Next Article in Journal
Biomaterial Inks from Peptide-Functionalized Silk Fibers for 3D Printing of Futuristic Wound-Healing and Sensing Materials
Next Article in Special Issue
Synthesis of Fluorescent, Dumbbell-Shaped Polyurethane Homo- and Heterodendrimers and Their Photophysical Properties
Previous Article in Journal
Ruthenium Complex HB324 Induces Apoptosis via Mitochondrial Pathway with an Upregulation of Harakiri and Overcomes Cisplatin Resistance in Neuroblastoma Cells In Vitro
Previous Article in Special Issue
Effect of Ionization Degree of Poly(amidoamine) Dendrimer and 5-Fluorouracil on the Efficiency of Complex Formation—A Theoretical and Experimental Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

NMR Studies of Two Lysine Based Dendrimers with Insertion of Similar Histidine-Arginine and Arginine-Histidine Spacers Having Different Properties for Application in Drug Delivery

by
Nadezhda N. Sheveleva
1,
Irina I. Tarasenko
2,
Mikhail A. Vovk
1,
Mariya E. Mikhailova
1,
Igor M. Neelov
3 and
Denis A. Markelov
1,*
1
Saint Petersburg State University, 7/9 Universitetskaya Nab, 199034 Saint Petersburg, Russia
2
Institute of Macromolecular Compounds, Russian Academy of Sciences, Bolshoi Prospect 31, V.O., 199004 Saint Petersburg, Russia
3
School of Computer Technologies and Control, Saint Petersburg National Research University of Information Technologies, Mechanics and Optics (ITMO University), Kronverkskiy Pr. 49, 197101 Saint Petersburg, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(2), 949; https://doi.org/10.3390/ijms24020949
Submission received: 29 November 2022 / Revised: 14 December 2022 / Accepted: 20 December 2022 / Published: 4 January 2023

Abstract

:
In this paper we study two lysine-based peptide dendrimers with Lys-His-Arg and Lys-Arg-His repeating units and terminal lysine groups. Combination of His and Arg properties in a dendrimer could be important for biomedical applications, especially for prevention of dendrimer aggregation and for penetration of dendrimers through various cell membranes. We describe the synthesis of these dendrimers and the confirmation of their structure using 1D and 2D Nuclear Magnetic Resonance (NMR) spectroscopy. NMR spectroscopy and relaxation are used to study the structural and dynamic properties of these macromolecules and to compare them with properties of previously studied dendrimers with Lys-2Arg and Lys-2His repeating units. Our results demonstrate that both Lys-His-Arg and Lys-Arg-His dendrimers have pH sensitive conformation and dynamics. However, properties of Lys-His-Arg at normal pH are more similar to those of the more hydrophobic Lys-2His dendrimer, which has tendency towards aggregation, while the Lys-Arg-His dendrimer is more hydrophilic. Thus, the conformation with the same amino acid composition of Lys-His-Arg is more pH sensitive than Lys-Arg-His, while the presence of Arg groups undoubtedly increases its hydrophilicity compared to Lys-2His. Hence, the Lys-His-Arg dendrimer could be a more suitable (in comparison with Lys-2His and Lys-Arg-His) candidate as a pH sensitive nanocontainer for drug delivery.

1. Introduction

Dendrimers are regular hyperbranched monodisperse macromolecules with a well-defined spherical structure. These nanosized macromolecules are widely used in biomedicine [1,2,3,4,5,6] as carriers in gene and drug delivery [7,8,9,10,11,12,13,14]. In the latter case, dendrimers contribute to an increase in the solubility of drugs, a decrease in their toxicity for normal cells, and a prolongation in their circulation time in blood flow [15,16,17]. The development of drugs and gene delivery carriers based on lysine and peptide dendrimers has been described in many studies [18,19,20,21,22]. The usage of these dendrimer carriers could provide a synergistic effect due to the bioactive action of the dendrimers themselves, which have antimicrobial [23,24,25,26,27,28] and antiangiogenic activity [29,30,31,32].
The search for new drugs and methods for their safe delivery, as well as the encouraging results of cytological studies of dendrimers [22,31,32,33,34,35,36,37], inspire researchers to develop new dendrimer macromolecules [38,39]. The possibility of step-by-step synthesis allows control over their structure and composition, as well as permitting the introduction of functional groups into the core, inner, or terminal segments. This functionalization is a reliable way to obtain dendrimers with tailored characteristics for various purposes [40,41,42]. Thus, an overall positive charge of the dendrimer macromolecules is an important characteristic, which is required for effective cellular uptake. The lysine-based dendrimers functionalized with additional lysine and glycine amino acid residues suppressed the growth of cancer cells [22]. Lysine dendrimers with double arginine insertions into spacers between neighboring lysine branching points were effective as siRNA carriers [35]. The introduction of histidine amino acid residues into the structure gives new properties [43] to the dendrimer. One of them is the ability to be deuterated. The proton at the C2 atom of the histidine imidazole ring is replaced by deuterium upon heating [44,45,46]. It showed the possibility of controlled deuterium labeling of the Lys-2His dendrimers. In addition, the chemical structure of these dendrimers remains stable after deuteration [47]. Another important effect is the change of the global conformation of histidine-containing dendrimer with the changing of the pH level. We have found that the Lys-2His dendrimer undergoes a conformational transition from a swollen conformation at low pH to a collapsed conformation at normal pH [48]. This circumstance can play an important role for using the histidine-containing dendrimer as a nanocontainer.
The above-mentioned results confirm that peptide dendrimers modified with various amino acid residues have great potential for biomedical applications. In particular, a change in the size of the Lys-2His dendrimer in the collapsed conformation can lead to the aggregation of dendrimer macromolecules, which, of course, can adversely affect their biomedical application. That is why this article is devoted to the synthesis and study of new Lys-His-Arg and Lys-Arg-His dendrimers with an intermediate chemical structure between the Lys-2His dendrimer (having a collapsed conformation at normal pH) and the Lys-2Arg dendrimer (having a swollen conformation at normal pH). In contrast to the lysine and lysine-based peptide dendrimers synthesized and studied by us earlier [47,48,49,50,51,52,53,54,55,56,57,58], the considered dendrimers contain two different types of amino acid residues between the branching points: histidine and arginine. The only difference between the two dendrimers is the order in which the histidine and arginine residues are inserted into spacers between lysine branching points. NMR spectroscopy and relaxation methods were used for their investigations. In this work, it has been shown that this method is sensitive to the collapsed conformation of the peptide dendrimer.

2. Results

2.1. Analysis of 1H and 13C NMR Spectra

Structural analysis of the newly synthesized second-generation dendrimers Lys-Arg-His and Lys-His-Arg was carried out using the methods of one-dimensional and two-dimensional NMR spectroscopy. Figure 1 and Figure 2 show the 1H and 13C NMR spectra of the Lys-Arg-His dendrimer. Figure 3 and Figure 4 show the 1H and 13C NMR spectra of the Lys-His-Arg dendrimer. Using 1H NMR spectra, the integral values of the peaks were calculated to estimate the contribution of protons of different groups to the signal.

2.2. Lys-Arg-His

On the proton NMR spectrum in Figure 1, we can see four regions where signals are observed. The signals in the aromatic region 8.30–7.00 ppm obviously refer to protons in the imidazole rings of the histidine residues. The signals in the region from 4.63 to 3.90 ppm belong to CH groups. The peaks in the range 3.30–2.90 ppm refer to the CH2 groups connected with nitrogen atoms or imidazole ring. The group of overlapping signals at 1.90–1.10 ppm corresponds to the CH2 groups of the aliphatic part of the dendrimer.
The 13C NMR spectrum of the Lys-Arg-His dendrimer is presented in Figure 2. The signals in the range of 178–170 ppm belong to carbons in carboxyl groups (symbols a, d, j, l, n in Figure 2). The peak at 156.70 ppm refers to quaternary carbon in guanidine groups of the arginine residues. The signals at 134.67, 130.30 and 117.24 ppm are attributed to carbons in the imidazole rings (symbols w, u, v, respectively, in Figure 1 and Figure 2). The signals in the range of 54–52.85 ppm belong to the carbon atoms in the CH groups. The peaks at 49.47 (symbols z, Figure 1 and Figure 2) and 39.07 ppm (symbols i and s, Figure 1 and Figure 2) refer to the CH2 groups adjacent to nitrogen atoms. The signals from the carbon atoms of the dendrimer’s aliphatic part are located in the region 30.65–16.80 ppm.
The 2D 1H-13C HMBC and HSQC spectra of this dendrimer are presented in Figures S1 and S2, respectively, in Supplementary Materials. Here, we present only the main results. It was found that the peaks at 8.10 and 7.10 ppm belong to the protons at the C2 carbon (symbol w, Figure 1) and at the C4 carbons (symbol v, Figure 1) of imidazole rings, respectively. We have the corresponding 1H-13C HSQC cross-peaks for the carbon atoms w (8.13, 134.92) and v (7.11, 117.23) (Figure S1 in Supplementary Materials). The cross-peaks (8.12, 117.11), (8.12, 130.37), (7.11, 130.40), and (7.10, 134.83) in the 1H-13C HMBC indicate a connection between the hydrogen and carbon atoms inside the imidazole ring In addition, the chemical shifts of carbon atoms in imidazole rings 134.67 (symbol w), 130.30 (symbol v) and 117.24 (symbol u) ppm are similar to the chemical shifts of these atoms in L-histidine: 138.87, 134.42 and 119.54 ppm, respectively.
The 1H-13C HMBC cross-peaks (3.11, 117.18), (3.11, 130.28) (Figure S2 in Supplementary Materials) confirm an interaction between carbons in the imidazole ring, with a proton at 3.11 ppm (symbol t, Figure 1). Then, according to the 1H-13C HSQC cross-peak (3.12, 27.52) the signal at 27.45 ppm refers to the CH2 groups connected to the imidazole rings (symbol t, Figure 1 and Figure 2) in histidine residues.
Cross peaks (3.15, 40.55) on 1H-13C HSQC and (3.14, 24.99), (3.14, 28.33) and (3.14, 156.65) on the 1H-13C HMBC spectra confirm that the carbons with the chemical shifts at 40.55 (symbol z, Figure 2), 24.99 (symbol y, Figure 2), 28.33 (symbol x, Figure 2), and 156.65 ppm (symbol u’, Figure 2) belong to arginine residues.
Further, based on the identified peaks we assigned the rest of the signals (see Supplementary Materials). In addition, we clarified that the protons of the CH2 groups marked by symbols i, t, z contribute to signal at 3.13 ppm (Figure 1). The signal at 2.97 ppm is attributed to protons of the CH2 groups (symbol s, Figure 1) adjacent to N-atoms in terminal lysine segments.

2.3. Lys-His-Arg

The Lys-His-Arg dendrimer differs from the previous one in the order of insertion of histidine and arginine moieties, so we can see from Figure 3 and Figure 4 that the proton and carbon spectra are very similar to those for the Lys-Arg-His dendrimer. However, some differences should be noted. On the 1H spectrum (Figure 3), three peaks can be clearly observed in the region of 3.27–2.85 ppm. According to the 1H-13C HSQC spectrum (Figure 5 and Figure S3) and the calculated integral values, the protons CH2-(N) groups of the inner Lys (symbol i, Figure 3) and side Arg segments (symbol z, Figure 3) make the main contribution to the peak at 3.11 ppm. The contribution from the protons of the CH2 groups connected to imidazole rings in the side histidine segments predominates (symbol t, Figure 3) at 3.01 ppm. The peak at 2.96 ppm is attributed to the protons of the CH2-(N) groups of the terminal lysine segments (symbol s, Figure 4). In the case of Lys-Arg-His, only two peaks are observed in this region: a peak of the CH2-(N) groups of the terminal lysine segments (symbol s, Figure 1) at 2.97 ppm, and a common peak for other considered CH2 groups (symbols i, z and t, Figure 1) at 3.13 ppm.
Detailed analyses of the 1H-13C HMBC and HSQC spectra are shown in Figures S3 and S4, respectively, in Supplementary Materials.
In the Lys-His-Arg dendrimer, the signals from the protons of the imidazole ring (symbols w and v in Figure 3) shifted towards upfield compared to the Lys-Arg-His dendrimer. The difference is about 0.1 ppm. At the same time, the 13C spectrum (Figure 4) shows a slight increase in the chemical shifts of the carbon atoms of the imidazole ring w and u, as well as the carbon atoms t of the CH2 groups of the histidine segments. It is quite possible that the imidazole rings are geometrically close to each other and a pairing effect occurs (Figure 6) [59,60].
Before further analysis, we must become acquainted with the terminology, which we use in this study. As mentioned above, the peaks in the range of 3.13–2.90 ppm correspond to the signals of the CH2 groups connected to nitrogen atoms: “inner” groups in the inner lysine (symbol i) segments, “side” groups in the side arginine (symbol z), and histidine (symbol t, bonded to imidazole ring) segments and “terminal” groups in the terminal lysine (symbol s) segments. Note that we use these groups to study orientational mobility in peptide dendrimers, which will be discussed below.
In order to confirm the assumption about the presence of paring between the imidazole rings of neighboring histidine residues in the Lys-His-Arg dendrimer, let us consider for comparison the 1H spectra of the studied dendrimers and the Lys-2Arg and Lys-2His dendrimers in the range of 3.13–2.90 ppm in Figure 7. As can be seen, the shape and position of the peaks for the dendrimers Lys-Arg-His and Lys-2Arg are similar and two main peaks are observed at 3.15–3.05 ppm (inner and side groups) and 2.95–2.90 (terminal groups). In the case of the Lys-His-Arg dendrimer, splitting of the peak for inner and side groups occurs, which leads to the appearance of an additional peak at about 3.0 ppm. A similar situation is observed for the Lys-2His dendrimer. We believe that such a change in the spectra of dendrimers is also associated with the presence/absence of the pairing effect between the imidazole rings. This assumption can be confirmed using the spectrum of the Lys-2His dendrimer at low pH 1.1 (Lys-2Hisp) in which the imidazole rings are charged. At the same time, according to the results of atomistic MD simulation of the Lys-2His and Lys-2Hisp dendrimers, a pairing effect between imidazole rings is observed at a distance of 0.4 nm for Lys-2His, and is absent for Lys-2Hisp [61]. Figure 7 shows that in the spectrum of Lys-2Hisp there is practically no additional peak at 3.0 ppm.
Thus, we conclude that in the case of the Lys-Arg-His dendrimer, the pairing effect is absent, while in the case of the Lys-His-Arg dendrimer, the pairing effect is observed. It is possible that this is the effect of the formation of a collapsed conformation of the Lys-2His dendrimer, as well as the Lys-His-Arg dendrimer (which is presented below using NMR relaxation data).

2.4. Local Orientational Mobility

To study the local orientational mobility of the Lys-His-Arg and Lys-Arg-His dendrimers, we used the signals from the inner, side, and terminal groups (Figure 1, Figure 3 and Figure 7). According to the spectral analysis, we observed a resolved signal from the protons of the terminal groups and average signals from the inner and side groups, since they overlap. Among the dendrimers under consideration, Lys-2Arg is an exception, since there is a separate peak for inner groups [48,51,62]. This was taken into account when considering the temperature dependences of the spin-lattice relaxation rate 1/T1H obtained (Figure 8).
In the framework of the dipole–dipole relaxation mechanism of 1H nuclei (protons), the 1/T1H function can be written as [63,64,65,66,67]:
1 / T 1 H = A 0 ( J ( ω H , τ i ) + 4 J ( 2 ω H , τ i ) )
where ωH is the cyclic resonance frequency (2πf0) for 1H nuclei; A0 is a constant that does not depend on temperature and frequency; and J is the spectral density which corresponds to Fourier transform from P2 orientational autocorrelation functions averaged over groups contributing to a corresponding peak. In the general case, the spectral density function for 1H nuclei has the form:
J ( n ω H , τ i ) = i C i τ i 1 + ( τ i n ω H ) 2
where τi and Ci are ith correlation times and their contribution to J, respectively, and n = 1, 2. The correlation time is determined by Arrhenius dependence
τ = τ 0 exp ( E a k b T )
where Ea is the activation energy for the chosen group, and T and kb are temperature and Boltzmann constant, respectively. The theory of orientational mobility in dendrimer [68,69,70,71,72,73,74,75,76,77] predicts that the main contribution to NMR relaxation is provided by two processes with different correlation times: (i) rotation dendrimer as a whole and branch (or subbranch) reorientation.
The local orientational mobility of groups in the dendrimer is determined by the position of the 1/T1H maximum. Moreover, the more to the left of the maximum (i.e., shifted towards high temperatures), the slower the group mobility. In dendrimer macromolecules, the inner groups are the slowest (Figure 8a). The most mobile are the terminal groups, for which an exponential growth is observed in the 1/T1H dependence, and the maximum is not reached due to the limited experimental temperature range (the freezing of the solvent) (Figure 8c). The mobility of side groups depends on the structure of the side segment and the position of the observed group in it. For example, in the case of the Lys-2Lys dendrimer, the mobility of the side groups is the same as the mobility of the terminal groups [51]. However, in Lys-2His, the mobility of the side groups coincides with the mobility of the inner groups [48].
As shown in [48], the position of the maximum of the 1/T1H temperature dependence for inner groups shifts to the left (towards high temperatures) due to the transition from the swollen conformation to the collapsed one. Thus, if the structures of the dendrimers are similar, then the criterion for contraction of the dendrimer can be the shift of the maximum of the 1/T1H dependence, to which the inner groups contribute, towards high temperatures. Figure 8a,b show the 1/T1H dependences for the inner groups of Lys-2Arg (swollen conformation) and Lys-2His (collapsed conformation), which were obtained earlier and used as references.
Let us now proceed to analyze the relaxation data for the Lys-Arg-His and Lys-His-Arg dendrimers. As can be seen from Figure 8a, the 1/T1H dependence for the side and inner groups of the Lys-Arg-His dendrimer practically coincides with the 1/T1H curve for the inner groups of the Lys-2Arg dendrimer. This means that the mobility of the inner groups of Lys-Arg-His indicates the swollen conformation of the dendrimer. However, it should be noted that in the Lys-Arg-His dendrimer, the side arginine groups that contribute to this dependence have a higher mobility than the inner groups and side groups in histidine residues. For illustration, in Figure 8a, the 1/T1H curve for the side and inner groups of the Lys-2Arg dendrimer is shown (solid red squares), the maximum for which is observed much more to the right than for the other groups. It can be expected that, in the case of Lys-Arg-His, the side arginine groups will also have a similar mobility, since its contribution is 40%. If it were possible to separate the signal only from the inner groups of the Lys-Arg-His dendrimer, then, obviously, the position of the 1/T1H maximum would be to the left of the same maximum for the inner groups in Lys-2Arg. Thus, it can be concluded that the mobility of the inner groups in Lys-Arg-His is slower than in the Lys-2Arg dendrimer.
In the case of the Lys-His-Arg dendrimer, the 1/T1H maximum is shifted by 10 K to the region of high temperatures compared to the similar curve for Lys-Arg-His. Therefore, taking into account the contribution of the side arginine groups, we can expect that the shift of the maximum will be even more significant and close to the 1/T1H dependence for the Lys-2His dendrimer (in the collapsed conformation). Such a slowdown in the mobility of the inner groups indicates that the conformation of the Lys-His-Arg dendrimer is close to the collapsed one.
Analogous conclusions can be drawn from the 1/T1H temperature dependence for the terminal groups (Figure 8c). The mobility of the terminal groups of the Lys-Arg-His dendrimer is similar to that of the Lys-2Arg dendrimer. At the same time, the 1/T1H dependencies for the terminal groups of Lys-Arg-His and Lys-2His have a similar behavior.
Thus, according to the NMR relaxation data, it can be argued that, despite the same chemical composition, the Lys-Arg-His and Lys-His-Arg dendrimers have different conformational structures. The global conformation of the Lys-Arg-His dendrimer is close to collapsed, and the conformation of the Lys-His-Arg dendrimer is more swollen.

3. Conclusions

In our recent work [48], it was found that a peptide dendrimer with double histidine insertions changes its global conformation, from swollen at low pH (Lys-2Hisp) to collapsed at normal pH (Lys-2His). However, the biomedical usage of the Lys-2His dendrimer in a collapsed conformation can be problematic due to its tendency to aggregate.
This work is a continuation of our previous study in [48]. Here, we present the synthesis of new peptide dendrimers with the same amino acid composition, but different amino acid sequences (Arg-His or His-Arg), in insertions between lysine branching points. These dendrimers were experimentally investigated using NMR spectroscopy and NMR relaxation methods. The main idea for this research was to obtain a dendrimer macromolecule that retains the properties of size change due to the recharging of imidazole rings, but prevents the problem of aggregation of dendrimers with each other due to the insertion of charged guanidine groups in arginine residues.
Despite the same amino acid composition, the Lys-Arg-His and Lys-His-Arg dendrimers have different structural properties. In the case of the Lys-Arg-His dendrimer, the pairing effect of imidazole rings of the histidine residues does not appear in the NMR spectra, and the NMR relaxation behavior indicates a swollen conformation of the macromolecule, similar to that of the Lys-2Arg dendrimer. The opposite situation is observed for the Lys-His-Arg dendrimer: the shape and position of the peaks of the side histidine groups are similar to the corresponding peaks of the Lys-2His dendrimer, in which the paring effect is observed. Moreover, according to the NMR relaxation data, the conformation of the Lys-His-Arg dendrimer becomes close to the collapsed, as in the case of the Lys-2His dendrimer.
Thus, we can make a conclusion that the Lys-His-Arg dendrimer is the most suitable candidate for use as a pH sensitive nanocontainer in biomedical applications.

4. Materials and Methods

4.1. Synthesis of Lys-His-Arg and Lys-Arg-His

Boc-amino acids were purchased from Bachem Holding (Torrance, CA, USA) and Iris Biotech GMBH (Marktredwitz, Germany); p-мethylbenzhydrylamine resin (MBHA-resin) was supplied by Bachem Holding (Torrance, CA, USA); trifluoromethanesulfonic acid (TFMSA), diisopropylcarbodiimide (DIC), 1-hydroxybenzotriazole (HOBt), thioanisole, ethanedithiol, and other reagents were purchased from Sigma-Aldrich (Munich, Germany). Triethylamine, dichloromethane (DCM) and dimethylformamide (DMF) were purchased from Vecton Ltd. (St. Petersburg, Russia). Trifluoroacetic acid (TFA) was purchased from Panreac (Barcelona, Spain) and distilled before application. All solvents were purified and distilled using standard procedures.
Lysine-based dendrimers with dipeptide insertions His-Arg and Arg-His between the branching points were obtained by a step-by-step BOC solid-phase peptide synthesis method (SPPS) [51]. The synthesis was carried out manually using 0.1 g MBHA-resin (amino group content was 0.72 mmol/g). The synthesis of the dendrimer molecules includes: (1) the introduction of L-alanine to initiate the synthesis; (2) branching using the trifunctional amino acid Boc-Lys(Boc); and (3) the formation of inner layers by insertion of histidine (His) and arginine (Arg) amino acid residues.
However, several points should be noted regarding the procedure for the synthesis of these dendrimers in order to obtain high quality products. Since the number of amino acids increased exponentially with each addition of Boc-Lys(Boc), the reaction was carefully controlled by test to avoid errors in the synthesis. During the formation of the first generation layers of macromolecules, we blocked unreacted peptide chains with an acetic anhydride/dichloromethane (1:1 v/v) solution for 30 min before washing the MBHA-resin with dimethylformamide (DMF), methanol, and dichloromethane (DCM) (twice each). When binding was difficult, especially on the terminal layers, N-methylpyrrolidone (NMP) was added to DMF. Probably some difficulties were caused by the presence of massive protective side groups: a p-benzyloxymethyl group (Bom) on histidine and a mesitylene-2-sulfonyl group (Mts) on arginine.
In the case of the Lys-His-Arg dendrimer, the average time of one acylation was increased to 10–12 h instead of 4–6 h, especially with increasing generations. In addition, the introduction of histidine residues in the first layer led to certain difficulties in the synthesis, compared with the Lys-Arg-His dendrimer, in which the arginine layer was the starting one. The first layer of histidine not only increased the binding time and the excess of reagents up to 6 equivalents, but also caused problems with the subsequent insertion of arginine residues. Furthermore, for complete conversion during the formation of the second generation layer of the Lys-His-Arg dendrimer, in addition to increasing the time and quantity of reagents, it was necessary to add 4-N,N-dimethylaminopyridine (DMAP) (0.1 eq.) as a catalyst.
In the case of the Lys-Arg-His dendrimer, practically no reacylation of the branching lysine layer was required, except for the terminal one; the other stages took twice as much time and twice as many reagents. Despite this, the amino groups in each generation remained available for complete conversion.

4.2. NMR Experiments

For the NMR study, the samples of the Lys-Arg-His and Lys-His-Arg dendrimers were dissolved in 0.158 M NaCl D2O at a concentration of about 1.54 g/dl and 1.59 g/dl, respectively.
NMR measurements were performed on a Bruker Avance III 500 MHz spectrometer (Karlsruhe, Germany) equipped with a standard 5 mm BBFO direct observation probe and a Great 1/60A gradient system with a MIC S2 Diff/30 diffusive probe with a 1H convolution compensation coil (EVT). One- and two-dimensional NMR spectra for peak assignments were recorded at 298 K. All spectra were obtained using standard pulse sequences. The 1H spin-lattice relaxation times, T1H, were acquired with an “inversion-recovery” pulse sequence, modified by the destructive gradient pulses at the beginning of the sequence (“spoiler-recovery” sequence) [78]. NMR relaxation experiments were carried out at the temperature range 283–343 K.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24020949/s1.

Author Contributions

Conceptualization, I.M.N.; Methodology, I.I.T.; Validation, N.N.S., M.A.V. and M.E.M.; Formal analysis, N.N.S. and M.A.V.; Investigation, N.N.S., I.I.T., M.A.V. and D.A.M.; Resources, I.I.T. and M.E.M.; Writing–original draft, N.N.S. and D.A.M.; Writing–review & editing, all authors; Visualization, N.N.S.; Supervision, I.M.N. and D.A.M. All authors have read and agreed to the published version of the manuscript.

Funding

Author I.M.N. was supported by RFBR grant 20-53-12036.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The NMR measurements were carried out in the Center for Magnetic Resonance of Research Park of St. Petersburg State University. Irina I. Tarasenko worked accordingly to State Assignment of the IMC RAS. Igor M. Neelov is used the equipment of the shared research facilities of HPC computing resources at Lomonosov Moscow State University [79].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chis, A.A.; Dobrea, C.; Morgovan, C.; Arseniu, A.M.; Rus, L.L.; Butuca, A.; Juncan, A.M.; Totan, M.; Vonica-Tincu, A.L.; Cormos, G.; et al. Applications and Limitations of Dendrimers in Biomedicine. Molecules 2020, 25, 3982. [Google Scholar] [CrossRef] [PubMed]
  2. Dias, A.P.; da Silva Santos, S.; da Silva, J.V.; Parise-Filho, R.; Igne Ferreira, E.; El Seoud, O.; Giarolla, J. Dendrimers in the context of nanomedicine. Int. J. Pharm. 2020, 573, 118814. [Google Scholar] [CrossRef] [PubMed]
  3. Lee, C.C.; MacKay, J.A.; Frechet, J.M.J.; Szoka, F.C. Designing dendrimers for biological applications. Nat. Biotech. 2005, 23, 1517–1526. [Google Scholar] [CrossRef] [PubMed]
  4. Majoros, I.; Baker, J.J.R. Dendrimer-Based Nanomedicine; Majoros, I.J., Baker, J.R., Eds.; Pan Stanford Publishing: Singapore, 2008; ISBN 9789814241045. [Google Scholar]
  5. Svenson, S.; Tomalia, D.A. Dendrimers in biomedical applications—Reflections on the field. Adv. Drug Deliv. Rev. 2012, 64, 102–115. [Google Scholar] [CrossRef]
  6. Jang, W.-D.; Kamruzzaman Selim, K.M.; Lee, C.-H.; Kang, I.-K. Bioinspired application of dendrimers: From bio-mimicry to biomedical applications. Prog. Polym. Sci. 2009, 34, 1–23. [Google Scholar] [CrossRef]
  7. Liu, M.; Fréchet, J.M.J. Designing dendrimers for drug delivery. Pharm. Sci. Technolo. Today 1999, 2, 393–401. [Google Scholar]
  8. Esfand, R.; Tomalia, D.A. Poly(amidoamine) (PAMAM) dendrimers: From biomimicry to drug delivery and biomedical applications. Drug Discov. Today 2001, 6, 427–436. [Google Scholar] [CrossRef]
  9. Choi, Y.; Thomas, T.; Kotlyar, A.; Islam, M.T.; Baker, J.R. Synthesis and Functional Evaluation of DNA-Assembled Polyamidoamine Dendrimer Clusters for Cancer Cell-Specific Targeting. Chem. Biol. 2005, 12, 35–43. [Google Scholar] [CrossRef] [Green Version]
  10. Tomalia, D.A.; Reyna, L.A.; Svenson, S. Dendrimers as multi-purpose nanodevices for oncology drug delivery and diagnostic imaging. Biochem. Soc. Trans. 2007, 35, 61–67. [Google Scholar] [CrossRef] [Green Version]
  11. Gillies, E.R.; Fréchet, J.M.J. pH-Responsive Copolymer Assemblies for Controlled Release of Doxorubicin. Bioconjug. Chem. 2005, 16, 361–368. [Google Scholar] [CrossRef]
  12. Boas, U.; Christensen, J.B.; Heegaard, P.M.H.; Peng, L. Dendrimers in Medicine and Biotechnology: New Molecular Tools; The Royal Society of Chemistry: London, UK, 2006; ISBN 978-0-85404-852-6. [Google Scholar]
  13. Tekade, R.K.; Kumar, P.V.; Jain, N.K. Dendrimers in Oncology: An Expanding Horizon. Chem. Rev. 2009, 109, 49–87. [Google Scholar] [CrossRef] [PubMed]
  14. Singh, S.K.; Sharma, V.K. Dendrimers: A Class of Polymer in the Nanotechnology for Drug Delivery. In Nanomedicine for Drug Delivery and Therapeutics; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2013; pp. 373–409. ISBN 9781118414095. [Google Scholar]
  15. Svenson, S.; Chauhan, A.S. Dendrimers for enhanced drug solubilization. Nanomedicine 2008, 3, 679–702. [Google Scholar] [CrossRef] [PubMed]
  16. Duncan, R.; Izzo, L. Dendrimer biocompatibility and toxicity. Adv. Drug Deliv. Rev. 2005, 57, 2215–2237. [Google Scholar] [CrossRef] [PubMed]
  17. Khan, M.K.; Nigavekar, S.S.; Minc, L.D.; Kariapper, M.S.T.; Nair, B.M.; Lesniak, W.G.; Balogh, L.P. In Vivo Biodistribution of Dendrimers and Dendrimer Nanocomposites—Implications for Cancer Imaging and Therapy. Technol. Cancer Res. Treat. 2005, 4, 603–613. [Google Scholar] [CrossRef]
  18. Sadler, K.; Tam, J.P. Peptide dendrimers: Applications and synthesis. Rev. Mol. Biotechnol. 2002, 90, 195–229. [Google Scholar] [CrossRef]
  19. Martinho, N.; Silva, L.C.; Florindo, H.F.; Brocchini, S.; Zloh, M.; Barata, T.S. Rational design of novel, fluorescent, tagged glutamic acid dendrimers with different terminal groups and in silico analysis of their properties. Int. J. Nanomedicine 2017, 12, 7053–7073. [Google Scholar] [CrossRef] [Green Version]
  20. Hsu, H.; Bugno, J.; Lee, S.; Hong, S. Dendrimer-based nanocarriers: A versatile platform for drug delivery. WIREs Nanomed. Nanobiotechnol. 2017, 9, e1409. [Google Scholar] [CrossRef]
  21. Palmerston Mendes, L.; Pan, J.; Torchilin, V. Dendrimers as Nanocarriers for Nucleic Acid and Drug Delivery in Cancer Therapy. Molecules 2017, 22, 1401. [Google Scholar] [CrossRef] [Green Version]
  22. Gorzkiewicz, M.; Konopka, M.; Janaszewska, A.; Tarasenko, I.I.; Sheveleva, N.N.; Gajek, A.; Neelov, I.M.; Klajnert-Maculewicz, B. Application of new lysine-based peptide dendrimers D3K2 and D3G2 for gene delivery: Specific cytotoxicity to cancer cells and transfection in vitro. Bioorg. Chem. 2020, 95, 103504. [Google Scholar] [CrossRef]
  23. Kuang, T.; Fu, D.; Chang, L.; Yang, Z.; Chen, Z.; Jin, L.; Chen, F.; Peng, X. Recent Progress in Dendrimer-based Gene Delivery Systems. Curr. Org. Chem. 2016, 20, 1820–1826. [Google Scholar] [CrossRef] [Green Version]
  24. Pompilio, A.; Geminiani, C.; Mantini, P.; Siriwardena, T.N.; Di Bonaventura, I.; Reymond, J.L.; Di Bonaventura, G. Peptide dendrimers as “lead compounds” for the treatment of chronic lung infections by Pseudomonas aeruginosa in cystic fibrosis patients: In vitro and in vivo studies. Infect. Drug Resist. 2018, 11, 1767–1782. [Google Scholar] [CrossRef] [PubMed]
  25. Ben Jeddou, F.; Falconnet, L.; Luscher, A.; Siriwardena, T.; Reymond, J.-L.; van Delden, C.; Köhler, T. Adaptive and Mutational Responses to Peptide Dendrimer Antimicrobials in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2020, 64, e02040-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Kawano, Y.; Jordan, O.; Hanawa, T.; Borchard, G.; Patrulea, V. Are Antimicrobial Peptide Dendrimers an Escape from ESKAPE? Adv. Wound Care 2020, 9, 378–395. [Google Scholar] [CrossRef] [PubMed]
  27. Cañas-Arranz, R.; de León, P.; Forner, M.; Defaus, S.; Bustos, M.J.; Torres, E.; Andreu, D.; Blanco, E.; Sobrino, F. Immunogenicity of a Dendrimer B2T Peptide Harboring a T-Cell Epitope from FMDV Non-structural Protein 3D. Front. Vet. Sci. 2020, 7, 498. [Google Scholar] [CrossRef]
  28. Shao, N.; Wang, F.; Chen, Y.; Wang, H.; Cheng, Y.; Wang, Y. Synergistic effect of amino acids modified on dendrimer surface in gene delivery. Biomaterials 2014, 35, 9187–9198. [Google Scholar]
  29. Crespo, L.; Sanclimens, G.; Pons, M.; Giralt, E.; Royo, M.; Albericio, F. Peptide and amide bond-containing dendrimers. Chem. Rev. 2005, 105, 1663–1681. [Google Scholar] [CrossRef]
  30. Johansson, E.M.V.; Crusz, S.A.; Kolomiets, E.; Buts, L.; Kadam, R.U.; Cacciarini, M.; Bartels, K.-M.; Diggle, S.P.; Cámara, M.; Williams, P.; et al. Inhibition and Dispersion of Pseudomonas aeruginosa Biofilms by Glycopeptide Dendrimers Targeting the Fucose-Specific Lectin LecB. Chem. Biol. 2008, 15, 1249–1257. [Google Scholar] [CrossRef]
  31. Gorain, B.; Choudhury, H.; Pandey, M.; Mohd Amin, M.C.I.; Singh, B.; Gupta, U.; Kesharwani, P. Dendrimers as Effective Carriers for the Treatment of Brain Tumor. In Nanotechnology-Based Targeted Drug Delivery Systems for Brain Tumors; Elsevier: Amsterdam, The Netherlands, 2018; pp. 267–305. [Google Scholar]
  32. Cooper, B.M.; Iegre, J.; O’ Donovan, D.H.; Ölwegård Halvarsson, M.; Spring, D.R. Peptides as a platform for targeted therapeutics for cancer: Peptide–drug conjugates (PDCs). Chem. Soc. Rev. 2021, 50, 1480–1494. [Google Scholar] [CrossRef]
  33. Maysinger, D.; Zhang, Q.; Kakkar, A. Dendrimers as Modulators of Brain Cells. Molecules 2020, 25, 4489. [Google Scholar] [CrossRef]
  34. Filipczak, N.; Yalamarty, S.S.K.; Li, X.; Parveen, F.; Torchilin, V. Developments in Treatment Methodologies Using Dendrimers for Infectious Diseases. Molecules 2021, 26, 3304. [Google Scholar] [CrossRef]
  35. Gorzkiewicz, M.; Kopeć, O.; Janaszewska, A.; Konopka, M.; Pędziwiatr-Werbicka, E.; Tarasenko, I.I.; Bezrodnyi, V.V.; Neelov, I.M.; Klajnert-Maculewicz, B. Poly(lysine) Dendrimers Form Complexes with siRNA and Provide Its Efficient Uptake by Myeloid Cells: Model Studies for Therapeutic Nucleic Acid Delivery. Int. J. Mol. Sci. 2020, 21, 3138. [Google Scholar] [CrossRef] [PubMed]
  36. Okuda, T.; Sugiyama, A.; Niidome, T.; Aoyagi, H. Characters of dendritic poly(L-lysine) analogues with the terminal lysines replaced with arginines and histidines as gene carriers in vitro. Biomaterials 2004, 25, 537–544. [Google Scholar] [CrossRef] [PubMed]
  37. Boas, U.; Heegaard, P.M.H. Dendrimers in drug research. Chem. Soc. Rev. 2004, 33, 43–63. [Google Scholar] [CrossRef] [PubMed]
  38. Stiriba, S.-E.; Frey, H.; Haag, R. Dendritic Polymers in Biomedical Applications: From Potential to Clinical Use in Diagnostics and Therapy. Angew. Chemie Int. Ed. 2002, 41, 1329–1334. [Google Scholar] [CrossRef]
  39. Agarwal, A.; Saraf, S.; Asthana, A.; Gupta, U.; Gajbhiye, V.; Jain, N.K. Ligand based dendritic systems for tumor targeting. Int. J. Pharm. 2008, 350, 3–13. [Google Scholar] [CrossRef]
  40. Yellepeddi, V.K.; Kumar, A.; Palakurthi, S. Surface modified poly(amido)amine dendrimers as diverse nanomolecules for biomedical applications. Expert Opin. Drug Deliv. 2009, 6, 835–850. [Google Scholar] [CrossRef]
  41. Kharwade, R.; More, S.; Warokar, A.; Agrawal, P.; Mahajan, N. Starburst pamam dendrimers: Synthetic approaches, surface modifications, and biomedical applications. Arab. J. Chem. 2020, 13, 6009–6039. [Google Scholar] [CrossRef]
  42. Wolinsky, J.; Grinstaff, M. Therapeutic and diagnostic applications of dendrimers for cancer treatment. Adv. Drug Deliv. Rev. 2008, 60, 1037–1055. [Google Scholar] [CrossRef]
  43. Chen, Y.; Tao, K.; Ji, W.; Kumar, V.B.; Rencus-Lazar, S.; Gazit, E. Histidine as a key modulator of molecular self-assembly: Peptide-based supramolecular materials inspired by biological systems. Mater. Today 2022, 60, 106–127. [Google Scholar] [CrossRef]
  44. Bradbury, J.H.; Chapman, B.E.; Crompton, M.W.; Norton, R.S.; Teh, J.S. Hydrogen–deuterium exchange of the C-2 protons of histidine and histidine peptides and proteins. J. Chem. Soc. Perkin Trans. 2 1980, 693–700. [Google Scholar] [CrossRef]
  45. Cebo, M.; Kielmas, M.; Adamczyk, J.; Cebrat, M.; Szewczuk, Z.; Stefanowicz, P. Hydrogen-deuterium exchange in imidazole as a tool for studying histidine phosphorylation. Anal. Bioanal. Chem. 2014, 406, 8013–8020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Matthews, H.R.; Matthews, K.S.; Opella, S.J. Selectively deuterated amino acid analogues synthesis, Incorporation into proteins and NMR properties. BBA-Gen. Subj. 1977, 497, 1–13. [Google Scholar] [CrossRef]
  47. Sheveleva, N.N.; Markelov, D.A.; Vovk, M.A.; Tarasenko, I.I.; Mikhailova, M.E.; Ilyash, M.Y.; Neelov, I.M.; Lahderanta, E. Stable Deuterium Labeling of Histidine-Rich Lysine-Based Dendrimers. Molecules 2019, 24, 2481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Sheveleva, N.N.; Bezrodnyi, V.V.; Mikhtaniuk, S.E.; Shavykin, O.V.; Neelov, I.M.; Tarasenko, I.I.; Vovk, M.A.; Mikhailova, M.E.; Penkova, A.V.; Markelov, D.A. Local Orientational Mobility of Collapsed Dendrimers. Macromolecules 2021, 54, 11083–11092. [Google Scholar] [CrossRef]
  49. Shavykin, O.V.; Mikhailov, I.V.; Darinskii, A.A.; Neelov, I.M.; Leermakers, F.A.M. Effect of an asymmetry of branching on structural characteristics of dendrimers revealed by Brownian dynamics simulations. Polymer 2018, 146, 256–266. [Google Scholar] [CrossRef]
  50. Shavykin, O.V.; Leermakers, F.A.M.; Neelov, I.M.; Darinskii, A.A. Self-Assembly of Lysine-Based Dendritic Surfactants Modeled by the Self-Consistent Field Approach. Langmuir 2018, 34, 1613–1626. [Google Scholar] [CrossRef]
  51. Sheveleva, N.N.; Markelov, D.A.; Vovk, M.A.; Mikhailova, M.E.; Tarasenko, I.I.; Neelov, I.M.; Lähderanta, E. NMR studies of excluded volume interactions in peptide dendrimers. Sci. Rep. 2018, 8, 8916. [Google Scholar] [CrossRef] [Green Version]
  52. Sheveleva, N.N.; Markelov, D.A.; Vovk, M.A.; Mikhailova, M.E.; Tarasenko, I.I.; Tolstoy, P.M.; Neelov, I.M.; Lähderanta, E. Lysine-based dendrimer with double arginine residues. RSC Adv. 2019, 9, 18018–18026. [Google Scholar] [CrossRef] [Green Version]
  53. Neelov, I.M.; Falkovich, S.; Markelov, D.; Paci, E.; Darinskii, A.; Tenhu, H. Molecular Dynamics of Lysine Dendrimers. Computer Simulation and NMR. In Dendrimers in Biomedical Applications; Royal Society of Chemistry: Cambridge, UK, 2013; pp. 99–114. ISBN 978-1-84973-611-4. [Google Scholar]
  54. Neelov, I.M.; Markelov, D.A.; Falkovich, S.G.; Ilyash, M.Y.; Okrugin, B.M.; Darinskii, A.A. Mathematical simulation of lysine dendrimers: Temperature dependences. Polym. Sci. Ser. C 2013, 55, 154–161. [Google Scholar] [CrossRef]
  55. Okrugin, B.; Ilyash, M.; Markelov, D.; Neelov, I. Lysine Dendrigraft Nanocontainers. Influence of Topology on Their Size and Internal Structure. Pharmaceutics 2018, 10, 129. [Google Scholar] [CrossRef] [Green Version]
  56. Mikhtaniuk, S.; Bezrodnyi, V.; Shavykin, O.; Neelov, I.; Sheveleva, N.; Penkova, A.; Markelov, D. Comparison of Structure and Local Dynamics of Two Peptide Dendrimers with the Same Backbone but with Different Side Groups in Their Spacers. Polymers 2020, 12, 1657. [Google Scholar] [CrossRef] [PubMed]
  57. Shavykin, O.V.; Neelov, I.M.; Darinskii, A.A. Is the manifestation of the local dynamics in the spin–lattice NMR relaxation in dendrimers sensitive to excluded volume interactions? Phys. Chem. Chem. Phys. 2016, 18, 24307–24317. [Google Scholar] [CrossRef] [PubMed]
  58. Okrugin, B.M.; Neelov, I.M.; Leermakers, F.A.M.; Borisov, O.V. Structure of asymmetrical peptide dendrimers: Insights given by self-consistent field theory. Polymer 2017, 125, 292–302. [Google Scholar] [CrossRef]
  59. Heyda, J.; Mason, P.E.; Jungwirth, P. Attractive Interactions between Side Chains of Histidine-Histidine and Histidine-Arginine-Based Cationic Dipeptides in Water. J. Phys. Chem. B 2010, 114, 8744–8749. [Google Scholar] [CrossRef] [PubMed]
  60. Chen, Y.; Yang, Y.; Orr, A.A.; Makam, P.; Redko, B.; Haimov, E.; Wang, Y.; Shimon, L.J.W.; Rencus-Lazar, S.; Ju, M.; et al. Self-Assembled Peptide Nano-Superstructure towards Enzyme Mimicking Hydrolysis. Angew. Chemie Int. Ed. 2021, 60, 17164–17170. [Google Scholar] [CrossRef]
  61. Bezrodnyi, V.V.; Mikhtaniuk, S.E.; Shavykin, O.V.; Neelov, I.M.; Sheveleva, N.N.; Markelov, D.A. Size and Structure of Empty and Filled Nanocontainer Based on Peptide Dendrimer with Histidine Spacers at Different pH. Molecules 2021, 26, 6552. [Google Scholar] [CrossRef]
  62. Bezrodnyi, V.V.; Shavykin, O.V.; Mikhtaniuk, S.E.; Neelov, I.M.; Sheveleva, N.N.; Markelov, D.A. Why the Orientational Mobility in Arginine and Lysine Spacers of Peptide Dendrimers Designed for Gene Delivery Is Different? Int. J. Mol. Sci. 2020, 21, 9749. [Google Scholar] [CrossRef]
  63. Abragam, A. The Principles of Nuclear Magnetism; Oxford University Press: Oxford, UK, 1983; ISBN 9780198520146. [Google Scholar]
  64. Bloembergen, N.; Purcell, E.M.; Pound, R.V. Relaxation effects in nuclear magnetic resonance absorption. Phys. Rev. 1948, 73, 679–712. [Google Scholar] [CrossRef] [Green Version]
  65. Solomon, I. Relaxation Processes in a System of Two Spins. Phys. Rev. 1955, 99, 559–565. [Google Scholar] [CrossRef]
  66. Chizhik, V.I.; Chernyshev, Y.S.; Donets, A.V.; Frolov, V.V.; Komolkin, A.V.; Shelyapina, M.G. Magnetic Resonance and Its Applications; Springer International Publishing: Berlin/Heidelberg, Germany, 2014; ISBN 978-3-319-05298-4. [Google Scholar]
  67. Levitt, M.H. Spin Dynamics: Basics of Nuclear Magnetic Resonance, 2nd ed.; Wiley: Chichester, UK; Hoboken, NJ, USA, 2008; ISBN 978-0-470-51118-3. [Google Scholar]
  68. Gotlib, Y.Y.; Markelov, D.A. Permittivity of a dendrimer containing polar groups. Polym. Sci. Ser. A 2004, 46, 815–832. [Google Scholar]
  69. Gotlib, Y.Y.; Markelov, D.A. Theory of orientational relaxation of individual specified units in a dendrimer. Polym. Sci. Ser. A 2007, 49, 1137–1154. [Google Scholar] [CrossRef]
  70. Markelov, D.A.; Dolgushev, M.; Gotlib, Y.Y.; Blumen, A. NMR relaxation of the orientation of single segments in semiflexible dendrimers. J. Chem. Phys. 2014, 140, 244904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Novoa-Carballal, R.; Säwén, E.; Fernandez-Megia, E.; Correa, J.; Riguera, R.; Widmalm, G. The dynamics of GATG glycodendrimers by NMR diffusion and quantitative 13C relaxation. Phys. Chem. Chem. Phys. 2010, 12, 6587. [Google Scholar] [CrossRef] [PubMed]
  72. Markelov, D.A.; Matveev, V.V.; Ingman, P.; Nikolaeva, M.N.; Lähderanta, E.; Shevelev, V.A.; Boiko, N.I. NMR Studies of Carbosilane Dendrimer with Terminal Mesogenic Groups. J. Phys. Chem. B 2010, 114, 4159–4165. [Google Scholar] [CrossRef]
  73. Pinto, L.F.; Correa, J.; Martin-Pastor, M.; Riguera, R.; Fernandez-Megia, E. The dynamics of dendrimers by NMR relaxation: Interpretation pitfalls. J. Am. Chem. Soc. 2013, 135, 1972–1977. [Google Scholar] [CrossRef]
  74. Pinto, L.F.; Riguera, R.; Fernandez-Megia, E. Stepwise filtering of the internal layers of dendrimers by transverse-relaxation-edited NMR. J. Am. Chem. Soc. 2013, 135, 11513–11516. [Google Scholar] [CrossRef]
  75. Markelov, D.A.; Dolgushev, M.; Lähderanta, E. Chapter One—NMR Relaxation in Dendrimers. In Annual Reports on NMR Spectroscopy; G.A. Webb, Ed.; Academic Press: Cambridge, MA, USA, 2017; Volume 91, pp. 1–66. ISBN 0066-4103. [Google Scholar]
  76. Markelov, D.A.; Shishkin, A.N.; Matveev, V.V.; Penkova, A.V.; Lähderanta, E.; Chizhik, V.I. Orientational Mobility in Dendrimer Melts: Molecular Dynamics Simulations. Macromolecules 2016, 49, 9247–9257. [Google Scholar] [CrossRef]
  77. Gupta, S.; Biswas, P. Orientational Relaxation of Poly(propylene imine) Dendrimers at Different pH. J. Phys. Chem. B 2020, 124, 4193–4202. [Google Scholar] [CrossRef]
  78. Sørland, G.H. The Spoiler Recovery Approach (SR). In Dynamic Pulsed-Field-Gradient NMR; Springer Series in Chemical Physics; Springer: Berlin/Heidelberg, Germany, 2014; Volume 110, pp. 169–190. ISBN1 978-3-662-44500-6. ISBN2 978-3-662-44499-3. [Google Scholar]
  79. Sadovnichy, V.; Tikhonravov, A.; Voevodin, V.l.; Opanasenko, V. “Lomonosov”: Supercomputing at Moscow State University. In Contemporary High Performance Computing: From Petascale toward Exascale; Vetter, J.S., Ed.; Chapman and Hall/CRC Press: London, UK, 2013; pp. 283–307. ISBN 9781466568341. [Google Scholar]
Figure 1. 1H NMR spectrum of the Lys-Arg-His dendrimer at 298 K. The dendrimer contains three types of methylene groups connected with nitrogen atoms (or imidazole rings in His residues), which have been used in NMR relaxation study: inner groups (green open circle), side groups (blue circles) and terminal groups (red circle). The peaks from small molecular weight impurities are marked by asterix (*).
Figure 1. 1H NMR spectrum of the Lys-Arg-His dendrimer at 298 K. The dendrimer contains three types of methylene groups connected with nitrogen atoms (or imidazole rings in His residues), which have been used in NMR relaxation study: inner groups (green open circle), side groups (blue circles) and terminal groups (red circle). The peaks from small molecular weight impurities are marked by asterix (*).
Ijms 24 00949 g001
Figure 2. 13C NMR spectrum of the Lys-Arg-His dendrimer at 298 K. The peaks from small molecular weight impurities are marked by asterix (*).
Figure 2. 13C NMR spectrum of the Lys-Arg-His dendrimer at 298 K. The peaks from small molecular weight impurities are marked by asterix (*).
Ijms 24 00949 g002
Figure 3. 1H NMR spectrum of the Lys-His-Arg dendrimer at 298 K. Three types of methylene groups connected with nitrogen atoms (or imidazole rings in His residues), which have been used in NMR relaxation study: inner groups (green open circle), side groups (blue circles) and terminal groups (red circle). The peaks from small molecular weight impurities are marked by asterix (*).
Figure 3. 1H NMR spectrum of the Lys-His-Arg dendrimer at 298 K. Three types of methylene groups connected with nitrogen atoms (or imidazole rings in His residues), which have been used in NMR relaxation study: inner groups (green open circle), side groups (blue circles) and terminal groups (red circle). The peaks from small molecular weight impurities are marked by asterix (*).
Ijms 24 00949 g003
Figure 4. 13C NMR spectrum of the Lys-His-Arg dendrimer at 298 K. The peaks from small molecular weight impurities are marked by asterix (*).
Figure 4. 13C NMR spectrum of the Lys-His-Arg dendrimer at 298 K. The peaks from small molecular weight impurities are marked by asterix (*).
Ijms 24 00949 g004
Figure 5. 1H-13C HSQC NMR spectrum of the Lys-His-Arg dendrimer in the range of 3.40–2.65 ppm at 298 K.
Figure 5. 1H-13C HSQC NMR spectrum of the Lys-His-Arg dendrimer in the range of 3.40–2.65 ppm at 298 K.
Ijms 24 00949 g005
Figure 6. The schematic illustration of a pairing between imidazole rings. Carbon is grey; nitrogen is blue; hydrogen is white.
Figure 6. The schematic illustration of a pairing between imidazole rings. Carbon is grey; nitrogen is blue; hydrogen is white.
Ijms 24 00949 g006
Figure 7. Comparison of 1H NMR spectra of different peptide dendrimers in the region of signals from inner CH2-(N) (green), side and inner (blue) and terminal (red) groups at a temperature of 298 K.
Figure 7. Comparison of 1H NMR spectra of different peptide dendrimers in the region of signals from inner CH2-(N) (green), side and inner (blue) and terminal (red) groups at a temperature of 298 K.
Ijms 24 00949 g007
Figure 8. Temperature dependences of the spin−lattice relaxation rate, 1/T1H: of (a) side and inner CH2 groups of the Lys-Arg-His and Lys-2Arg dendrimers and inner groups of Lys-2Arg; (b) side and inner CH2 groups of the Lys-His-Arg and Lys-2His dendrimers; (c) terminal groups of Lys-Arg-His, Lys-His-Arg, Lys-2Arg and Lys-2His dendrimers.
Figure 8. Temperature dependences of the spin−lattice relaxation rate, 1/T1H: of (a) side and inner CH2 groups of the Lys-Arg-His and Lys-2Arg dendrimers and inner groups of Lys-2Arg; (b) side and inner CH2 groups of the Lys-His-Arg and Lys-2His dendrimers; (c) terminal groups of Lys-Arg-His, Lys-His-Arg, Lys-2Arg and Lys-2His dendrimers.
Ijms 24 00949 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sheveleva, N.N.; Tarasenko, I.I.; Vovk, M.A.; Mikhailova, M.E.; Neelov, I.M.; Markelov, D.A. NMR Studies of Two Lysine Based Dendrimers with Insertion of Similar Histidine-Arginine and Arginine-Histidine Spacers Having Different Properties for Application in Drug Delivery. Int. J. Mol. Sci. 2023, 24, 949. https://doi.org/10.3390/ijms24020949

AMA Style

Sheveleva NN, Tarasenko II, Vovk MA, Mikhailova ME, Neelov IM, Markelov DA. NMR Studies of Two Lysine Based Dendrimers with Insertion of Similar Histidine-Arginine and Arginine-Histidine Spacers Having Different Properties for Application in Drug Delivery. International Journal of Molecular Sciences. 2023; 24(2):949. https://doi.org/10.3390/ijms24020949

Chicago/Turabian Style

Sheveleva, Nadezhda N., Irina I. Tarasenko, Mikhail A. Vovk, Mariya E. Mikhailova, Igor M. Neelov, and Denis A. Markelov. 2023. "NMR Studies of Two Lysine Based Dendrimers with Insertion of Similar Histidine-Arginine and Arginine-Histidine Spacers Having Different Properties for Application in Drug Delivery" International Journal of Molecular Sciences 24, no. 2: 949. https://doi.org/10.3390/ijms24020949

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