Sara K. Murasea,
Nurit Haspel*b,
Luis J. del Vallea,
Eric A. Perpète*c,
Catherine Michauxc,
Ruth Nussinovde,
Jordi Puiggalí*af and
Carlos Alemán*af
aDepartment of Chemical Engineering, ETSEIB, Universitat Politècnica de Catalunya, Av. Diagonal 647, 08028, Barcelona, Spain. E-mail: jordi.puiggali@upc.edu; carlos.aleman@upc.edu
bUniversity of Massachusetts Boston, Department of Computer Science, Boston, MA 02125, USA. E-mail: nurit.haspel@umb.edu
cUnité de Chimie Physique Théorique et Structurale, University of Namur, rue de Bruxelles, 61, 5000 Namur, Belgium. E-mail: eric.perpete@unamur.be
dBasic Science Program, Leidos Biomedical Research, Inc., Cancer and Inflammation Program, National Cancer Institute, Frederick, MD 21702, USA
eSackler Institute of Molecular Medicine, Department of Human Genetics and Molecular Medicine, Sackler School of Medicine, Tel Aviv University, Tel Aviv 69978, Israel
fCenter for Research in Nano-Engineering, Universitat Politècnica de Catalunya, Campus Sud, Edifci C', C/Pasqual i Vila s/n, Barcelona E-08028, Spain
First published on 15th May 2014
Peptide–polymer conjugates made of poly(L-lactide) and L-phenylalanine or L,L-diphenylalanine (F–PLA and FF–PLA, respectively) have been synthesized by the ring-opening polymerization of L-lactide using the peptide fragment as an initiator. The structure of the conjugates was confirmed by 1H NMR, FT-IR, GPC, UV-Vis and CD. Molecular dynamics simulations have been used to identify both the conformational preferences of the FF–PLA conjugate in solution and the potential intramolecular interactions between the peptide and polymer blocks, while TD-DFT calculations have been applied to model the electronic transitions observed by the UV-Vis absorption spectroscopy. Results show that the polymer fragment prefers a random coil or a mix of helix/strand while the peptide fragment tends to have folded and helical conformations. Although the degree of interaction between the two fragments is slightly higher than that reported for other peptide–polymer conjugates, it is small enough to suggest that FF–PLA is a potential candidate to aggregate forming peptide-guided organizations via self-assembly. On the other hand, quantum mechanical calculations have allowed us to identify the π → π* transition, which is typically observed in helical peptides and proteins, as well as the n → π* transition along the N–C–O backbone.
The fabrication of well-defined peptide–polymer conjugates is usually performed using one of the following approaches:12,13 (1) coupling the peptide segments to pre-formed synthetic molecules through one or more reactive sites; (2) growing the polymer from the peptide segment; (3) assembling the peptide on an already formed polymer; and (4) polymerizing short peptides with a polymerizable functionality. In all cases, these approaches are expected to provide hybrid molecules able to adopt 3D organizations, which should be defined by the peptide secondary structure.14 Some of these approaches have been recently used to prepare peptide–PEO conjugates based on phenylalanine (F) homopeptides (e.g. FFFF–PEO,15,16 KLVFF–PEO,17 FFKLVFF–PEO18 and F–ω-hydroxyl-PEO19).
In this work we present the preparation and characterization at the molecular level of the peptide–polymer conjugate resulting from the combination of FF and poly(L-lactide) (PLA), hereafter denoted FF–PLA. Conjugates based on PLA, which is a biodegradable material obtained from renewable resources, are expected to present important medical and biotechnological applications because of their interesting properties. PLA can be obtained by polycondensation from lactic acid or by ring-opening polymerization (ROP) of its cyclic dimer, lactide. Lactic acid is a new renewable resource derived from the starch of either corn or sugar beets, which is fermented to form glucose and then converted into lactic acid. Although the variety of stereoisomeric polymers that can be obtained from lactic acid or lactide polymerization are complete as single-use packaging materials, films and other commodity uses, they find their most outstanding use in biomedical devices in which they slowly hydrolyze back to lactic acid and re-enter the Krebs cycle. Also, these polymers have been employed as scaffolds for tissue engineering and in drug delivery systems. Excellent reviews of many chemical and practical aspects of PLA have been published.20–22
In a recent study, the synthesis of L-phenylalanine-terminated PLA, F–PLA, was described using a three-step procedure.23 First, the hydroxyl-terminated PLA was synthesized through the ROP of L-lactide initiated by n-butanol under the presence of tin(II) ethylhexanoate. Subsequently, the complete capping of the hydroxyl end group of PLA with Boc-L-phenylalanine was achieved by using a mixed anhydride of Boc-L-phenylalanine under the catalysis of 4-(1-pyrrolidinyl) pyridine. Finally, the free amino end group was obtained by removal of the t-butoxycarbonyl group through trifluoroacetic acid treatment under anhydrous conditions. Unfortunately, this product was not characterized but used as macroinitiator for the synthesis of poly(L-lactide)-b-poly(L-lysine) block copolymers. More recently, this strategy has been also used to prepare other block copolymers, as for example those derived from poly(ester amide) and polyester.24
This work is focused on the synthesis and basic characterization of thermal and electronic properties of F–PLA and FF–PLA. The process followed for this purpose is the contrary to that reported in ref. 23 and discussed above, since peptide fragments have been used as initiators of the polymerization reactions (i.e. the polymer has been grown from the peptide segment). Furthermore, the molecular structure of FF–PLA has been examined at the atomistic level using Molecular Dynamics (MD) simulations to determine the organization of the FF and PLA fragments as well as to ascertain the level of interactions among them. After this, the size of the model used for classical MD simulations has been conveniently reduced to investigate the electronic transitions provoked by the FF fragment through quantum mechanical calculations. More specifically, time-dependent density functional theory (TD-DFT) calculations have been applied to identify the electronic transitions detected by UV-Vis spectroscopy.
1H NMR spectra were recorded with a Bruker AMX-300 spectrometer operating at 300.1 MHz. Chemical shifts were calibrated using tetramethylsilane as an internal standard and CDCl3 (δ(1H) = 7.24 ppm) was used as the solvent.
Infrared absorption spectra were recorded with a Fourier Transform FTIR 4100 Jasco spectrometer in a 4000–600 cm−1 range. A Specac-Teknokroma model MKII Golden Gate attenuated total reflection (ATR) set-up was also employed.
Calorimetric data were obtained by differential scanning calorimetry with a TA Instruments Q100 series equipped with a refrigerated cooling system (RCS). Experiments were conducted with around 5 mg of sample under a flow of dry nitrogen and at a heating rate of 20 °C min−1. Samples were previously quenched from the melt state at the maximum rate allowed by the equipment. Calibration was performed with indium.
Circular dichroism (CD) spectra were measured in a 1 mm path length using a Jasco J-815 spectropolarimeter (Jasco, Japan) and analyzed with the supplied Spectra-Manager software. The temperature was kept constant at 20 °C using a temperature controlled water bath. The samples were dissolved in HFIP.
Ultraviolet spectra were recorded with a Shimadzu UV-3600 ultraviolet-visible spectrometer from polymer solutions in HFIP.
:
1. Reaction was performed under magnetic stirring in an oil bath at a temperature of 135–140 °C. The reaction mixture was firstly kept under a nitrogen stream for ten minutes and then the tube was sealed. Polymerization process was monitored by 1H NMR to confirm the total lactide conversion. After the required time (ca. 240 h for both H–Phe–OH and H–Phe–Phe–OH initiated reactions), reaction tubes were cooled in ice to stop polymerization. The obtained pale solids were repeatedly washed with methanol and recovered by centrifugation. Finally, polymers were purified by re-precipitation with methanol from chloroform solutions.
![]() | ||
| Fig. 1 Synthesis of F–PLA and FF–PLA conjugates initiated by L-phenylalanine (H–Phe–OH) and L,L-diphenylalanine (H–Phe–Phe–OH). | ||
430 atoms and was placed in an orthorhombic box of dimensions 62 × 104 × 112 Å3.
All MD simulations described in this work were performed by using AMBER ff03 potential28 and the NAMD computer package.29 The Berendsen thermostat-barostat30 was used to control temperature and pressure, with a relaxation time of 800 ps. The SHAKE31 algorithm was applied to bonds involving hydrogen atoms. The integration time step was set to 2 fs. The cutoff distance for nonbonding interactions was 14 Å. Particle Mesh of Ewald (PME)32 was applied for computing electrostatic interactions beyond the cutoff distance. The real space term was determined by the van der Waals cut off (14 Å), whereas the reciprocal term was estimated by interpolation of the effective charge into a charge mesh with a grid thickness of one point per cubic Å. Periodic boundary conditions were applied in all simulations in this paper using the nearest image convention.
The equilibration protocol consisted of the following: initially, the potential energy of the system was minimized for 2000 steps using steepest descent algorithm. Next, the system was heated up to 500 K during 500 ps in a NVT-MD simulation. After this, 500 ps at 298 K of NVT-MD were run. Finally, 500 ps of NPT-MD were run for density relaxation with a pressure of 1 atm. The box size was optimized until the density of the system reached approximately 1.6 g cm−3, which is the density of the HFIP solvent. The production runs were carried out in constant volume.
The conformational exploration was performed using a Simulated Annealing/Molecular Dynamics (SA-MD) protocol. This strategy used the final outputs of the density relaxation, heated them to 900 K and cooled them to 300 K over 10 ns, saving a snapshot every 2 ps (5000 per cycle). The potential energy of the obtained structures was relaxed through 500 steps of conjugate gradient minimization in order to resolve minor clashes and reach a nearby local minimum. The five lowest potential energy conformations were selected as starting points for the next round. The protocol was run for 8 rounds. Each round produced 3250 snapshots, a total of 26
000 structures being obtained.
In recent years some powerful and advanced sampling methods has been tested,33 even though these are significantly demanding in terms of computer resources. In this work the choice of the SA-MD methodology is based on the fact that it has proven useful for conformational exploration of flexible molecules in solution,34,35 for which sophisticated methods are not required. These advantages include the ability of the method to overcome high potential energy barriers and to proficiently explore flat potential energy surfaces compared to plain MD, which is highly dependent on the initial geometry. To make sure that the sampling was sufficient, we followed a conformational clustering protocol.34,35 After each round the conformations were clustered using the VMD clustering protocol36 with a 6.5 Å root-mean square-deviation (RMSD) threshold on the C-alpha atoms. The VMD clustering protocol uses the Quality Threshold method that determines the number of clusters in advance. We determined the number of clusters per round to be the smallest number that assigns a cluster to at least 90% of the conformations. Fig. 2 shows the number of clusters as a function of the number of SA-MD rounds. As seen, the number of clusters stabilized on approximately 30 clusters after 8 rounds.
Subsequently, the electronic absorption spectrum was computed for the FF–4PLA, in order to compare to the experimental UV-Vis data and help interpreting them. It is well-known that such simulations can be sensitive to the selected level of approximation40 and several DFT functionals as well as Pople's BS have consequently been tested on the above-optimized structure to ensure the simulations reliability. On the BS side, we considered extending the basis set in terms of zeta, polarization and diffusion functions, while on the DFT side, various blends of exchange–correlation formalisms were assessed with respect to the stability of respective results.
:
1) allowed us to obtain peptide–polymer conjugates with a relatively high molecular weight. Specifically, number average molecular weights of 49
000 g mol−1 and 66
000 g mol−1 and polydispersity indices of 1.41 and 1.48 were attained for samples coming from H–Phe–OH (F–PLA) and H–Phe–Phe–OH (FF–PLA), respectively.
1H NMR spectra were monitored during ROP to assure a total conversion since signals attributed to the lactide monomer (i.e. 5.37 and 1.68 ppm) were clearly differentiated from those corresponding to the polymer in the conjugate (Fig. 3). Long reaction times were needed for both polymerizations, with the ROP reaction being slightly faster when H–Phe–Phe–OH was employed (e.g. the time necessary to achieve a conversion of 70% increased from 80 to 160 h when H–Phe–OH was used). PLA in conjugates were obtained with a yield close to 70% after purification by re-precipitation when the reaction was performed for 240 h, a reaction time that assured a practically complete conversion without any sign of thermal degradation.
1H NMR spectra of synthesized conjugates showed the typical signals of PLA: a quadruplet at 5.19, 5.17, 5.15 and 5.13 ppm corresponding to the proton of the methine unit, and the doublet at 1.76 and 1.73 ppm corresponding to the protons of methyl groups. Furthermore, small signals attributed to the terminal lactoyl unit (4.4–4.3 ppm) and also to the phenylalanine terminal units (aromatic protons at 7.3–7.2 ppm) could also be detected as shown in the insets of Fig. 3.
FTIR spectra of synthesized samples were similar and showed the typical absorption bands of PLA41–43 (Fig. S1†). Thus, the most intense signals were observed at 1749 cm−1 ν(C
O); 1452 cm−1 δas (CH3); 1361 cm−1 δ(CH); 1183 cm−1 νas(C–O–C) + ras(CH3) for the crystalline phase; 1128 cm−1 rs(CH3); 1085 cm−1 νs(C–O–C); and 1043 cm−1 ν(C–CH3).
Polymers in the synthesized conjugates are semicrystalline, as shown in Fig. S2† where the DSC heating scan of a representative melt quenched sample (i.e. FF–PLA) is displayed. Typical features such as a glass transition with a significant relaxation peak, an exothermic cold crystallization peak and a complex melting peak close to 140 °C were detected. In fact, a degree of crystallinity of 30–33% could be estimated for both F–PLA and FF–PLA samples from the melting enthalpy (106 J g−1) for a 100% crystalline sample.44 Small signals of FTIR spectra observed at 1209 cm−1 (νas (C–O–C) + ras(CH3) for a 107 helical conformation) and 921 cm−1 (coupling of the C–C stretching and CH3 rocking mode for long segments with a 107 helical conformation) were in agreement with a preferred molecular conformation of PLA.41–43,45,46
UV-Vis spectra clearly demonstrated that ROP was effectively initiated by the phenylalanine derivatives since typical absorption peaks associated with the aromatic rings could always be observed. Thus, Fig. 4 shows that the characteristic five peaks at 246, 250, 256, 262 and 266 nm of the H–Phe–Phe–OH compound were still detected in the spectra of the synthesized FF–PLA samples. As expected, the considered peaks were more significant when polymerization was initiated by the dipeptide. It is also remarkable that commercial PLA did not show any significant absorption in the region comprised between 246 and 266 nm, and consequently absorbance peaks of synthesized PLA samples should be unambiguously assigned to linked phenylalanine units.
We used CD spectroscopy to characterize the secondary structure of PLA, H–Phe–OH, H–Phe–Phe–OH, F–PLA and FF–PLA dissolved in HFIP (Fig. 5). Amorphous and crystalline PLA dissolved in HFIP exhibited a positive CD band at 210 nm similar to that registered in acetonitrile solution.47 The CD spectra of H–Phe–OH and H–Phe–Phe–OH show one positive band at wavelength of 226 nm and 237 nm, respectively, and several negative several dichroic bands, the ones centered at 222 and 208 nm being typically attributed to n–π* and parallel-polarized π–π* excitonic transitions in peptides. These two negative bands disappear upon the conjugation of these peptides to PLA (Fig. 5, inset). A positive band at 230 nm is observed for F–PLA, this position being very similar to that registered for H–Phe–OH (Δ = 4 nm). For FF–PLA the positive band is centered at 227 nm, reflecting that the difference with respect to the individual peptide (Δ = 10 nm) is slightly higher than for F–PLA. Although CD spectra cannot provide high resolution information on the secondary structure, these results suggest that the interactions between the peptide and polymer blocks are practically inexistent in F–PLA and weak in FF–PLA.
![]() | ||
| Fig. 5 CD spectra in HFIP of commercial PLA at 0.05 mg mL−1 (i), H–Phe–OH at 5 mg mL−1 (b), F–PLA at 5 mg mL−1 (c), H–Phe–Phe–OH at 5 mg mL−1 (d) and FF–PLA at 2.5 mg mL−1 (e). | ||
The average number of structures per cluster is 23 ± 16. The lowest energy conformations in each cluster were within 26 kcal mol−1 compared to the lowest energy structure detected. Fig. 6 shows a histogram of the energy distribution of all the 10% low-energy conformations in the clusters. The energy units are relative to the lowest energy conformation. As seen, the energies distribute between 0 kcal mol−1 (above the lowest energy conformation) and about 50 kcal mol−1, with a peak around 30–40 kcal mol−1 above the minimum. The energy distribution can give indications about the conformational accessibility of the system. This is based on the fact that the probability of a state to exist at a certain temperature can be expressed as a function of the energy difference between that state and the lowest energy state through a Maxwell–Boltzmann distribution. Moreover, a small number of states within an energy range indicates fewer thermodynamically accessible structures at a given temperature and consequently low conformational variability in terms of energy. In this case, as also shown below, the system exhibits some structural variability.
Fig. 7a shows the Ramachandran plot for flexible backbone dihedral angles of the 10% low-energy conformations of the PLA in FF–PLA, which have been denoted ξ1 and ξ2 (Fig. 7b). As seen, the majority of the conformations exhibit a random coil or a mix of helix/strand. Since we used the L-lactide configuration, the Ramachandran plot of PLA is a mirror image of the typical L-amino acid plot. The ξ1–ξ2 distribution is generally compatible with the allowed regions. Approximately 13% of the dihedral angles are in the helical region (ξ1 between 35° and 90°, and ξ2 between 15° and 70°) and 8% in the extended/semi-extended region (ξ1 between 90° and 180°, and ξ2 between −90° and −180°). Fig. 7c shows the values of the dihedrals associated with the ester bond of the PLA block, which have been denoted as ω (Fig. 7b). As expected, ω distributes mainly around ±180°. Indeed, only 9.6% of the ω angles are in the range of −150° to 150°, and only about 1% of the ω angles are between −120° and 120°.
Fig. 7d represents the Ramachandran plot for flexible backbone dihedral angles of the 10% low-energy conformations of the peptide fragment (φ and ψ in Fig. 7b) in FF–PLA. The analysis has been focused on the second phenylalanine residue since definition of the ψ dihedral is not possible for the first one. As can be seen, the most populated φ–ψ distributions correspond to folded and helical regions, even though regions associated with extended and semiextended conformations are also occasionally visited. The conformational preferences of the N-acetyl-N′-methylamide derivative of FF were explored in a recent study by combining a building procedure with quantum mechanical calculations at the B3LYP/6-31G(d) level.48 The 100 initial geometrical arrangements for FF, which were constructed using a combinatorial growth of the peptide chain by considering 10 minimum energy conformations for F, converged over a total of 78 minimum energy structures. However, clustering analysis converged 78 such minimized structures into 20 different conformations. For each cluster the conformational characteristics of the backbone were practically identical, differences being only found in the arrangements of the side groups. Among the 20 basic organizations, the conformation that presented the lowest energy corresponded to a double γ-turn (also denoted 2.27 ribbon), which was the outcome of replicating the most favored arrangement of F. This regular organization has only been observed in short peptides.49 In the three following organizations each phenylalanine residue adopted a different conformation. Thus, the second organization presented a folded structure that was the outcome of forming an intramolecular hydrogen bond between the C
O and NH moieties of the blocking groups. This particular arrangement was associated with a β-turn motif or with an embryonic helical organization that would render a 310 helix if the peptide could be prolonged beyond the two residues of phenylalanine. The third organization resulted from the combination of distorted extended and γ-turn motifs while the fourth showed a hydrogen bond associated with the γ-turn motif of the second phenylalanine residue and an inter-residue N–H⋯π interaction. The rest of the organizations resulted from repetition of two extended or semi-extended conformations, from the combination of the latter conformation with a folded pattern, or from the repetition of two helical conformations. Details about all such structures were provided in ref. 48. It is worth noting that the conformational profile predicted by quantum mechanical calculation for the N-acetyl-N′-methylamide derivative of FF is fully consistent with the conformational preferences displayed in Fig. 7d.
The crystal structure of H–Phe–Phe–OH was reported by Görbitz more than one decade ago.50 Görbitz discovered that this small peptide crystallizes with hydrogen bonded head-to-tail chains forming hydrophilic channels embedded in a hydrophobic matrix created by the peptide side chains. The sequence formed by two consecutive phenylalanine residues was proposed to be an attractive model for membrane channels due to the substantial size of the hydrophilic channels. Interestingly, the conformation adopted by the unblocked peptide in the solid state, which was identified in ref. 48 as a minimum energy conformation, is one of the visited conformations depicted in Fig. 7d. Indeed, an overall comparison of Fig. 7d with the conformational profile of N-acetyl-N′-methylamide derivative of FF48 suggests that the influence of the polymer on the peptide fragment conformation in the conjugate is relatively small.
Representative low-energy conformations obtained for FF–PLA by our search and clustering process are shown in Fig. 8a. As seen, the low-energy conformations are partially-folded, mostly with a small hydrophobic core, and exhibit a mix of secondary structure and random coil conformations, in accordance with the dihedral angle analysis shown above. Fig. 8b shows the pairwise RMSD between the energy minima, evidencing values ranging from 2 to 12 Å. This corresponds to a wide distribution of conformations. The RMSD was calculated for all but the hydrogen atoms.
Fig. 9a and b shows the radius of gyration (Rg) and the end-to-end distance (de–e), respectively, for each PLA block of the low-energy conformations, measured between the O1 atom of the first repeat unit and the C-terminus. The average Rg and de–e were 20.1 ± 2.5 Å and 52.9 ± 8.0 Å, respectively. This result indicates that most of the conformations adopted by the PLA block in HFIP solution were generally rather compact with respect to the size of the extended polymer, which is expected considering its hydrophobic nature.
![]() | ||
| Fig. 9 (a) Radius of gyration (Rg) and (b) end-to-end distance (de–e) of the low-energy conformations. (c) Representative interactions between FF and PLA blocks. | ||
Analysis of the interactions between the FF and PLA blocks in the conjugate reveals that the non-terminal phenylalanine residue was frequently involved (∼30% of low energy conformations) in a hydrogen bond with the main chain of the PLA block (i.e. usually with the repeat unit immediately after it). Another hydrogen bond between the main chain of the non-terminal phenylalanine residue and the second PLA repeat unit was observed in approximately 13% of the low-energy conformations while the interaction between the C-terminus and the second PLA repeat unit appears in approximately 17% of the conformations (see Fig. 9c). However, no other significant inter-block secondary interaction (i.e. with a population higher than 5%) was detected suggesting that FF and PLA tend to organize independently. Although the interactions between FF and PLA fragments are scarce, these are more frequent than those recently observed for FF–PEG, where PEG refers to poly(ethylene glycol), in which the two counterparts remained independent, preserving their inner physical properties.51
Results obtained in this section suggest that FF–PLA possesses an enormous potential to form self-assemblies for biomedical applications via peptide-guided organization. The self-assembling capabilities of FF are well known. For example, Boc–FF52 and Fmoc–FF53 [where Boc and Fmoc refer to tert-butyloxycarbonyl and N-(fluorenyl-9-methoxycarbonyl, respectively)] organize in microspheres and nanofibrils, respectively. While FF is able to assemble into peptide nanotubes,54 disassembly can be achieved depending on the solution that surrounds the structure. No major changes are expected in the ability of FF to assemble, as their interactions with the polymer fragment are scarce. We are aware that the high crystallinity of PLA can interfere with the FF assembly since it is well known that simultaneous polymer crystallization and peptide organization become less likely.55 However, the synthetic procedure used in this work allows controlling the molecular weight of PLA, the crystallinity of PLA decreasing with the molecular weight. Our research is currently oriented towards the self-assembly of FF–PLA using polymer fragments of low molecular weight.
Fig. 10a represents the variation of σav with n for FF–nPLA models. As expected, σav increases with decreasing n. More specifically, σav ranges from 9.7° to 22.0° and from 15.6° to 40.6° for n = 16 and 4, respectively. Although in some cases the reduction of n represents a significant increment of σav (e.g. for models #3 and #6 the value of σav increases from 11.7° to 20.6° and from 15.9° to 35.4°, respectively, when n decreases from 10 to 4), the computational effort required to carry out TD-DFT calculations on FF–nPLA decreases by a huge amount with such reduction. Fig. 10b depicts the variation of the relative energy (ΔE) with n for the six calculated models. Although the energy gap (i.e. ΔE of the most unfavored model) decreases from 31.2 to 14.4 kcal mol−1 when n decreases from 16 to 4, the ΔE order remains practically identical. Thus, the most remarkable change occurs for models #1, #2 and #3, which are practically isoenergetic when n = 4 while they are separated by an energy gap of 5.4 kcal mol−1 when n = 16. On the other hand, it is worth noting that σav is 15.6°, 17.3° and 29.6° for the three more stable models of FF–4PLA. Fig. 10c represents more stable reduced models, which are within an ΔE interval of 0.2 kcal mol−1 (Fig. 10b), evidencing that the inter-block interaction between FF and PLA is preserved after reduction of the number of polymer repeat units from 40 to 4.
Typically for FF–4PLA models, three more or less pronounced stabilizing interactions appear in the structures: (i) a hydrogen bond between the C
O and N–H, as classically observed in a β-turn; (ii) a quadruple interaction between both phenyl groups. Several recent experimental studies have emphasized the importance of aromatic interactions in stabilizing helices and beta-hairpins in designed synthetic peptides;56,57 (iii) A C–H⋯π interaction between the methyl and one of the phenyl groups. Such interaction is a weak attractive force occurring between CH groups and π-systems. They are known to make important contributions to bio-molecular structure and function.58–60 The presence of such weak interactions between the FF and PLA blocks is fully consistent with the CD spectra displayed in Fig. 5.
The simulation of the electronic excitations has been carried out for the most stable conformer identified for the FF–PLA tetramer at the structural study stage, namely FF4–PLA, by using the well-established TD-DFT computational scheme. The same WB97XD level of approximation was first selected, and the accuracy of the spectrum was then increased by enlarging the original BS, i.e. Pople's 6-311+G(d,p) was used, and by considering a Polarizable Continuum Model (PCM) to take the solvation effects into account. Doing so allowed keeping in the limits of accepted precision for excited states calculation in this upper part of the UV-Vis absorption spectra.
Indeed, as depicted in Fig. 12, FF4–PLA UV spectrum nicely reproduces the shape of the observed FF–PLA absorption spectrum with a consistent shift of ∼35 nm with respect to the de-convoluted experimental data (see Fig. 11). That is, we have the five excitation wavelengths computed (A to E in Fig. 12) at 208, 215, 222, 229 and 232 nm for observed peaks (1 to 5 in Fig. 12) at 246, 251, 257, 263 and 268 nm, with well-reproduced intervals between the spectral lines.
The topology of the molecular orbitals involved in these transitions has also been investigated, and two major contributions have been highlighted (Fig. 13). First, the typical π → π* arising in this part of the spectrum were indeed observed. Second, a n → π* transition can be identified along the N–C–O backbone, corresponding to one of the two essential transitions observed in a peptide bond. Consistently with Bulheller et al.,61 the non-bonding orbital is centered about one of the lone pairs of the oxygen in CO while the anti-bonding orbital lies along the peptide bond. In an asymmetric molecule such a transition can show very low total absorption of light but high optical activity, as observed for helical polypeptides and proteins. Besides, n → π* interactions between carbonyl groups have been shown to play a role in protein folding,62 while carbonyl–carbonyl interactions between adjacent backbone amides have indeed been implicated in the conformational stability of proteins.
MD simulations of FF–PLA in HFIP revealed that the peptide fragment retains the intrinsic conformational preferences of diphenylalanine. This feature is consistent with the relatively scarce interactions found between FF and PLA blocks, which essentially consist of hydrogen bonds between the non-terminal phenylalanine residue and the L-lactide unit immediately after it. Thus, PLA tend to organize independently, which is a prerequisite for the formation of peptide-guided assemblies, corroborating the interpretation of CD spectra. However, in order to promote the formation of nanostructures through FF-guided assemblies, the crystallinity of PLA should be reduced by controlling the molecular weight during the ROP. Thus, future investigations on FF–PLA should be focused on exploiting the conjugate by avoiding the competition between the intrinsic tendency of PLA to crystallize and the intrinsic tendency of FF to organize in nanostructures.
Finally, TD-DFT calculations were achieved to model the electronic transitions observed by the UV-Vis absorption spectroscopy of the most stable FF–PLA conformation. They confirm the five excitations experimentally observed, and thanks to the involved molecular orbital topology, we were able to assign them to the usual π → π* transitions, as well as to the n → π* transition along the N–C–O backbone.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra01534g |
| This journal is © The Royal Society of Chemistry 2014 |