Next Article in Journal
Classical and Bayesian Inference of an Exponentiated Half-Logistic Distribution under Adaptive Type II Progressive Censoring
Next Article in Special Issue
Entropy Effects in Intermolecular Associations of Crown-Ethers and Cyclodextrins with Amino Acids in Aqueous and in Non-Aqueous Media
Previous Article in Journal
Clients’ Emotional Experiences Tied to Therapist-Led (but Not Client-Led) Physiological Synchrony during Imagery Rescripting
Previous Article in Special Issue
The Composition of Saturated Vapor over 1-Butyl-3-methylimidazolium Tetrafluoroborate Ionic Liquid: A Multi-Technique Study of the Vaporization Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermodynamic Properties of the First-Generation Hybrid Dendrimer with “Carbosilane Core/Phenylene Shell” Structure

by
Semen S. Sologubov
1,
Alexey V. Markin
1,*,
Natalia N. Smirnova
1,
Elena S. Chamkina
2,
Irina Yu. Krasnova
2,
Sergey A. Milenin
3,
Olga A. Serenko
2,
Zinaida B. Shifrina
2 and
Aziz M. Muzafarov
2,3
1
Chemistry Department, National Research Lobachevsky State University of Nizhny Novgorod, 23/5 Gagarin Av., 603950 Nizhny Novgorod, Russia
2
Nesmeyanov Institute of Organoelement Compounds of Russian Academy of Sciences, 28 Vavilov St., 119334 Moscow, Russia
3
Enikolopov Institute of Synthetic Polymeric Materials of Russian Academy of Sciences, 70 Profsoyuznaya St., 117393 Moscow, Russia
*
Author to whom correspondence should be addressed.
Entropy 2021, 23(12), 1557; https://doi.org/10.3390/e23121557
Submission received: 25 October 2021 / Revised: 21 November 2021 / Accepted: 22 November 2021 / Published: 23 November 2021

Abstract

:
The molar heat capacity of the first-generation hybrid dendrimer with a “carbosilane core/phenylene shell” structure was measured for the first time in the temperature range T = 6–600 K using a precise adiabatic vacuum calorimeter and DSC. In the above temperature interval, the glass transition of the studied compound was observed, and its thermodynamic characteristics were determined. The standard thermodynamic functions (the enthalpy, the entropy, and the Gibbs energy) of the hybrid dendrimer were calculated over the range from T = 0 to 600 K using the experimentally determined heat capacity. The standard entropy of formation of the investigated dendrimer was evaluated at T = 298.15 K. The obtained thermodynamic properties of the studied hybrid dendrimer were compared and discussed with the literature data for some of the first-generation organosilicon and pyridylphenylene dendrimers.

Graphical Abstract

1. Introduction

Dendrimers are a unique class of perfect monodisperse macromolecules having a highly branched three-dimensional architecture. The well-defined chemical structure of dendrimers consists of three major components: a central core (the multifunctional atomic group), an inner sphere (tree-like branches organized in generations G1, G2, G3, etc.), and an outer shell (comprising a large number of terminal functional groups) [1,2,3]. These structural features of dendrimers lead to their versatile physical and chemical properties, as well as contribute to their applications in nanotechnologies, biomedical fields (anticancer vaccines, contrast agents in magnetic resonance imaging, nanocarriers in targeted drug delivery, and gene therapy), electronics, and catalysis [4,5,6,7,8,9,10,11,12]. In contrast to traditional polymers, dendrimers are synthesized in a stepwise controlled manner. They can be constructed by two distinct synthetic strategies, namely, divergent and convergent approaches. The structural evolution of dendrimers is inextricably associated with a continuous improvement in traditional synthetic strategies, as well as the development of new synthetic tools for fundamental discoveries and practical applications [13].
Among various dendritic macromolecules, carbosilane (Si–C) dendrimers represent one of the most important and promising classes of silicon-containing dendrimers [14,15,16]. These compounds exhibit high flexibility, manifested by low glass transition temperatures, biological inertness, and kinetic and thermodynamic stability due to the low polarity and high strength of the Si–C bond. Therefore, carbosilane dendrimers have attracted much attention as functional molecules for successful applications in electrochemistry, catalysis, and nanoengineering [17]. Polyphenylene dendrimers are highly branched monodisperse macromolecules entirely consisting of substituted benzene rings that result in their rigid, shape-persistent chemical structure [18,19,20,21]. These dendrimers can be synthetically modified by the introduction of different chemical functionalities that influence their applications [22,23]. Polyphenylene dendrimers are extremely chemical and thermal stable molecules, but with greatly increased solubility in a wide range of organic solvents [24,25]. These advantages lead to their perspective applications as active materials in OLED devices, templates for composite materials, molecular sensors, magnetically recoverable catalysts, and bioapplications [26,27,28,29,30,31,32,33,34].
A new direction in dendrimer chemistry can be attributed to the synthesis and study of physicochemical properties of hybrid dendrimers consisting of two chemically different dendritic segments, each of which has an independent highly branched molecular architecture. The arrangement of the dendritic blocks in a different manner (i.e., a flexible core and a rigid shell or vice versa) allows for a controllable synthesis of principally new dendrimers with the desired characteristics [35]. The results of our calorimetric investigations of the widespread families such as carbosilane [36,37,38], siloxane [39,40,41], and polyphenylene [42,43] dendrimers with different molecular structure of the inner sphere and the surface functional groups were published earlier.
The thermodynamic properties of the first-generation hybrid dendrimer with a “carbosilane core/phenylene shell” structure have not yet been studied. These are the key data for technological and thermophysical calculations of processes with this dendrimer. The accumulation of reliable thermodynamic values for this class of macromolecular nanoobjects will subsequently make it possible to obtain practically important “structure–property” dependences. It allows predicting the physicochemical properties of dendrimer molecules that have not yet been studied.
As a continuation of publications on thermodynamics of dendrimers, the present study is dedicated to the calorimetric determination of the molar heat capacity of the first-generation hybrid dendrimer with a “carbosilane core/phenylene shell” structure, as well as the calculation of the standard thermodynamic functions (ΔH°, ΔS°, and ΔG°) in the range of T = 6–600 K and the estimation of the standard entropy of its formation at T = 298.15 K from the obtained experimental data. A comparative analysis of thermodynamic properties of the investigated hybrid dendrimer with the literature data for the previously studied organosilicon (carbosilane [36,37,38] and siloxane [39]) and pyridylphenylene dendrimers [42,43] of the first generation was performed to develop the fundamental principles of the regulation of different properties of dendrimers by varying their molecular parameters.

2. Experimental

2.1. Sample

The first-generation hybrid dendrimer with a “carbosilane core/phenylene shell” structure is denoted as G1{Si}13[Ar]32, where G1 indicates the generation number of dendrimer, {Si}13 represents the number of silicon atoms in the inner sphere of dendrimer, and [Ar]32 corresponds to the number of phenylene units on the outer layer of dendrimer. The molecular structure of the studied dendrimer is presented in Figure 1.
The sample of the hybrid dendrimer G1{Si}13[Ar]32 was synthesized at the Nesmeyanov Institute of Organoelement Compounds of Russian Academy of Sciences (Moscow) via the Diels–Alder cycloaddition reaction according to the procedure described in detail elsewhere [44]. The synthesized compound was a light-brown powder under the standard conditions and characterized by monodispersity on the generation number. The structure and the composition of dendrimer G1{Si}13[Ar]32 were confirmed by elemental analysis, 1H-, 13C-, and 29Si-NMR spectroscopy (Bruker AVANCE III HD spectrometer; 500.13 MHz for 1H; 125.76 MHz for 13C; CDCl3 was used as a solvent), and MALDI-TOF mass spectrometry (Bruker BIFLEX III spectrometer; a 337 nm nitrogen laser; indole-3-acrylic acid was used as a matrix, and silver trifluoroacetate was used as a cationization agent). The elemental analysis results for G1{Si}13[Ar]32 (brutto formula C304H332Si13), reported in [44], are as follows:
  • Found (%): C, 83.18; H, 7.51; Si, 8.41.
  • Calculated (%): C, 83.92; H, 7.69; Si, 8.39.
The sample of dendrimer G1{Si}13[Ar]32 was purified by gel permeation chromatography (GPC) in tetrahydrofuran (THF) using a Shimadzu system. According to the GPC analysis results, reported in [44], the final purity of dendrimer G1{Si}13[Ar]32 was 98%, and the yield of the pure compound was 0.45 g (82%).
The molar mass of dendrimer G1{Si}13[Ar]32 (M = 4350.95 g·mol−1) was calculated using the IUPAC table of atomic masses [45]. The sample information for dendrimer G1{Si}13[Ar]32 is summarized in Table 1.
The preliminary DSC measurements confirmed that the presence of phenylene fragments on the outer layer of hybrid dendrimer G1{Si}13[Ar]32 significantly increases the glass transition temperature compared to the entirely carbosilane dendrimers (Table 2). The thermal stability of hybrid dendrimer G1{Si}13[Ar]32 was studied using TG analysis. The decomposition process of the investigated sample in air is illustrated in Figure 2 and described by two steps: the destruction of the carbosilane skeleton (570–670 K) and the decomposition of the aromatic shell (770–970 K). On the contrary, a one-step mass loss (620–820 K), observed during the thermal decomposition of dendrimer G1{Si}13[Ar]32 in argon, is shown in Figure 3. As a result, the studied sample of hybrid dendrimer G1{Si}13[Ar]32 is thermally stable up to T ~620 K. The difference in TG curves is caused by the oxidation contribution in the decomposition process of the studied hybrid dendrimer. The results of DSC and TG analyses show that the surface layer of hybrid dendrimers strongly influences their thermal properties. In particular, the rigid phenylene shell considerably increases the glass transition temperature of hybrid dendrimers and makes them more thermally stable macromolecules than carbosilane and siloxane dendrimers.

2.2. Apparatus and Measurement Procedure

2.2.1. Adiabatic Vacuum Calorimetry

The heat capacity of the hybrid dendrimer with a “carbosilane core/phenylene shell” structure G1{Si}13[Ar]32 was measured using a thoroughly automated precise adiabatic calorimeter in the temperature range T = 6–350 K. The calorimeter construction and the operation procedure were described in detail elsewhere [46]. The calorimetric titanium cell (V = 1 cm3), containing the studied sample of dendrimer (m = 0.1989 g), was kept in vacuum and then filled with gaseous helium for improving thermal conduction. For the temperature measurements, a miniature iron–rhodium resistance thermometer (R = 100 Ω), calibrated according to the International Temperature Scale ITS-90, was used; the standard uncertainty for the temperature u(T) = 0.01 K. The temperature difference between the calorimeter and the adiabatic shield was measured using a four-junction copper-iron/chromel thermocouple. Liquid nitrogen and liquid helium were used in the low-temperature (T = (6–80) K) calorimetric experiments. The “AKSAMIT” data acquisition system, connected with a personal computer, performed the control of the measurement process, as well as the determination of the heat capacity values.
The calorimeter reliability was verified by the heat capacity measurements of the standard reference materials (benzoic acid C6H5COOH, corundum α-Al2O3, copper Cu) over the range of T = 6–350 K [47]. The relative expanded uncertainties for the heat capacity (Ur(Cp) = 0.02 below T = 15 K, Ur(Cp) = 0.005 between T = 15 and 40 K, and Ur(Cp) = 0.002 in the interval of T = 40–350 K) were evaluated on the basis of the calibration results.

2.2.2. Differential Scanning Calorimetry

A differential scanning calorimeter DSC 204 F1 Phoenix (NETZSCH) was used to determine the heat capacity of hybrid dendrimer G1{Si}13[Ar]32 in the temperature range T = 350–600 K. The calorimeter design and the experimental procedure were described in detail elsewhere [48]. The DSC 204 F1 Phoenix calorimeter was calibrated using the high-purity (>99.99%) reference samples, namely, indium (In), tin (Sn), bismuth (Bi), mercury (Hg), biphenyl ((C6H5)2), and cyclohexane (C6H12), recommended by the IUPAC Technical Report [49]. The calibration experiments were performed at a heating rate of 5 K·min−1 in an argon atmosphere with a gas flow of 25 mL·min−1. The heat capacity of the studied hybrid dendrimer was calculated using the ratio approach with synthetic sapphire (α-Al2O3) as a reference [50]. The dendrimer mass loaded into the DSC crucible was 12.53 mg. The DSC measurements were carried out in the interval of T = 350–600 K at a heating rate of 5 K·min−1 under an argon purge (25 mL·min−1). The NETZSCH Proteus software was used to analyze the obtained DSC results. As a result, the standard uncertainty for the temperature u(T) = 0.2 K and the relative expanded uncertainty for the heat capacity Ur(Cp) = 0.02 were determined in the range of T = 350–600 K.

3. Results and Discussion

3.1. Heat Capacity

The temperature dependence of the molar heat capacity of the hybrid dendrimer with a “carbosilane core/phenylene shell” structure G1{Si}13[Ar]32 is presented in Figure 4. The Cp,m values of the studied dendrimer in the experimental temperature range T = 6–600 K are listed in Table S1 (Supplementary Materials). The mass loss of hybrid dendrimer G1{Si}13[Ar]32 was not observed after the calorimetric experiments, and this is in good agreement with the results of TG analysis. The heat capacity of hybrid dendrimer G1{Si}13[Ar]32 was not more than 25–60% of the total heat capacity of the calorimetric ampoule filled with the substance.

3.2. Standard Thermodynamic Characteristics of the Glass Transition and the Glassy State

The studied hybrid dendrimer G1{Si}13[Ar]32 was devitrified during the heat capacity measurements in the interval of T = 300–350 K. The glass transition temperature T g o = 323 ± 1 K was determined using the method of Alford and Dole [51]. The heat capacity increase at the glass transition temperature Δ C p o ( T g o ) = 1312 ± 16 J·K−1·mol−1 was evaluated graphically. The configuration entropy S conf o = 334 ± 5 J·K−1·mol−1 was calculated using Equation (1) [52].
S conf o = Δ C p o ( T g o ) ln ( T g o / T K ) ,
where TK is the Kauzmann temperature [53], and the ratio ( T g o / T K ) = 1.29 ± 0.14 [54]. The assumption S conf o S°(0) was used to determine the absolute entropy S°(T) of dendrimer G1{Si}13[Ar]32.
The standard thermodynamic characteristics of the glass transition and glassy state of the studied hybrid dendrimer G1{Si}13[Ar]32 are presented in Table 2 in comparison to the data for the investigated organosilicon [35,40,41,42] and pyridylphenylene [45,46] dendrimers of the first generation. Molecular structures of the calorimetrically studied G1 dendrimers are shown in Figure 5.
As can be seen from Table 2, the studied hybrid dendrimer G1{Si}13[Ar]32 was characterized by the highest glass transition temperature in comparison to the entirely carbosilane dendrimer G1{Si}5[CH2CH=CH2]8 [36] and siloxane dendrimer G1[OSi(CH3)3]6 [39]. Furthermore, the studied hybrid dendrimer had the highest S conf o value among the other dendrimers of the first generation. The configuration entropy can characterize the mobility of the individual macromolecule parts. Thus, the S conf o can take a minimum value or is equal to zero for strictly ordered systems. In our case, the value of S conf o = 334 J·K−1·mol−1 for hybrid dendrimer G1{Si}13[Ar]32 may indicate a disordered structure of the dendritic macromolecule. The intermediate T g o values were revealed in case of carbosilane dendrimers G1{Si}13[(C6H4)C3H5O2]8 [37] and G1{Si}13[CH2CH2C6H5]8 [38] containing the cyclic phenyl and heterocyclic dioxolane units on the surface layer. The previously studied pyridylphenylene dendrimer G1[C5H4N]12 [43] had exactly the same T g o value as the investigated hybrid dendrimer G1{Si}13[Ar]32. Moreover, the absolutely identical behavior of the heat capacity curves can be explained by the similar structure of these macromolecules due to the presence of rigid phenylene fragments. As a result, the glass transition temperature of dendrimers with different molecular skeleton and functional terminal groups largely depends on the structure of their periphery layer.

3.3. Standard Thermodynamic Functions

The standard thermodynamic functions of hybrid dendrimer G1{Si}13[Ar]32 were calculated in the interval of T = 6–600 K. The graphical extrapolation of the low-temperature Cp,m values to absolute zero was performed in accordance with the Debye T3 law [55].
C p o = n D ( Θ D / T ) ,
where D is the Debye function, and n and ΘD are specially selected parameters (n = 30, ΘD = 49.0 K). The experimental Cp,m values of the investigated dendrimer were described by Equation (2) with the relative expanded uncertainty Ur(Cp,m) = 0.014.
The standard thermodynamic functions (the enthalpy (H°(T) − H°(0)), the entropy (S°(T) − S°(0)), and the Gibbs energy (G°(T) − H°(0))) of dendrimer G1{Si}13[Ar]32 are listed in Table 3. The calculation procedure was described in detail elsewhere [56,57].
The standard entropy of formation ΔfS° of dendrimer G1{Si}13[Ar]32 was calculated at a reference T = 298.15 K using the value of (S°(T) − S°(0)) = 5865 J·K−1·mol−1 (Table 3), the residual entropy S°(0) = 334 J·K−1·mol−1 (Table 2), and the absolute entropies of the elemental substances C(gr) (S° = 5.74 J·K−1·mol−1), H2(g) (S° = 130.68 J·K−1·mol−1), and Si(cr) (S° = 18.82 J·K−1·mol−1) recommended by Chase [58]. The obtained ΔfS° value,
ΔfS°(G1{Si}13[Ar]32, 298.15) = −(17,484 ± 97) J·K−1·mol−1,
corresponds to the following formation reaction of dendrimer:
304C(gr) + 166H2(g) + 13Si(cr) → C304H332Si13(gl),
where the physical states are given in parentheses (gr—graphite, g—gas, cr—crystal, gl—glassy state).

4. Conclusions

In this work, we reported the results of a calorimetric study of the first-generation hybrid dendrimer with a “carbosilane core/phenylene shell” structure. The temperature dependence of the heat capacity of the studied compound was determined in the range of T = 6–600 K by precise adiabatic calorimetry and DSC. The glass transition of hybrid dendrimer G1{Si}13[Ar]32 was observed in the experimental temperature interval. The standard thermodynamic characteristics of this transformation were determined and compared with those of the studied earlier organosilicon and pyridylphenylene dendrimers with different functional terminal groups. The standard thermodynamic functions of hybrid dendrimer G1{Si}13[Ar]32 were calculated over the range T = 0–600 K. The standard entropy of formation of the studied compound was evaluated at T = 298.15 K. It was confirmed that the structure of the surface layer of dendrimers strongly influences their thermal behavior. The rigid phenylene shell considerably increases the glass transition temperature of hybrid dendrimers and makes them more thermally stable macromolecules than carbosilane and siloxane dendrimers.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/e23121557/s1, Experimental values of the molar heat capacity of hybrid dendrimer G1{Si}13[Ar]32 (Table S1).

Author Contributions

Conceptualization, A.V.M.; methodology, A.V.M. and N.N.S.; validation, S.S.S., A.V.M. and N.N.S.; formal analysis, S.S.S.; investigation, S.S.S., E.S.C., I.Y.K. and S.A.M.; writing—original draft preparation, S.S.S.; writing—review and editing, S.S.S. and A.V.M.; supervision, A.V.M.; project administration, A.V.M., O.A.S., Z.B.S. and A.M.M.; funding acquisition, O.A.S. and Z.B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Russian Science Foundation (Grant No. 20-13-00025).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fréchet, J.M.J.; Tomalia, D.A. Dendrimers and Other Dendritic Polymers; John Wiley & Sons: Chichester, UK, 2001. [Google Scholar] [CrossRef]
  2. Newkome, G.R.; Moorefield, C.N.; Vögtle, F. Dendrimers and Dendrons: Concepts, Syntheses, Applications; Wiley-VCH: Weinheim, Germany, 2001. [Google Scholar] [CrossRef]
  3. Tomalia, D.A.; Christensen, J.B.; Boas, U. Dendrimers, Dendrons, and Dendritic Polymers: Discovery, Applications, and the Future; Cambridge University Press: Cambridge, UK, 2012. [Google Scholar] [CrossRef]
  4. Inoue, K. Functional dendrimers, hyperbranched and star polymers. Prog. Polym. Sci. 2000, 25, 453–571. [Google Scholar] [CrossRef]
  5. Grayson, S.M.; Fréchet, J.M.J. Convergent dendrons and dendrimers: From synthesis to applications. Chem. Rev. 2001, 101, 3819–3867. [Google Scholar] [CrossRef]
  6. van Heerbeek, R.; Kamer, P.C.J.; van Leeuwen, P.W.N.M.; Reek, J.N.H. Dendrimers as support for recoverable catalysts and reagents. Chem. Rev. 2002, 102, 3717–3756. [Google Scholar] [CrossRef] [PubMed]
  7. Boas, U.; Heegaard, P.M.H. Dendrimers in drug research. Chem. Soc. Rev. 2004, 33, 43–63. [Google Scholar] [CrossRef]
  8. Gade, L.H. Dendrimer Catalysis; Springer: Heidelberg, Germany, 2006. [Google Scholar] [CrossRef]
  9. Astruc, D.; Boisselier, E.; Ornelas, C. Dendrimers Designed for Functions: From Physical, Photophysical, and Supramolecular Properties to Applications in Sensing, Catalysis, Molecular Electronics, Photonics, and Nanomedicine. Chem. Rev. 2010, 110, 1857–1959. [Google Scholar] [CrossRef] [PubMed]
  10. Svenson, S.; Tomalia, D.A. Dendrimers in biomedical applications—Reflections on the field. Adv. Drug Deliv. Rev. 2012, 64, 102–115. [Google Scholar] [CrossRef]
  11. Sebestik, J.; Reinis, M.; Jezek, J. Biomedical Applications of Peptide-, Glyco- and Glycopeptide Dendrimers, and Analogous Dendrimeric Structures; Springer: Vienna, Austria, 2012. [Google Scholar] [CrossRef]
  12. Yang, J.; Zhang, Q.; Chang, H.; Cheng, Y. Surface-Engineered Dendrimers in Gene Delivery. Chem. Rev. 2015, 115, 5274–5300. [Google Scholar] [CrossRef] [PubMed]
  13. Carlmark, A.; Hawker, C.; Hult, A.; Malkoch, M. New methodologies in the construction of dendritic materials. Chem. Soc. Rev. 2009, 38, 352–362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Majoral, J.-P.; Caminade, A.-M. Dendrimers Containing Heteroatoms (Si, P, B, Ge, or Bi). Chem. Rev. 1999, 99, 845–880. [Google Scholar] [CrossRef]
  15. Frey, H.; Schlenk, C. Silicon-based dendrimers. Top. Curr. Chem. 2000, 210, 69–129. [Google Scholar] [CrossRef]
  16. Rebrov, E.A.; Leshchiner, I.D.; Muzafarov, A.M. Synthesis of carbosilane dendrimers with variable distance between branching nodes. Macromolecules 2012, 45, 8796–8804. [Google Scholar] [CrossRef]
  17. Dvornic, P.R.; Owen, M.J. Silicon-Containing Dendritic Polymers; Springer: Dordrecht, The Netherlands, 2009. [Google Scholar] [CrossRef]
  18. Miller, T.M.; Neenan, T.X. Convergent synthesis of monodisperse dendrimers based upon 1,3,5-trisubstituted benzenes. Chem. Mater. 1990, 2, 346–349. [Google Scholar] [CrossRef]
  19. Morgenroth, F.; Reuther, E.; Müllen, K. Polyphenylene dendrimers: From three-dimensional to two-dimensional structures. Angew. Chem. Int. Ed. 1997, 36, 631–634. [Google Scholar] [CrossRef]
  20. Morgenroth, F.; Kübel, C.; Müllen, K. Nanosized polyphenylene dendrimers based upon pentaphenylbenzene units. J. Mater. Chem. 1997, 7, 1207–1211. [Google Scholar] [CrossRef]
  21. Morgenroth, F.; Müllen, K. Dendritic and hyperbranched polyphenylenes via a simple Diels–Alder route. Tetrahedron 1997, 53, 15349–15366. [Google Scholar] [CrossRef]
  22. Hammer, B.A.G.; Müllen, K. Dimensional evolution of polyphenylenes: Expanding in all directions. Chem. Rev. 2016, 116, 2103–2140. [Google Scholar] [CrossRef] [PubMed]
  23. Hammer, B.A.G.; Wu, Y.; Fischer, S.; Liu, W.; Weil, T.; Müllen, K. Controlling cellular uptake and toxicity of polyphenylene dendrimers by chemical functionalization. ChemBioChem 2017, 18, 960–964. [Google Scholar] [CrossRef] [PubMed]
  24. Shifrina, Z.B.; Rajadurai, M.S.; Firsova, N.V.; Bronstein, L.M.; Huang, X.; Rusanov, A.L.; Muellen, K. Poly(phenylene-pyridyl) dendrimers: Synthesis and templating of metal nanoparticles. Macromolecules 2005, 38, 9920–9932. [Google Scholar] [CrossRef]
  25. Rajadurai, M.S.; Shifrina, Z.B.; Kuchkina, N.V.; Rusanov, A.L.; Müllen, K. Rigid aromatic dendrimers. Russ. Chem. Rev. 2007, 76, 767–783. [Google Scholar] [CrossRef]
  26. Bronstein, L.M.; Shifrina, Z.B. Dendrimers as encapsulating, stabilizing, or directing agents for inorganic nanoparticles. Chem. Rev. 2011, 111, 5301–5344. [Google Scholar] [CrossRef]
  27. Yuzik-Klimova, E.Y.; Kuchkina, N.V.; Sorokina, S.A.; Morgan, D.G.; Boris, B.; Nikoshvili, L.Z.; Lyubimova, N.A.; Matveeva, V.G.; Sulman, E.M.; Stein, B.D.; et al. Bronstein, Magnetically recoverable catalysts based on polyphenylenepyridyl dendrons and dendrimers. RSC Adv. 2014, 4, 23271–23280. [Google Scholar] [CrossRef]
  28. Kuchkina, N.V.; Morgan, D.G.; Kostopoulou, A.; Lappas, A.; Brintakis, K.; Boris, B.S.; Yuzik-Klimova, E.Y.; Stein, B.D.; Svergun, D.I.; Spilotros, A.; et al. Hydrophobic periphery tails of polyphenylenepyridyl dendrons control nanoparticle formation and catalytic properties. Chem. Mater. 2014, 26, 5654–5663. [Google Scholar] [CrossRef] [Green Version]
  29. Serenko, O.; Strashnov, P.; Kapustin, G.; Kalinin, M.; Kuchkina, N.; Serkova, E.; Shifrina, Z.; Muzafarov, A. Adsorption properties of pyridylphenylene dendrimers. RSC Adv. 2017, 7, 7870–7875. [Google Scholar] [CrossRef] [Green Version]
  30. Sorokina, S.A.; Kuchkina, N.V.; Lawson, B.P.; Krasnova, I.Y.; Nemygina, N.A.; Nikoshvili, L.Z.; Talanova, V.N.; Stein, B.D.; Pink, M.; Morgan, D.G.; et al. Pyridylphenylene dendrons immobilized on the surface of chemically modified magnetic silica as efficient stabilizing molecules of Pd species. Appl. Surf. Sci. 2019, 488, 865–873. [Google Scholar] [CrossRef]
  31. Schmaltz, B.; Weil, T.; Müllen, K. Polyphenylene-based materials: Control of the electronic function by molecular and supramolecular complexity. Adv. Mater. 2009, 21, 1067–1078. [Google Scholar] [CrossRef]
  32. Türp, D.; Nguyen, T.T.T.; Baumgarten, M.; Müllen, K. Uniquely versatile: Nano-site defined materials based on polyphenylene dendrimers. New J. Chem. 2012, 36, 282–298. [Google Scholar] [CrossRef]
  33. Stangenberg, R.; Wu, Y.; Hedrich, J.; Kurzbach, D.; Wehner, D.; Weidinger, G.; Kuan, S.L.; Jansen, M.I.; Jelezko, F.; Luhmann, H.J.; et al. A polyphenylene dendrimer drug transporter with precisely positioned amphiphilic surface patches. Adv. Healthc. Mater. 2015, 4, 377–384. [Google Scholar] [CrossRef] [PubMed]
  34. Wagner, J.; Dillenburger, M.; Simon, J.; Oberländer, J.; Landfester, K.; Mailänder, V.; Ng, D.Y.W.; Müllen, K.; Weil, T. Amphiphilic dendrimers control protein binding and corona formation on liposome nanocarriers. Chem. Commun. 2020, 56, 8663–8666. [Google Scholar] [CrossRef]
  35. Milenin, S.A.; Cherkaev, G.V.; Demchenko, N.V.; Serkova, E.S.; Krasnova, I.Y.; Selezneva, E.V.; Buzin, M.I.; Bakirov, A.V.; Vasil’ev, V.G.; Shifrina, Z.B.; et al. Influence of the growing flexible shell on the molecular behavior of hybrid dendrimers. Macromolecules 2020, 53, 9706–9716. [Google Scholar] [CrossRef]
  36. Lebedev, B.V.; Ryabkov, M.V.; Tatarinova, E.A.; Rebrov, E.A.; Muzafarov, A.M. Thermodynamic properties of the first to fifth generations of carbosilane dendrimers with allyl terminal groups. Russ. Chem. Bull. 2003, 52, 545–551. [Google Scholar] [CrossRef]
  37. Smirnova, N.N.; Sologubov, S.S.; Sarmini, Y.A.; Markin, A.V.; Novozhilova, N.A.; Tatarinova, E.A.; Muzafarov, A.M. Thermodynamic properties of first- and third-generation carbosilane dendrimers with terminal phenyldioxolane groups. Russ. J. Phys. Chem. A 2017, 91, 2317–2325. [Google Scholar] [CrossRef]
  38. Sologubov, S.S.; Markin, A.V.; Smirnova, N.N.; Novozhilova, N.A.; Tatarinova, E.A.; Muzafarov, A.M. Thermodynamic properties of a first-generation carbosilane dendrimer with terminal phenylethyl groups. Russ. J. Phys. Chem. A 2018, 92, 235–243. [Google Scholar] [CrossRef]
  39. Markin, A.V.; Sarmini, Y.A.; Sologubov, S.S.; Smirnova, N.N.; Boldyrev, K.L.; Tatarinova, E.A.; Meshkov, I.B.; Muzafarov, A.M. Thermodynamic properties of a first-generation siloxane dendrimer with terminal trimethylsilyl groups. Russ. J. Phys. Chem. A 2020, 94, 240–248. [Google Scholar] [CrossRef]
  40. Sologubov, S.S.; Markin, A.V.; Sarmini, Y.A.; Samosudova, Y.S.; Smirnova, N.N.; Boldyrev, K.L.; Tatarinova, E.A.; Meshkov, I.B.; Muzafarov, A.M. Calorimetric study of siloxane dendrimer of the third generation with trimethylsilyl terminal groups. J. Therm. Anal. Calorim. 2019, 138, 3301–3310. [Google Scholar] [CrossRef]
  41. Sologubov, S.S.; Markin, A.V.; Sarmini, Y.A.; Smirnova, N.N.; Boldyrev, K.L.; Tatarinova, E.A.; Meshkov, I.B.; Muzafarov, A.M. Thermodynamic investigation of G2 and G4 siloxane dendrimers with trimethylsilyl terminal groups. J. Chem. Thermodyn. 2021, 153, 106318. [Google Scholar] [CrossRef]
  42. Smirnova, N.N.; Markin, A.V.; Zakharova, Y.A.; Kuchkina, N.V.; Rusanov, A.L.; Shifrina, Z.B. Thermodynamic properties of pyridine-containing polyphenylene dendrimers of the first–fourth generations. Russ. Chem. Bull. 2011, 60, 132–138. [Google Scholar] [CrossRef]
  43. NSmirnova, N.N.; Zakharova, Y.A.; Markin, A.V.; Kuchkina, N.V.; Yuzik-Klimova, E.Y.; Shifrina, Z.B. Thermodynamics of hard poly(phenylene-pyridyl) dendrimers. Russ. Chem. Bull. 2013, 62, 2258–2262. [Google Scholar] [CrossRef]
  44. Serkova, E.S.; Krasnova, I.Y.; Milenin, S.; Selezneva, E.V.; Tatarinova, E.A.; Boldyrev, K.L.; Korlyukov, A.A.; Zubavichus, Y.V.; Buzin, M.I.; Serenko, O.A.; et al. Core/shell hybrid dendrimers: Controllable rigidity determines molecular behaviour. Polymer 2018, 138, 83–91. [Google Scholar] [CrossRef]
  45. Meija, J.; Coplen, T.B.; Berglund, M.; Brand, W.A.; De Bièvre, P.; Gröning, M.; Holden, N.E.; Irrgeher, J.; Loss, R.D.; Walczyk, T.; et al. Atomic weights of the elements 2013 (IUPAC Technical Report). Pure Appl. Chem. 2016, 88, 265–291. [Google Scholar] [CrossRef]
  46. Varushchenko, R.M.; Druzhinina, A.I.; Sorkin, E.L. Low-temperature heat capacity of 1-bromoperfluorooctane. J. Chem. Thermodyn. 1997, 29, 623–637. [Google Scholar] [CrossRef]
  47. Sabbah, R.; Xu-Wu, A.; Chickos, J.S.; Leitão, M.L.P.; Roux, M.V.; Torres, L.A. Reference materials for calorimetry and differential thermal analysis. Thermochim. Acta 1999, 331, 93–204. [Google Scholar] [CrossRef]
  48. Höhne, G.W.H.; Hemminger, W.F.; Flammersheim, H.-J. Differential Scanning Calorimetry, 2nd ed.; Springer: Berlin/Heidelberg, Germany, 2003. [Google Scholar] [CrossRef]
  49. della Gatta, G.; Richardson, M.J.; Sarge, S.M.; Stølen, S. Standards, calibration, and guidelines in microcalorimetry. Part 2. Calibration standards for differential scanning calorimetry (IUPAC Technical Report). Pure Appl. Chem. 2006, 78, 1455–1476. [Google Scholar] [CrossRef] [Green Version]
  50. Kaisersberger, E.; Janoschek, J.; Wassmer, E. A heat flux DSC for enthalpy and specific heat determinations to 1700 K. Thermochim. Acta 1989, 148, 499–505. [Google Scholar] [CrossRef]
  51. Alford, S.; Dole, M. Specific heat of synthetic high polymers. VI. A study of the glass transition in polyvinyl chloride. J. Am. Chem. Soc. 1955, 77, 4774–4777. [Google Scholar] [CrossRef]
  52. Adam, G.; Gibbs, J.H. On the Temperature Dependence of Cooperative Relaxation Properties in Glass-Forming Liquids. J. Chem. Phys. 1965, 43, 139–146. [Google Scholar] [CrossRef] [Green Version]
  53. Kauzmann, W. The Nature of the Glassy State and the Behavior of Liquids at Low Temperatures. Chem. Rev. 1948, 43, 219–256. [Google Scholar] [CrossRef]
  54. Bestul, A.B.; Chang, S.S. Excess Entropy at Glass Transformation. J. Chem. Phys. 1964, 40, 3731–3733. [Google Scholar] [CrossRef]
  55. Debye, P. Zur Theorie der spezifischen Wärmen. Ann. Phys. 1912, 344, 789–839. [Google Scholar] [CrossRef] [Green Version]
  56. JMcCullough, P.; Scott, D.W. Experimental Thermodynamics. Calorimetry of Non-Reacting Systems; Butterworth & Co. Ltd.: London, UK, 1968; Volume I, Available online: https://sciencedirect.com/science/book/9781483213279 (accessed on 22 November 2021).
  57. Lebedev, B. Application of precise calorimetry in study of polymers and polymerization processes. Thermochim. Acta 1997, 297, 143–149. [Google Scholar] [CrossRef]
  58. Chase, M.W. NIST-JANAF Thermochemical Tables, 4th ed.; Monograph No. 9; National Institute of Standards and Technology: Gaithersburg, MD, USA, 1998; pp. 1–1951. [CrossRef]
Figure 1. Molecular structure of the first-generation hybrid dendrimer with a “carbosilane core/phenylene shell” structure, G1{Si}13[Ar]32.
Figure 1. Molecular structure of the first-generation hybrid dendrimer with a “carbosilane core/phenylene shell” structure, G1{Si}13[Ar]32.
Entropy 23 01557 g001
Figure 2. Thermogravimetric curve of hybrid dendrimer G1{Si}13[Ar]32 in air.
Figure 2. Thermogravimetric curve of hybrid dendrimer G1{Si}13[Ar]32 in air.
Entropy 23 01557 g002
Figure 3. Thermogravimetric curve of hybrid dendrimer G1{Si}13[Ar]32 in argon.
Figure 3. Thermogravimetric curve of hybrid dendrimer G1{Si}13[Ar]32 in argon.
Entropy 23 01557 g003
Figure 4. Temperature dependence of the molar heat capacity of hybrid dendrimer G1{Si}13[Ar]32 (circles (6–350 K) and the solid line—(350–600 K) correspond to measurements in adiabatic calorimeter and DSC, respectively).
Figure 4. Temperature dependence of the molar heat capacity of hybrid dendrimer G1{Si}13[Ar]32 (circles (6–350 K) and the solid line—(350–600 K) correspond to measurements in adiabatic calorimeter and DSC, respectively).
Entropy 23 01557 g004
Figure 5. Molecular structures of the calorimetrically studied organosilicon and pyridylphenylene dendrimers of the first generation.
Figure 5. Molecular structures of the calorimetrically studied organosilicon and pyridylphenylene dendrimers of the first generation.
Entropy 23 01557 g005aEntropy 23 01557 g005b
Table 1. Sample information.
Table 1. Sample information.
Designation of DendrimerBrutto FormulaSourcePurification MethodFinal Mass Fraction PurityAnalysis Method
G1{Si}13[Ar]32C304H332Si13Synthesis [44]Gel permeation chromatography0.98Elemental analysis, NMR spectroscopy, MALDI-TOF mass spectrometry
Table 2. The standard thermodynamic characteristics of the glass transition and glassy state of the first-generation dendrimers with different molecular skeleton and functional terminal groups at p° = 0.1 MPa a.
Table 2. The standard thermodynamic characteristics of the glass transition and glassy state of the first-generation dendrimers with different molecular skeleton and functional terminal groups at p° = 0.1 MPa a.
Type of DendrimerDesignation of DendrimerΔT (K) T g o ±   1   ( K ) Δ C p o ( T g o ) ( J · K 1 · mol 1 ) S c o n f o   S ° ( 0 )   ( J · K 1 · mol 1 ) Source
HybridG1{Si}13[Ar]32300–3503231312 ± 16334 ± 5This work
CarbosilaneG1{Si}5[CH2CH=CH2]8150–160154406 ± 5103 ± 2[36]
G1{Si}13[(C6H4)C3H5O2]8200–2602311180 ± 11301 ± 3[37]
G1{Si}13[CH2CH2C6H5]8176–215198960 ± 10245 ± 3[38]
SiloxaneG1[OSi(CH3)3]6137–153147245 ± 363 ± 1[39]
PyridylphenyleneG1[C5H4N]12290–350323225 ± 357 ± 1[43]
a The standard uncertainty for pressure u(p) = 10 kPa. The reported expanded uncertainties correspond to the 0.95 level of confidence (coverage factor k ≈ 2).
Table 3. The standard thermodynamic functions of hybrid dendrimer G1{Si}13[Ar]32 (M(C304H332Si13) = 4350.95 g·mol−1) at p° = 0.1 MPa a.
Table 3. The standard thermodynamic functions of hybrid dendrimer G1{Si}13[Ar]32 (M(C304H332Si13) = 4350.95 g·mol−1) at p° = 0.1 MPa a.
T (K) C p o ( T ) ( kJ · K 1 · mol 1 ) (H°(T) − H°(0))
(kJ·mol−1)
(S°(T) − S°(0))
(kJ·K−1·mol−1)
−(G°(T) − H°(0))
(kJ·mol−1)
Amorphous (glassy) state
50.01960.02490.006670.00838
100.09480.2980.041610.118
150.2241.080.1030.468
200.36892.5690.18791.189
300.64527.6380.38974.053
400.911215.460.61299.056
501.14225.730.841216.33
601.35938.231.06825.88
701.57252.871.29437.69
801.79869.741.51951.75
902.00588.691.74168.04
1002.196109.71.96386.56
1503.034240.93.015211.4
2003.862413.84.004387.1
2504.692628.14.957611.1
298.155.675877.05.865871.6
3005.706887.65.900882.5
3206.07010066.2821004
3236.10010226.3311020
Amorphous (devitrified) state
3237.41210226.3311020
3307.53410766.4991068
3507.73612296.9491203
4008.21616288.0141577
4508.63820509.0062003
5009.01624919.9362477
5509.358295110.812996
6009.671342711.643557
a The standard uncertainty for pressure u(p) = 10 kPa. The standard uncertainties for temperature u(T) = 0.01 K in the interval of T = 5–350 K and u(T) = 0.5 K in the range of T = 350–600 K. The combined expanded relative uncertainties Uc,r( C p o ( T ) ) = 0.02, 0.005, 0.002, and 0.02; Uc,r([H°(T) − H°(0)]) = 0.022, 0.007, 0.005, and 0.022; Uc,r([S°(T) − S°(0)]) = 0.023, 0.008, 0.006, and 0.023; Uc,r([G°(T) − H°(0)]) = 0.03, 0.01, 0.009, and 0.03 in the intervals of T = 5–15 K, T = 15–40 K, T = 40–350 K, and T = 350–600 K, respectively. The reported expanded uncertainties correspond to the 0.95 level of confidence (coverage factor k ≈ 2).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sologubov, S.S.; Markin, A.V.; Smirnova, N.N.; Chamkina, E.S.; Krasnova, I.Y.; Milenin, S.A.; Serenko, O.A.; Shifrina, Z.B.; Muzafarov, A.M. Thermodynamic Properties of the First-Generation Hybrid Dendrimer with “Carbosilane Core/Phenylene Shell” Structure. Entropy 2021, 23, 1557. https://doi.org/10.3390/e23121557

AMA Style

Sologubov SS, Markin AV, Smirnova NN, Chamkina ES, Krasnova IY, Milenin SA, Serenko OA, Shifrina ZB, Muzafarov AM. Thermodynamic Properties of the First-Generation Hybrid Dendrimer with “Carbosilane Core/Phenylene Shell” Structure. Entropy. 2021; 23(12):1557. https://doi.org/10.3390/e23121557

Chicago/Turabian Style

Sologubov, Semen S., Alexey V. Markin, Natalia N. Smirnova, Elena S. Chamkina, Irina Yu. Krasnova, Sergey A. Milenin, Olga A. Serenko, Zinaida B. Shifrina, and Aziz M. Muzafarov. 2021. "Thermodynamic Properties of the First-Generation Hybrid Dendrimer with “Carbosilane Core/Phenylene Shell” Structure" Entropy 23, no. 12: 1557. https://doi.org/10.3390/e23121557

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