicious dosing regimens [31]. On the other hand, effective in vitro chaperoning is a requirement that may or may not translate into effective in vivo chaperoning, depending on the specific pharmacokinetics and pharmacodynamics of each PC candidate. In a previous work, the lead IVPC was shown to display its effect on PAH in both cell and animal models [11]. Functional studies to test the chaperoning effect of this second generation of IVPC analogues are out of the scope of this work, but their chaperoning effect will be tested in the future to assess the impact of increasing the binding affinity of IVPC on the in vivo chaperoning potency and PKU-variant specificity of the new compounds. As parent compound IVPC is a weak competitive inhibitor of PAH [11], the new family of derivatives here described (some with tighter and some with weaker binding) may help fine tune its chaperoning and inhibitory effects on PAH.

3. Materials and Methods

3.1. Reagents and Chemicals

All reagents were of analytical grade and used as obtained from commercial sources. Thiophenes 1 and 2 and pyridin-2-amines 5, 6, 10, 11 and 12 are commercially available and were acquired from AK Scientific, Inc. The other compounds were synthesized. A detailed description of the synthesis and characterization of all the compounds involved in this work is reported in Supplementary Materials. All IVPC analogues were dissolved in 100% dimethylsulfoxide (DMSO) and stored frozen at −20 °C.

3.2. Parameterization of the Metal Center and IVPC Analogues, and Molecular Dynamics (MD) Preparation Setup for the Alchemical (AFEC) Simulations

PAH is a tetrameric metalloenzyme carrying an iron atom per subunit. Prior to performing MD simulations of the complexes between PAH and the different IVPC analogues, the coordinated center including the iron ion and either compound IVPC or one out of their analogues (IVx) was parameterized. Parameterization was done by following an ad hoc methodology relying on a set of programs in the AmberTools18 package [32] (mainly MCPB.py [33]). The coordinated metal center was modeled both with the iron ion with charge 2+ (FeII) and 3+ (FeIII). The protein residues His285, His290 and Glu330 were included in the Fe coordination sphere given the proximity (< 3.0 Å) of their side chain coordinated atom (Nε2 in His, Oε2 in Glu) to the iron ion in the crystal structure of the complex [12] (Figure S1b). The N-ter and C-ter atoms of those residues were capped by acetyl (ACE) or N-methylamide (NME) moieties, respectively, before running Gaussian09 [34] (DFT level: B3LYP/6-31G*) to minimize the systems. Then, force constants of the bonds and angles were updated (Z-Matrix method [35]), and the Merz–Kollman charges [36] extracted and subsequently fitted by the RESP method [37] (see final charges in Table S1). The parameters and charges obtained were then combined with those for the rest of the protein, as taken from the Amber99SB force field [38].
On the other hand, IVPC and IVx analogues were parameterized by means of the DFT function B3LYP/6-31G* (bonds and angles force constants updated), the General Amber Force Field (GAFF) [39] was set for providing the remaining van der Waals (vdW) and coulombic parameters for these compounds, and the Merz–Kollman charges [36] fitted through the RESP method [37]. Since compound IVPC in complex PAH-IVPC appears directly coordinated through the nitrogen N1 of the pyrimidine central ring [12], it seems this nitrogen loses its hydrogen before coordinating the metal center (otherwise it may impede proper coordination). How this occurs, either through a tautomeric mechanism (passing this hydrogen to the 2-thioxo sulfur atom) or of a reduction inflicted by the iron or another reducing agent, is unclear and not be addressed here. In this protocol, we thus setup and parameterized compound IVPC and most of its analogues (the exceptions being IVi, IVj, IVk, IVl and IVp, which are all neutral; see structures in Table 1) with an overall charge of minus one (−1), which means that we removed the referred hydrogen atom from nitrogen N1 on these compounds.
For the MD simulations, the compounds (IVPC and its analogues) and their corresponding complexes (PAH-IVPC or PAH-IVx with either FeII or FeIII) were embedded in a 3 nm and an 8 nm diameter octahedral box, respectively, filled (solvated) with Tip3p [40] water molecules and neutralized with Na+ counterions (when required). The preparation phase was completed with the following sequence of steps: minimization, heating (NVT) and equilibration (NPT). Namely, a 10,000-step steepest-descent minimization was run and then the systems were heated to 300 K through a T-ladder (consisting in running 6 consecutive NVT steps of 50 ns each, with T constant over individual steps and increasing 50 K when passing to the next one, using the Berendsen thermostat [41]). This was followed by a 200 ns NVT step (at the final simulation T of 300 K) that was performed to change to the v-rescale thermostat [42], and by two NPT steps (1 atm), the first one a 250 ns step with the Berendsen barostat [41] and the second one a 250 ns step with the Parrinello–Rahman barostat [43,44,45]. A cutoff of 0.9 nm was set as the maximum radius to account for short-range vdW and coulombic interactions. For the short-range vdW interactions, a potential-shift-Verlet modifier was setup, while for the long-range vdW and electrostatics interactions a PME scheme was implemented. ”All-bonds” restraints were applied (LINCS algorithm [46]). In the AFEC productive phase (5 ns) Hamiltonian replica exchange molecular dynamics (H-REMD) simulations [47,48,49] were performed to enhance conformational sampling and ensure convergence when calculating alchemical free energies [50] for the targeted PAH binders. Sixteen lambdas (see Table S2) were settled and optimized to turn on/off the forces acting on the atoms being transmuted. The GROMACS 4.6.1 package [51] was used to run all the MD simulations.

3.3. Alchemical Free Energy Calculation (AFEC)

Calculation of the relative alchemical binding free energies (∆∆Gb) of the targeted IVx analogues versus IVPC in the PAH complex was implemented, as shown in Figure S1a. AFEC transformations (IVPCIVx) on solvated compounds enabled obtaining the ”solvating” free energy (∆Gsolv), whereas the corresponding AFEC transformations on compounds bound to the metal center in the complex allowed to extract the free-energy change for the ”bound” state (∆Gbound). The later, ∆Gbound, was calculated both for the systems setup with FeII and FeIII (Table S3), and the relative ”binding” free energy ( Δ Δ G b IV PC IVx = Δ G b IV PC Δ G b IVx ) was then calculated Equation (1) subtracting the ”solvating” free energy from the ”bound” free energy:
Δ Δ G b IV PC IVx = Δ G b o u n d IV PC IVx Δ G s o l v IV PC IVx
The multistate Bennet acceptance ratio (MBAR) method [52] was used to calculate the (alchemical) free energy differences (“solvating” and ”bound”).

3.4. Expression and Purification of Human Recombinant PAH

Wild-type human PAH was recombinantly expressed in E. coli BL21 (DE3) cells, purified as described [12,25] and obtained as a tetramer. Essentially, the enzyme was overexpressed as a fusion protein with maltose-binding protein (MBP), purified to homogeneity by affinity chromatography, cleaved from MBP, MBP removed in a second affinity step, and PAH in its tetrameric form finally recovered after a molecular exclusion chromatography step. Final fractions were analyzed by SDS-PAGE, and their concentration determined spectrophotometrically using the theoretical molar extinction coefficient [53].

3.5. Fluorescence Thermal Denaturation Measurements

Thermal denaturation of PAH was monitored by tryptophan fluorescence emission (λexc = 295 nm and λem = 345 nm), from 20 to 90 °C at a heating rate of 1 °C × min−1 on a Cary Eclipse Fluorescence Spectrophotometer (Varian). PAH samples (2 µM monomer concentration) in 20 mM Tris, pH 7.4, with 200 mM NaCl and 100 µM compound (either IVPC or IVx) were prepared for differential scanning fluorimetry analysis. Compounds were initially dissolved in 100% DMSO so that the final DMSO concentration in the samples was 2.5% in all cases. Controls containing PAH (2 µM monomer concentration) and 2.5% DMSO were included.
For simplicity, data analysis was performed by fitting each experimental fluorescence curve to a two-state unfolding model, as described [54], using Equation (2):
F = F N o + m N × T + F U o + m U × T × e Δ G / R × T 1 + e Δ G / R × T
where F corresponds to the fluorescence intensity signal at a given temperature while F N o + m N × T and F U o + m U × T represent the temperature-dependent fluorescence signal of the native and unfolded states of the protein, respectively. F N o and F U o represent the fluorescence signal of the native and unfolded protein at a reference T = 0 K, and m N and m U are the slopes of their linear temperature dependencies. Furthermore, ∆G is the unfolding Gibbs energy change, R is the universal gas constant and T is the temperature.
The fraction of unfolded protein ( χ U ) was determined using Equation (3):
χ U = F F N o + m N × T F U o + m U × T F N o + m N × T
where the temperature of mid-denaturation, Tm, is the temperature at which half of the protein molecules are in the unfolded state ( χ U = 0.5 ) and the other half are native. Thermal shift (ΔTm) is calculated as the difference between the Tm values determined in presence and absence of compound.

3.6. Isothermal Titration Calorimetry (ITC)

ITC measurements were carried out in an Auto-iTC200 (MicroCal, Malvern-Panalytical, Malvern, United Kingdom) using carefully degassed ligand (IVPC or IVx analogues) and PAH solutions. A 300 µM solution of IVPC or analogues dissolved in PBS, pH 7.4, was titrated into PAH (20 µM monomer concentration) in the same buffer. Ligand solutions were prepared from stock compound solutions in 100% DMSO, and all working solutions contained the same residual DMSO concentration (2.5%). Binding titrations were performed at 25 °C by successive injections of 2 µL ligand solution into the reference cell every 150 s, with a stirring speed of 750 rpm. Thermodynamic parameters of the binding equilibrium (the binding constant, Kb, the binding enthalpy change, ∆Hb, and the binding stoichiometry, n) were calculated through nonlinear least squares regression analysis of the data, by using a one-site binding model implemented in the MicroCal LLC ITC module from the Origin 7.0 software package (OriginLab, Northampton, MA, USA). The binding Gibbs energy change, ∆Gb, the dissociation constant, Kd and the entropic component, −T ×S, were obtained from basic thermodynamic relationships with the previously calculated thermodynamic binding parameters [55].

4. Conclusions

AFEC makes it possible to anticipate the change in the affinity of the PAH-IVPC complex upon modification of the chemical structure of the bound pharmacological chaperone. Good correlations between calculated and experimental Δ Δ G b values are obtained by using a model with a FeII parameterized iron ion coordinating the residues of the catalytic triad. Based on this computational approach, a new generation of IVPC analogues was obtained exhibiting improved binding affinity for the target enzyme, which translates into higher PAH thermostabilization. AFEC, and the computational calculation of properties such as those available using absorption, distribution, metabolism, excretion or toxicity predictors, can play an important role in medicinal chemistry by guiding the early selection of the more promising analogues, thereby reducing the time and cost required for their synthesis and testing.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/ijms23094502/s1.

Author Contributions

Conceptualization, J.S.; methodology, M.C.-G., J.J.G.-F., M.G.-C., A.M., B.L.V., S.S., A.V.-C., R.M.M.B., J.A.G., M.D.D.-d.-V. and J.S.; formal analysis, M.C.-G., J.J.G.-F., B.L.V.; A.V.-C., R.M.M.B., M.D.D.-d.-V. and J.S.; resources, R.M.M.B., J.A.G., M.D.D.-d.-V. and J.S.; writing—original draft preparation, M.C.-G., J.J.G.-F. and M.D.D.-d.-V.; writing—review and editing, M.C.-G., J.J.G.-F., M.G.-C., A.M., B.L.V., S.S., A.V.-C., R.M.M.B., J.A.G., M.D.D.-d.-V. and J.S.; funding acquisition, J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by MINECO, Spain, grants BFU2016-78232-P and PID2019-107293GB-I00; EU (Interreg-SUDOE), grant NEUROMED; FECYT-PRECIPITA; and Gobierno de Aragón, Spain, grants LMP30_18 and E45_20R. M.C.-G. was recipient of a predoctoral contract from Gobierno de Aragón, Spain.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

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. Scriver, C.R.; Kaufman, S. Hyperphenylalaninaemia: Phenylalanine hydroxylase deficiency. In The Metabolicand Molecular Basesof Inherited Disease, 8th ed.; Scriver, C.R., Beaudet, A.L., Slyetal, W.S., Eds.; McGraw-Hill: New York, NY, USA, 2001; pp. 1667–1724. [Google Scholar]
  2. Güttler, F. Hyperphenylalaninemia: Diagnosis and classification of the various types of phenylalanine hydroxylase deficiency in childhood. Acta Paediatr Scand. 1980, 280, 180. [Google Scholar]
  3. Van Wegberg, A.M.J.; Macdonald, A.; Ahring, K.; BéLanger-Quintana, A.; Blau, N.; Bosch, A.M.; Burlina, A.; Campistol, J.; Feillet, F.; Giżewska, M.; et al. The complete European guidelines on phenylketonuria: Diagnosis and treatment. Orphanet J. Rare Dis. 2017, 12, 162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Blau, N.; van Spronsen, F.J.; Levy, H.L. Phenylketonuria. Lancet 2010, 376, 1417–1427. [Google Scholar] [CrossRef]
  5. Bickel, H.; Gerrard, J.; Hickmans, E. Preliminary Communication. Lancet 1953, 262, 812–813. [Google Scholar] [CrossRef] [Green Version]
  6. Harding, C.O. New era in treatment for phenylketonuria: Pharmacologic therapy with sapropterin dihydrochloride. Biol. Targets Ther. 2010, 4, 231–236. [Google Scholar] [CrossRef] [Green Version]
  7. Pey, A.L.; Pérez, B.; Desviat, L.R.; Martínez, M.A.; Aguado, C.; Erlandsen, H.; Gámez, A.; Stevens, R.; Thórólfsson, M.; Ugarte, M.; et al. Mechanisms underlying responsiveness to tetrahydrobiopterin in mild phenylketonuria mutations. Hum. Mutat. 2004, 24, 388–399. [Google Scholar] [CrossRef]
  8. Ho, G.; Christodoulou, J. Phenylketonuria: Translating research into novel therapies. Transl. Pediatr. 2014, 3, 49–62. [Google Scholar]
  9. Thomas, J.; Levy, H.; Amato, S.; Vockley, J.; Zori, R.; Dimmock, D.; Harding, C.O.; Bilder, D.A.; Weng, H.H.; Olbertz, J.; et al. Pegvaliase for the treatment of phenylketonuria: Results of a long-term phase 3 clinical trial program (PRISM). Mol. Genet. Metab. 2018, 124, 27–38. [Google Scholar] [CrossRef]
  10. Longo, N.; Dimmock, D.; Levy, H.; Viau, K.; Bausell, H.; Bilder, D.A.; Burton, B.; Gross, C.; Northrup, H.; Rohr, F.; et al. Evidence- and consensus-based recommendations for the use of pegvaliase in adults with phenylketonuria. Genet. Med. 2019, 21, 1851–1867. [Google Scholar] [CrossRef] [Green Version]
  11. Pey, A.L.; Ying, M.; Cremades, N.; Velázquez-Campoy, A.; Scherer, T.; Thöny, B.; Sancho, J.; Martínez, A. Identification of pharmacological chaperones as potential therapeutic agents to treat phenylketonuria. J. Clin. Investig. 2008, 118, 2858–2867. [Google Scholar] [CrossRef] [Green Version]
  12. Torreblanca, R.; Lira-Navarrete, E.; Sancho, J.; Hurtado-Guerrero, R. Structural and Mechanistic Basis of the Interaction between a Pharmacological Chaperone and Human Phenylalanine Hydroxylase. ChemBioChem 2012, 13, 1266–1269. [Google Scholar] [CrossRef] [PubMed]
  13. De Ruiter, A.; Oostenbrink, C. Free energy calculations of protein–ligand interactions. Curr. Opin. Chem. Biol. 2011, 15, 547–552. [Google Scholar] [CrossRef] [PubMed]
  14. Chodera, J.D.; Mobley, D.L.; Shirts, M.R.; Dixon, R.W.; Branson, K.; Pande, V.S. Alchemical free energy methods for drug discovery: Progress and challenges. Curr. Opin. Struct. Biol. 2011, 21, 150–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Schindler, C.E.M.; Baumann, H.; Blum, A.; Böse, D.; Buchstaller, H.-P.; Burgdorf, L.; Cappel, D.; Chekler, E.; Czodrowski, P.; Dorsch, D.; et al. Large-Scale Assessment of Binding Free Energy Calculations in Active Drug Discovery Projects. J. Chem. Inf. Model. 2020, 60, 5457–5474. [Google Scholar] [CrossRef]
  16. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 2001, 46, 3–26. [Google Scholar] [CrossRef]
  17. Veber, D.F.; Johnson, S.R.; Cheng, H.-Y.; Smith, B.R.; Ward, K.W.; Kopple, K.D. Molecular Properties That Influence the Oral Bioavailability of Drug Candidates. J. Med. Chem. 2002, 45, 2615–2623. [Google Scholar] [CrossRef]
  18. Egan, W.J.; Merz, K.M.; Baldwin, J.J. Prediction of Drug Absorption Using Multivariate Statistics. J. Med. Chem. 2000, 43, 3867–3877. [Google Scholar] [CrossRef]
  19. Kim, B.R.; Lee, H.-G.; Kang, S.-B.; Sung, G.H.; Kim, J.-J.; Park, J.K.; Lee, S.-G.; Yoon, Y.-J. tert-Butoxide-Assisted Amidationof Estersunder Green Conditions. Synthesis 2012, 44, 42–50. [Google Scholar]
  20. Kim, S.; Yi, K.Y. 1,1′-Thiocarbonyldi-2,2′-pyridone. A new useful reagent for functional group conversions under essentially neutral conditions. J. Org. Chem. 1986, 51, 2613–2615. [Google Scholar] [CrossRef]
  21. Arya, D.P.; Bruice, T.C. Positively charged deoxynucleic methylthioureas: Synthesis and binding properties of pentameric thymidyl methylthiourea. J. Am. Chem. Soc. 1998, 120, 12419–12427. [Google Scholar] [CrossRef]
  22. Wang, K.; Li, J.; Degterev, A.; Hsu, E.; Yuan, J.; Yuan, C. Structure–activity relationship analysis of a novel necroptosis inhibitor, Necrostatin-5. Bioorganic Med. Chem. Lett. 2007, 17, 1455–1465. [Google Scholar] [CrossRef] [PubMed]
  23. Gewald, K.; Schinke, E.; Böttcher, H. Heterocyclen aus CH-aciden Nitrilen, VIII. 2-Amino-thiophene aus methylenaktiven Nitrilen, Carbonylverbindungen und Schwefel. Eur. J. Inorg. Chem. 1966, 99, 94–100. [Google Scholar] [CrossRef]
  24. Briel, D.; Rybak, A.; Kronbach, C.; Unverferth, K. Substituted 2-Aminothiopen-derivatives: A potential new class of GluR6-Antagonists. Eur. J. Med. Chem. 2010, 45, 69–77. [Google Scholar] [CrossRef] [PubMed]
  25. Conde-Giménez, M.; Sancho, J. Unravelling the Complex Denaturant and Thermal-Induced Unfolding Equilibria of Human Phenylalanine Hydroxylase. Int. J. Mol. Sci. 2021, 22, 6539. [Google Scholar] [CrossRef]
  26. Celej, M.S.; Montich, G.G.; Fidelio, G.D. Protein stability induced by ligand binding correlates with changes in protein flexibility. Protein Sci. 2003, 12, 1496–1506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Claveria-Gimeno, R.; Vega, S.; Abian, O.; Velazquez-Campoy, A. A Look at Ligand Binding Thermodynamics in Drug Discovery. Expert Opin. Drug Discov. 2017, 12, 363–377. [Google Scholar] [CrossRef] [Green Version]
  28. Abu-Omar, M.M.; Loaiza, A.; Hontzeas, N. Reaction Mechanisms of Mononuclear Non-Heme Iron Oxygenases. Chem. Rev. 2005, 105, 2227–2252. [Google Scholar] [CrossRef]
  29. Tran, M.L.; Génisson, Y.; Ballereau, S.; Dehoux, C. Second-Generation Pharmacological Chaperones: Beyond Inhibitors. Molecules 2020, 25, 3145. [Google Scholar] [CrossRef] [PubMed]
  30. Hole, M.; Jorge-Finnigan, A.; Underhaug, J.; Teigen, K.; Martinez, A. Pharmacological Chaperones that Protect Tetrahydrobiopterin Dependent Aromatic Amino Acid Hydroxylases Through Different Mechanisms. Curr. Drug Targets 2016, 17, 1515–1526. [Google Scholar] [CrossRef]
  31. Parenti, G.; Andria, G.; Valenzano, K.J. Pharmacological Chaperone Therapy: Preclinical Development, ClinicalTranslation, and Prospects for the Treatment of Lysosomal Storage Disorders. Mol. Ther. 2015, 23, 1138–1148. [Google Scholar] [CrossRef] [Green Version]
  32. Case, D.A.; Ben-Shalom, I.Y.; Brozell, S.R.; Cerutti, D.S. AMBER2018; University of California: Los Angeles, CA, USA, 2018. [Google Scholar]
  33. Li, P.; Merz, J.K.M. MCPB.py: A Python Based Metal Center Parameter Builder. J. Chem. Inf. Model. 2016, 56, 599–604. [Google Scholar] [CrossRef] [PubMed]
  34. Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.; et al. Gaussian09 (RevisionA02); Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  35. Parsons, J.; Holmes, J.B.; Rojas, J.M.; Tsai, J.; Strauss, C.E.M. Practical conversion from torsion space to Cartesian space forin silico protein synthesis. J. Comput. Chem. 2005, 26, 1063–1068. [Google Scholar] [CrossRef] [PubMed]
  36. Besler, B.H.; Merz, K.M., Jr.; Kollman, P.A. Atomic charges derived from semiempirical methods. J. Comput. Chem. 1990, 11, 431–439. [Google Scholar] [CrossRef]
  37. Woods, R.; Chappelle, R. Restrained electrostatic potential atomic partial charges for condensed-phase simulations of carbohydrates. J. Mol. Struct. THEOCHEM 2000, 527, 149–156. [Google Scholar] [CrossRef] [Green Version]
  38. Hornak, V.; Abel, R.; Okur, A.; Strockbine, B.; Roitberg, A.; Simmerling, C. Comparison of multiple Amber force fields and development of improved protein backbone parameters. Proteins: Struct. Funct. Bioinform. 2006, 65, 712–725. [Google Scholar] [CrossRef] [Green Version]
  39. Wang, J.; Wolf, R.M.; Caldwell, J.W.; Kollman, P.A.; Case, D.A. Development and testing of a general amber force field. J. Comput. Chem. 2004, 25, 1157–1174. [Google Scholar] [CrossRef]
  40. Jorgensen, W.L.; Chandrasekhar, J.; Madura, J.D.; Impey, R.W.; Klein, M.L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 1983, 79, 926–935. [Google Scholar] [CrossRef]
  41. Berendsen, H.J.C.; Postma, J.P.M.; Van Gunsteren, W.F.; DiNola, A.; Haak, J.R. Molecular dynamics with coupling to an external bath. J. Chem. Phys. 1984, 81, 3684–3690. [Google Scholar] [CrossRef] [Green Version]
  42. Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101. [Google Scholar] [CrossRef] [Green Version]
  43. Parrinello, M.; Rahman, A. Strain fluctuations and elastic constants. J. Chem. Phys. 1982, 76, 2662–2666. [Google Scholar] [CrossRef]
  44. Parrinello, M.; Rahman, A. Polymorphic transitions in single crystals: A new molecular dynamics method. J. Appl. Phys. 1981, 52, 7182–7190. [Google Scholar] [CrossRef]
  45. Parrinello, M.; Rahman, A. Crystal Structure and Pair Potentials: A Molecular-Dynamics Study. Phys. Rev. Lett. 1980, 45, 1196–1199. [Google Scholar] [CrossRef]
  46. Hess, B.; Bekker, H.; Berendsen, H.J.C.; Fraaije, J.G.E.M. LINCS: A linear constraint solver for molecular simulations. J. Comput. Chem. 1997, 18, 1463–1472. [Google Scholar] [CrossRef]
  47. Fukunishi, H.; Watanabe, O.; Takada, S. On the Hamiltonian replica exchange method for efficient sampling of biomolecular systems: Application to protein structure prediction. J. Chem. Phys. 2002, 116, 9058–9067. [Google Scholar] [CrossRef]
  48. Liu, P.; Kim, B.; Friesner, R.A.; Berne, B.J. Replica exchange with solute tempering: A method for sampling biological systems in explicit water. Proc. Natl. Acad. Sci. USA 2005, 102, 13749–13754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Liu, P.; Voth, G.A. Smart resolution replica exchange: An efficient algorithm for exploring complex energy landscapes. J. Chem. Phys. 2007, 126, 045106. [Google Scholar] [CrossRef] [PubMed]
  50. Meng, Y.; Dashti, D.S.; Roitberg, A.E. Computing alchemical free energy differences with Hamiltonian replica exchange molecular dynamics (H-REMD) simulations. J. Chem. Theory Comput. 2011, 79, 2721–2727. [Google Scholar] [CrossRef] [Green Version]
  51. Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A.E.; Berendsen, H.J.C. GROMACS: Fast, flexible, and free. J. Comput. Chem. 2005, 26, 1701–1718. [Google Scholar] [CrossRef]
  52. Shirts, M.R.; Chodera, J.D. Statistically optimal analysis of samples from multiple equilibrium states. J. Chem. Phys. 2008, 129, 124105. [Google Scholar] [CrossRef] [Green Version]
  53. Gasteiger, E.; Hoogland, C.; Gattiker, A.; Duvaud, S.; Wilkins, M.R.; Appel, R.D.; Bairoch, A. Protein identification and analysis tools on the ExPASy server. In The Proteomics Protocols Handbook; Walker, J.M., Ed.; Humana Press: Totowa, NJ, USA, 2005; Volume XVIII. [Google Scholar]
  54. Sancho, J. The stability of 2-state, 3-state and more-state proteins from simple spectroscopic techniques… plus the structure of the equilibrium intermediates at the same time. Arch. Biochem. Biophys. 2013, 531, 4–13. [Google Scholar] [CrossRef]
  55. Velazquez-Campoy, A.; Ohtaka, H.; Nezami, A.; Muzammil, S.; Freire, E. Isothermal Titration Calorimetry. Curr. Protoc. Cell Biol. 2004, 23, 17.8.1–17.8.24. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Close view of the PAH catalytic site showing bound IVPC depicted in cyan (PDB ID 4 ANP [12]). The catalytic iron atom (gray sphere) is five-coordinated. Its interaction with N1 of IVPC (in CPK with carbon atoms in cyan) does not perturb its coordination with the residues of the catalytic triad: H285, H290 and E330. Additional compound interactions with water molecules (small red spheres) and PAH residues (in CPK) are also represented.
Figure 1. Close view of the PAH catalytic site showing bound IVPC depicted in cyan (PDB ID 4 ANP [12]). The catalytic iron atom (gray sphere) is five-coordinated. Its interaction with N1 of IVPC (in CPK with carbon atoms in cyan) does not perturb its coordination with the residues of the catalytic triad: H285, H290 and E330. Additional compound interactions with water molecules (small red spheres) and PAH residues (in CPK) are also represented.
Ijms 23 04502 g001
Scheme 1. Synthesis of mixed thioureas 79.
Scheme 1. Synthesis of mixed thioureas 79.
Ijms 23 04502 sch001
Scheme 2. Synthesis of mixed thioureas 1315.
Scheme 2. Synthesis of mixed thioureas 1315.
Ijms 23 04502 sch002
Scheme 3. Cyclization of mixed thioureas to thieno[2,3-d]pyrimidin-4-ones.
Scheme 3. Cyclization of mixed thioureas to thieno[2,3-d]pyrimidin-4-ones.
Ijms 23 04502 sch003
Scheme 4. Conversion of analogue IVc into analogue IVa.
Scheme 4. Conversion of analogue IVc into analogue IVa.
Ijms 23 04502 sch004
Scheme 5. Conversion of compound 17 into chaperone IVb.
Scheme 5. Conversion of compound 17 into chaperone IVb.
Ijms 23 04502 sch005
Scheme 6. Conversion of compound 16 into chaperone IVd.
Scheme 6. Conversion of compound 16 into chaperone IVd.
Ijms 23 04502 sch006
Scheme 7. S-Alkylation of chaperone IVc.
Scheme 7. S-Alkylation of chaperone IVc.
Ijms 23 04502 sch007
Scheme 8. Conversion of chaperone IVk into chaperone IVl.
Scheme 8. Conversion of chaperone IVk into chaperone IVl.
Ijms 23 04502 sch008
Scheme 9. Synthesis of chaperones IVmIVo.
Scheme 9. Synthesis of chaperones IVmIVo.
Ijms 23 04502 sch009
Scheme 10. Conversion of analogue IVn into analogue IVp.
Scheme 10. Conversion of analogue IVn into analogue IVp.
Ijms 23 04502 sch010
Scheme 11. Synthesis of analogue IVq.
Scheme 11. Synthesis of analogue IVq.
Ijms 23 04502 sch011
Figure 2. In vitro effects of IVPC and IVx analogues on PAH thermostability. Thermal unfolding curves of PAH in absence or presence of IVPC and (a) IVx analogues with thermodestabilizing or moderate effect or (b) with thermostabilizing effect. The samples contain PAH (2 µM monomer concentration) and 100 µM compound IVPC (green), IVa (yellow), IVb (gray), IVc (pink), IVd (blue), IVe (brown), IVf (red), IVg (violet), IVh (light magenta), IVi (olive), IVj (light cyan), IVk (light yellow), IVl (navy), IVm (ochre), IVn (orange), IVo (dark gray), IVp (turquoise) and IVq (magenta). Control in absence of compound (black) contains the same % of DMSO as the rest of samples. Continuous lines correspond to two-state fittings of the unfolding curves.
Figure 2. In vitro effects of IVPC and IVx analogues on PAH thermostability. Thermal unfolding curves of PAH in absence or presence of IVPC and (a) IVx analogues with thermodestabilizing or moderate effect or (b) with thermostabilizing effect. The samples contain PAH (2 µM monomer concentration) and 100 µM compound IVPC (green), IVa (yellow), IVb (gray), IVc (pink), IVd (blue), IVe (brown), IVf (red), IVg (violet), IVh (light magenta), IVi (olive), IVj (light cyan), IVk (light yellow), IVl (navy), IVm (ochre), IVn (orange), IVo (dark gray), IVp (turquoise) and IVq (magenta). Control in absence of compound (black) contains the same % of DMSO as the rest of samples. Continuous lines correspond to two-state fittings of the unfolding curves.
Ijms 23 04502 g002
Figure 3. Thermodynamic binding profile of IVPC and IVx analogues for which a significant (ΔTm > 5 °C) thermostabilizing effect was observed: binding Gibbs energy (blue), binding enthalpy (green) and entropic contribution (−T × ΔS) to binding Gibbs energy (red) obtained from ITC experiments.
Figure 3. Thermodynamic binding profile of IVPC and IVx analogues for which a significant (ΔTm > 5 °C) thermostabilizing effect was observed: binding Gibbs energy (blue), binding enthalpy (green) and entropic contribution (−T × ΔS) to binding Gibbs energy (red) obtained from ITC experiments.
Ijms 23 04502 g003
Figure 4. Correlation between computational (by AFEC) and experimental (by ITC) relative binding free energies for the binding of IVPC and the IVx analogues to the enzyme PAH. The AFEC data were obtained by simulating the PAH enzyme as coordinating an FeII iron ion. The straight line corresponds to a linear fit. The fitting equation and the square Pearson coefficient of the fit are also depicted.
Figure 4. Correlation between computational (by AFEC) and experimental (by ITC) relative binding free energies for the binding of IVPC and the IVx analogues to the enzyme PAH. The AFEC data were obtained by simulating the PAH enzyme as coordinating an FeII iron ion. The straight line corresponds to a linear fit. The fitting equation and the square Pearson coefficient of the fit are also depicted.
Ijms 23 04502 g004
Figure 5. Direct and sequential alchemical transformations of IVPC into IVd. Computational Δ Δ G b values relative to alchemical transformations IVPCIVa, IVPCIVb and IVPCIVd performed with complexes parameterized with coordinated FeII are shown near the corresponding arrows. Question marks indicate transformations for which the relative binding energies, Δ Δ G b   I V a I V d and Δ Δ G b   I V b I V d can be arithmetically obtained from the other alchemical data shown in the figure, in addition to by performing the corresponding alchemical transformations.
Figure 5. Direct and sequential alchemical transformations of IVPC into IVd. Computational Δ Δ G b values relative to alchemical transformations IVPCIVa, IVPCIVb and IVPCIVd performed with complexes parameterized with coordinated FeII are shown near the corresponding arrows. Question marks indicate transformations for which the relative binding energies, Δ Δ G b   I V a I V d and Δ Δ G b   I V b I V d can be arithmetically obtained from the other alchemical data shown in the figure, in addition to by performing the corresponding alchemical transformations.
Ijms 23 04502 g005
Table 1. Synthesized analogues of lead compound IVPC. The chemical modifications introduced are distributed throughout the three heterocycles, thiophene, pyridine and pyrimidine, as indicated.
Table 1. Synthesized analogues of lead compound IVPC. The chemical modifications introduced are distributed throughout the three heterocycles, thiophene, pyridine and pyrimidine, as indicated.
CompoundChemical StructureMW (g/mol)Modified Heterocycle
IVPC Ijms 23 04502 i001303.4-
IVa Ijms 23 04502 i002319.4Thiophene
IVb Ijms 23 04502 i003319.4Pyridine
IVc Ijms 23 04502 i004361.4Thiophene
IVd Ijms 23 04502 i005335.4Thiophene and Pyridine
IVe Ijms 23 04502 i006323.8Pyridine
IVf Ijms 23 04502 i007319.4Pyridine
IVg Ijms 23 04502 i008314.4Pyridine
IVh Ijms 23 04502 i009451.6Thiophene and Pyrimidine
IVi Ijms 23 04502 i010400.5Thiophene and Pyrimidine
IVj Ijms 23 04502 i011415.5Thiophene and Pyrimidine
IVk Ijms 23 04502 i012401.5Thiophene and Pyrimidine
IVl Ijms 23 04502 i013359.5Thiophene and Pyrimidine
IVm Ijms 23 04502 i014316.4Pyridine (substitution)
IVn Ijms 23 04502 i015338.5Pyridine (substitution)
IVo Ijms 23 04502 i016342.5Pyridine (substitution)
IVp Ijms 23 04502 i017379.5Pyridine (substitution) and Pyrimidine
IVq Ijms 23 04502 i018361.4Thiophene
Table 2. Calculated and experimental relative binding free energies, and thermostabilizing effect of compounds on PAH.
Table 2. Calculated and experimental relative binding free energies, and thermostabilizing effect of compounds on PAH.
CompoundIn Silico ∆∆Gb IVPCIVx (kJ/mol) aExperimental ∆∆Gb (kJ/mol) bΔTm
(°C) c
with FeIIwith FeIII
IVPC---5.0 ± 0.6
IVa−1.62 ± 0.18−0.66 ± 0.25−1.25 ± 1.195.5 ± 0.8
IVb2.12 ± 0.79−0.08 ± 0.231.89 ± 1.601.2 ± 0.9
IVc−2.78 ± 1.70−1.65 ± 0.93−1.92 ± 1.316.5 ± 0.8
IVd0.63 ± 0.37−0.73 ± 0.190.42 ± 1.298.6 ± 0.5
IVe−3.14 ± 0.19−3.98 ± 0.09−2.50 ± 1.128.0 ± 1.4
IVf−1.06 ± 0.37−3.04 ± 0.33−0.88 ± 1.254.1 ± 0.8
IVg0.21 ± 0.37−4.01 ± 0.370.09 ± 1.177.3 ± 0.8
IVhn.d. cn.d.c−0.10 ± 1.13−4.1 ± 0.7
IVi0.64 ± 3.21−22.94 ± 4.36−2.46 ± 1.5514.1 ± 0.6
IVjn.d. c−9.52 ± 8.951.65 ± 1.10−4.9 ± 0.8
IVk3.01 ± 0.62n.d. c−0.43 ± 1.14−7.4 ± 0.6
IVl8.36 ± 2.64n.d. c0.50 ± 1.081.3 ± 0.6
IVm−3.39 ± 0.80−3.28 ± 0.93−2.03 ± 1.2610.0 ± 0.9
IVn−4.17 ± 0.89−7.33 ± 0.32−2.68 ± 1.2713.6 ± 1.4
IVo−5.99 ± 0.64−3.17 ± 1.09−4.28 ± 1.378.8 ± 0.6
IVp−1.00 ± 2.87−11.85 ± 3.02−3.57 ± 1.1017.9 ± 0.9
IVq−4.86 ± 0.76−3.86 ± 0.31−1.54 ± 1.326.7 ± 0.6
a Relative AFEC binding free energies obtained as ∆∆GbIVPC→IVx = ∆GboundIVPC→IVx–∆GsolvIVPC→IVx. Partial errors and values for the solvated and bound AFEC transformations are given in Table S3. Abbreviation n.d. indicates the energies that could not be determined. b Experimental relative binding energy obtained from the following equation: Δ Δ G b I V P C I V x = R T l n K d I V P C K d I V x   , where K d I V P C and K d I V x were determined by ITC experiments. c Increase in PAH mid-point unfolding temperature in presence of IVPC or analogues (IVx).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Conde-Giménez, M.; Galano-Frutos, J.J.; Galiana-Cameo, M.; Mahía, A.; Victor, B.L.; Salillas, S.; Velázquez-Campoy, A.; Brito, R.M.M.; Gálvez, J.A.; Díaz-de-Villegas, M.D.; et al. Alchemical Design of Pharmacological Chaperones with Higher Affinity for Phenylalanine Hydroxylase. Int. J. Mol. Sci. 2022, 23, 4502. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23094502

AMA Style

Conde-Giménez M, Galano-Frutos JJ, Galiana-Cameo M, Mahía A, Victor BL, Salillas S, Velázquez-Campoy A, Brito RMM, Gálvez JA, Díaz-de-Villegas MD, et al. Alchemical Design of Pharmacological Chaperones with Higher Affinity for Phenylalanine Hydroxylase. International Journal of Molecular Sciences. 2022; 23(9):4502. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23094502

Chicago/Turabian Style

Conde-Giménez, María, Juan José Galano-Frutos, María Galiana-Cameo, Alejandro Mahía, Bruno L. Victor, Sandra Salillas, Adrián Velázquez-Campoy, Rui M. M. Brito, José Antonio Gálvez, María D. Díaz-de-Villegas, and et al. 2022. "Alchemical Design of Pharmacological Chaperones with Higher Affinity for Phenylalanine Hydroxylase" International Journal of Molecular Sciences 23, no. 9: 4502. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23094502

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