Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 18 de 18
Filter
Add more filters










Publication year range
1.
Soft Matter ; 17(17): 4496-4503, 2021 May 05.
Article in English | MEDLINE | ID: mdl-33949603

ABSTRACT

We report a mechanical metamaterial-like behavior as a function of the micro/nanostructure of otherwise chemically identical aliphatic polyurea aerogels. Transmissibility varies dramatically with frequency in these aerogels. Broadband vibration mitigation is provided at low frequencies (500-1000 Hz) through self-assembly of locally resonant metastructures wherein polyurea microspheres are embedded in a polyurea web-like network. A micromechanical constitutive model based on a discrete element method is established to explain the vibration mitigation mechanism. Simulations confirm the metamaterial-like behavior with a negative dynamic material stiffness for the micro-metastructured aerogels in a much wider frequency range than the majority of previously reported locally resonant metamaterials.

2.
ACS Appl Mater Interfaces ; 11(37): 34292-34304, 2019 Sep 18.
Article in English | MEDLINE | ID: mdl-31490651

ABSTRACT

Polyurethane aerogels were prepared from a rigid aromatic triisocyanate (tris(4-isocyanatophenyl)methane) and cage-shaped α- and ß-cyclodextrins as rigid polyols. Gelation was carried out in DMF using dibutyltin dilaurate as catalyst. Wet-gels were dried to aerogels (abbreviated as α- or ß-CDPU-xx) with supercritical fluid CO2. "xx" stands for the percent weight of the two monomers in the sol and was varied at two levels for each cyclodextrin: 2.5% and 15%. All aerogels were characterized with solid-state 13C and 15N NMR, CHN analysis, FTIR, XPS, SEM, and gas (N2 and CO2) sorption porosimetry. α- and ß-CDPU-xx aerogels were investigated as desiccants at room temperature. All materials had relatively higher capacities for water adsorption from high-humidity environments (99%) than typical commercial desiccants like silica or Drierite. However, α-CDPU-2.5 aerogels did stand out with a water uptake capacity reaching 1 g of H2O per gram of material. Most importantly though, adsorbed water could be released quantitatively without heating, by just reducing the relative humidity of the environment to 10%. All α- and ß-CDPU-xx aerogel samples were cycled between humid and dry environments 10 times. Their unusual behavior was traced to filling smaller mesopores with water and was attributed to a delicate balance of enthalpic (H-bonding) and entropic factors, whereas the latter are a function of pore sizes.

3.
ACS Appl Mater Interfaces ; 11(25): 22668-22676, 2019 Jun 26.
Article in English | MEDLINE | ID: mdl-31184858

ABSTRACT

A new route to metallic aerogels that bypasses the use of supercritical fluids and handling fragile wet-gel and aerogel precursors is exemplified by the carbothermal synthesis of monolithic Co(0) aerogels from compressed cobaltia xerogel powders coated conformally (cross-linked) at the primary particle level with a carbonizable polyurea. Residual carbon is removed, and carbon-free samples are obtained by high-temperature treatment of as-prepared Co(0) aerogels under a flowing stream of H2O/H2 that prevents oxidation of the Co(0) network. The durability of Co(0) aerogels is demonstrated under harsh processing conditions in their application as thermites. For this, Co(0) aerogel discs are infiltrated with LiClO4 from a melt and are ignited at about 1100 °C with an electric resistor. As Co(0) "burns" to CoO, temperature exceeds 1500 °C, and the heat released (-552 ± 2 kcal mol-1) is near to both the theoretical value (-58.47 kcal mol-1) and that from well-known pressed-pellet iron/perchlorate thermites (-66.6 kcal mol-1). The advantage of nanostructured thermites based on Co(0) aerogels is the efficiency (100%) by which the metal is consumed during its reaction with LiClO4 filling the pores.

4.
ACS Nano ; 13(3): 3677-3690, 2019 Mar 26.
Article in English | MEDLINE | ID: mdl-30839204

ABSTRACT

Morphology is a qualitative property of nanostructured matter and is articulated by visual inspection of micrographs. For deterministic procedures that relate nanomorphology to synthetic conditions, it is necessary to express nano- and microstructures numerically. Selecting polyurea aerogels as a model system with demonstrated potential for rich nanomorphology and guided by a statistical design-of-experiments model, we prepared a large array of materials (208) with identical chemical composition but quite different nanostructures. By reflecting on SEM imaging, it was realized that our first preverbal impression about a nanostructure is related to its openness and texture; the former is quantified by porosity ( Π), and the latter is oftentimes related to hydrophobicity, which, in turn, is quantified by the contact angle (θ) of water droplets resting on the material. Herewith, the θ-to-Π ratio is referred to as the K-index, and it was noticed that all polyurea samples of this study could be put in eight K-index groups with separate nanomorphologies ranging from caterpillar-like assemblies of nanoparticles, to thin nanofibers, to cocoon-like structures, to large bald microspheres. A first validation of the K-index as a morphology descriptor was based on compressing samples to different strains: it was observed that as the porosity decreases, the water-contact angle decreases proportionally, and thereby the K-index remains constant. The predictive power of the K-index was demonstrated with 20 polyurea aerogels prepared in 8 binary solvent systems. Subsequently, several material properties were correlated to nanomorphology through the K-index and that, in turn, provided insight about the root cause of the diversity of the nanostructure in polyurea aerogels. Finally, using response surface methodology, K-indexes and other material properties of practical interest were correlated to the monomer, water, and catalyst concentrations as well as the three Hansen solubility parameters of the sol. That enabled the synthesis of materials with up to six prescribed properties at a time, including nanomorphology, bulk density, BET surface area, elastic modulus, ultimate compressive strength, and thermal conductivity.

5.
ACS Appl Mater Interfaces ; 10(27): 23321-23334, 2018 Jul 11.
Article in English | MEDLINE | ID: mdl-29947503

ABSTRACT

Thermodynamic-kinetic relationships are not uncommon, but rigorous correlations are rare. On the basis of the parabolic free-energy profiles of elastic deformation, a generalized Marcus-type thermodynamic-kinetic relationship was identified between the shape recovery rate, Rt( N), and the elastic modulus, E, in poly(isocyanurate-urethane) shape memory aerogels. The latter were prepared with mixtures of diethylene, triethylene, and tetraethylene glycol and an aliphatic triisocyanate. Synthetic conditions were selected using a statistical design of experiments method. Microstructures obtained in each formulation could be put into two groups, one consisting of micron-size particles connected with large necks and a second one classified as bicontinuous. The two types of microstructures could be explained consistently by spinodal decomposition involving early versus late phase separation relative to the gel point. Irrespective of microstructure, all samples showed a shape memory effect with shape fixity and shape recovery ratios close to 100%. Larger variations (0.35-0.71) in the overall figure of merit, the fill factor, were traced to a variability in the shape recovery rates, Rt( N), which in turn were related to the microstructure. Materials with bicontinuous microstructures were stiffer and showed slower recovery rates. Thereby, using the elastic modulus, E, as a proxy for microstructure, the correlation of Rt( N) with E was traced to a relationship between the activation barrier for shape recovery, Δ A#, and the specific energy of deformation, (reorganization energy, λ), which in turn is proportional to the elastic modulus. Data were fitted well ( R2 = 0.92) by the derived equations. The inverse correlation between Rt( N) and the elastic modulus, E, provides a means for qualitative predictability of the shape recovery rates, the fill factors, and the overall quality of the shape memory effect.

6.
RSC Adv ; 8(38): 21214-21223, 2018 Jun 08.
Article in English | MEDLINE | ID: mdl-35539905

ABSTRACT

Scalable, low-density and flexible aerogels offer a unique combination of excellent mechanical properties and scalable manufacturability. Herein, we report the fabrication of a family of low-density, ambient-dried and hydrophobic poly(isocyanurate-urethane) aerogels derived from a triisocyanate precursor. The bulk densities ranged from 0.28 to 0.37 g cm-3 with porosities above 70% v/v. The aerogels exhibit a highly stretchable behavior with a rapid increase in the Young's modulus with bulk density (slope of log-log plot > 6.0). In addition, the aerogels are very compressible (more than 80% compressive strain) with high shape recovery rate (more than 80% recovery in 30 s). Under tension even at high strains (e.g., more than 100% tensile strain), the aerogels at lower densities do not display a significant lateral contraction and have a Poisson's ratio of only 0.22. Under dynamic conditions, the properties (e.g., complex moduli and dynamic stress-strain curves) are highly frequency- and rate-dependent, particularly in the Hopkinson pressure bar experiment where in comparison with quasi-static compression results, the properties such as mechanical strength were three orders of magnitude stiffer. The attained outcome of this work supports a basis on the understanding of the fundamental mechanical behavior of a scalable organic aerogel with potential in engineering applications including damping, energy absorption, and substrates for flexible devices.

7.
ACS Appl Mater Interfaces ; 9(15): 13520-13536, 2017 Apr 19.
Article in English | MEDLINE | ID: mdl-28379692

ABSTRACT

Polymeric aerogels (PA-xx) were synthesized via room-temperature reaction of an aromatic triisocyanate (tris(4-isocyanatophenyl) methane) with pyromellitic acid. Using solid-state CPMAS 13C and 15N NMR, it was found that the skeletal framework of PA-xx was a statistical copolymer of polyamide, polyurea, polyimide, and of the primary condensation product of the two reactants, a carbamic-anhydride adduct. Stepwise pyrolytic decomposition of those components yielded carbon aerogels with both open and closed microporosity. The open micropore surface area increased from <15 m2 g-1 in PA-xx to 340 m2 g-1 in the carbons. Next, reactive etching at 1,000 °C with CO2 opened access to the closed pores and the micropore area increased by almost 4× to 1150 m2 g-1 (out of 1750 m2 g-1 of a total BET surface area). At 0 °C, etched carbon aerogels demonstrated a good balance of adsorption capacity for CO2 (up to 4.9 mmol g-1), and selectivity toward other gases (via Henry's law). The selectivity for CO2 versus H2 (up to 928:1) is suitable for precombustion fuel purification. Relevant to postcombustion CO2 capture and sequestration (CCS), the selectivity for CO2 versus N2 was in the 17:1 to 31:1 range. In addition to typical factors involved in gas sorption (kinetic diameters, quadrupole moments and polarizabilities of the adsorbates), it is also suggested that CO2 is preferentially engaged by surface pyridinic and pyridonic N on carbon (identified with XPS) in an energy-neutral surface reaction. Relatively high uptake of CH4 (2.16 mmol g-1 at 0 °C/1 bar) was attributed to its low polarizability, and that finding paves the way for further studies on adsorption of higher (i.e., more polarizable) hydrocarbons. Overall, high CO2 selectivities, in combination with attractive CO2 adsorption capacities, low monomer cost, and the innate physicochemical stability of carbon render the materials of this study reasonable candidates for further practical consideration.

8.
ACS Appl Mater Interfaces ; 6(9): 6872-82, 2014 May 14.
Article in English | MEDLINE | ID: mdl-24758407

ABSTRACT

Polyurea (PUA) develops H-bonding with water and is inherently hydrophilic. The water contact angle on smooth dense PUA derived from an aliphatic triisocyanate and water was measured at θ=69.1±0.2°. Nevertheless, texture-related superhydrophobic PUA aerogels (θ'=150.2°) were prepared from the same monomer in one step with no additives, templates, or surfactants via sol-gel polymerization carried out in polar, weakly H-bonding acetonitrile. Those materials display a unique nanostructure consisting of micrometer-size spheres distributed randomly and trapped in a nanofiber web of the same polymer. Morphostructurally, as well as in terms of their hydrophobic properties, those PUA aerogels are analogous to well-studied electrospun fiber mats incorporating particle-like defects. PUA aerogels have the advantage of easily scalable synthesis and low cost of the raw materials. Despite large contact angles and small contact areas, water droplets (5 µL) stick to the aerogels surface when the substrate is turned upside-down. That so-called Petal effect is traced to H-bonding at the points of contact between the water droplet and the apexes of the roughness of the aerogel surface. Monoliths are flexible and display oleophilicity in inverse order to their hydrophobicity; oil fills all the available open porosity (94% v/v) of cocoon-in-web like aerogels with bulk density ρb=0.073 g cm(-3); that capacity for oil absorption is >10:1 w/w and translates into ∼6:1 w/v relative to state-of-the-art materials (e.g., graphene-derived aerogels). Oil soaked monoliths float on water and can be harvested off.

9.
ACS Appl Mater Interfaces ; 6(7): 4891-902, 2014 Apr 09.
Article in English | MEDLINE | ID: mdl-24665966

ABSTRACT

Biocompatible dysprosia aerogels were synthesized from DyCl3·6H2O and were reinforced mechanically with a conformal nano-thin-polyurea coating applied over their skeletal framework. The random mesoporous space of dysprosia aerogels was filled up to about 30% v/v with paracetamol, indomethacin, or insulin, and the drug release rate was monitored spectrophotometrically in phosphate buffer (pH = 7.4) or 0.1 M aqueous HCl. The drug uptake and release study was conducted comparatively with polyurea-crosslinked random silica aerogels, as well as with as-prepared (native) and polyurea-crosslinked mesoporous silica perforated with ordered 7 nm tubes in hexagonal packing. Drug uptake from random nanostructures (silica or dysprosia) was higher (30-35% w/w) and the release rate was slower (typically >20 h) relative to ordered silica (19-21% w/w, <1.5 h, respectively). Drug release data from dysprosia aerogels were fitted with a flux equation consisting of three additive terms that correspond to drug stored successively in three hierarchical pore sites on the skeletal framework. The high drug uptake and slow release from dysprosia aerogels, in combination with their low toxicity, strong paramagnetism, and the possibility for neutron activation render those materials attractive multifunctional vehicles for site-specific drug delivery.


Subject(s)
Acetaminophen/chemistry , Drug Carriers/chemistry , Drug Delivery Systems/instrumentation , Dysprosium/chemistry , Indomethacin/chemistry , Insulin/chemistry , Silicon Dioxide/chemistry , Acetaminophen/pharmacology , Blood Platelets/drug effects , Blood Platelets/physiology , Gels/chemistry , Humans , Indomethacin/pharmacology , Insulin/pharmacology , Porosity
10.
J Org Chem ; 78(17): 8297-304, 2013 Sep 06.
Article in English | MEDLINE | ID: mdl-23952066

ABSTRACT

(1)H NMR, ESI-MS, and DFT calculations with the M062X/6-31G* method show that, in water, the bistetrafluoroborate salt of N,N'-dimethyl-2,6-diaza-9,10-anthraquinonediium dication (DAAQ·2BF4(-)) exists in equilibrium with both its gem-diol and several aggregates (from dimers to at least octamers). With high concentrations of HCl (e.g., 1.2-1.5 M), all aggregates break up and the keto-to-gem-diol equilibrium is shifted quantitatively toward the quinone form. The same effect is observed with 1.5 mol equiv of cucurbit[7]uril, CB[7], with which all equilibria are shifted toward the quinone form, which undergoes slow exchange with the CB[7] cavity as both the free and the CB[7]-intercalated quinone (DAAQ@CB[7]) are observed simultaneously by (1)H NMR. The affinity of DAAQ for the CB[7] cavity (Keq = 4 × 10(6) M(-1)) is in the range found for tricyclic dyes (0.4-5.4 × 10(6) M(-1)), and among the highest observed to date. A computational comparative study of the corresponding CB[7] complex of the N,N'-dimethyl-4,4'-bipyridinium dication (N,N'-dimethyl viologen, MeV) suggests that the higher binding constant for intercalation of DAAQ may be partially attributed to a lesser distortion of CB[7] (required to maximize favorable nonbonding interactions) as a result of the flat geometry of DAAQ.


Subject(s)
Alcohols/chemistry , Bridged-Ring Compounds/chemistry , Imidazoles/chemistry , Isoquinolines/chemistry , Ketones/chemistry , Quinones/chemistry , Cations/chemistry , Crystallography, X-Ray , Models, Molecular , Molecular Structure , Quantum Theory
11.
J Org Chem ; 77(5): 2263-71, 2012 Mar 02.
Article in English | MEDLINE | ID: mdl-22339810

ABSTRACT

According to recent reports, supramolecular complexes of the pyrylium cation with cucurbit[x]urils (CB[x], x = 7, 8) show promising photoluminescence suitable for electroluminescent devices. In turn, photoluminescence seems to be related to the stereochemistry of the complexes; however, that has been controversial. Here, we report that in H(2)O, 2,6-disubsituted-4-phenyl pyryliums (Pylm) form dimers quantitatively (equilibrium constants >10(4) M(-1)), but they enter as such only in the larger CB[8]. In terms of orientation, (1)H NMR shows that Me-Pylm, Ph-Pylm, and t-Bu-Pylm insert their 4-phenyl groups in either the CB[7] or CB[8] cavity. The orientation of iPr-Pylm in the iPr-Pylm@CB[7] complex is similar. Experimental conclusions are supported by DFT calculations using the M062X functional and the 6-31G(d) basis set. In the case of (iPr-Pylm)(2)@CB[8], (1)H NMR of both the guest and the host indicates that both guests might enter CB[8] from the same side with their iPr groups in the cavity, but DFT calculations leave room for ambiguity. In addition to the size and hydrophobicity of the 2,6-substituents of the guests, as well as the size and flexibility of the hosts, theory reveals the importance of explicit solvation (H(2)O) and finite temperature effects (particularly for (1)H NMR shielding calculations) in the determination of the stereochemistry of those complexes.


Subject(s)
Bridged-Ring Compounds/chemistry , Imidazoles/chemistry , Pyrans/chemistry , Models, Molecular , Molecular Structure , Quantum Theory
12.
Org Lett ; 11(7): 1595-8, 2009 Apr 02.
Article in English | MEDLINE | ID: mdl-19271743

ABSTRACT

N-Substituted 4-benzoylpyridinium monocations form stable host-guest complexes with cucurbit[7]uril (CB[7]) in DMSO (K(eq) approximately 0.6-1.9 x 10(3) M(-1)). Observation of simultaneous reversible and quasi-reversible e-transfer processes from the free and intercalated quests, respectively, is attributed to the pre-e-transfer host-guest equilibrium. The standard rate constant for Me-BP@CB[7] (k(s) = 1.0 x 10(-4) cm.s(-1)) reflects e-transfer across 5.7 A, corresponding to the distance of the intercalated guest from the outmost perimeter of CB[7] (5.3 A).


Subject(s)
Bridged-Ring Compounds/chemistry , Imidazoles/chemistry , Pyridinium Compounds/chemistry , Cations , Electron Transport , Molecular Structure , Thermodynamics
13.
J Am Chem Soc ; 131(13): 4576-7, 2009 Apr 08.
Article in English | MEDLINE | ID: mdl-19292451

ABSTRACT

For many applications ranging from catalysis to sensors to energetic materials, it is desirable to produce intimate mixtures of nanoparticles. For instance, to improve the reaction rates of energetic materials, the oxidizing agent and the fuel need to be mixed as intimately as possible, ideally at the nanoscopic level. In this context, the acidity of a hydrated CuCl(2) solution reacting toward a network of CuO nanoparticles (a good oxidant) is used to induce one-pot cogelation of a nanostructured network of a resorcinol-formaldehyde resin (RF, the fuel). The resulting wet gels are dried to aerogels, and upon pyrolysis under Ar, the interpenetrating CuO/RF network undergoes a smelting reaction toward metallic Cu. Upon ignition in the open air, pure RF aerogels do not burn, while CuO/RF composites, even with substoichiometric CuO, sustain combustion, burning completely leaving only a solid residue of CuO whose role then has been that of a redox mediator through the smelting reaction.

14.
Org Lett ; 10(6): 1131-4, 2008 Mar 20.
Article in English | MEDLINE | ID: mdl-18278929

ABSTRACT

In water, N-methyl-4-(p-substituted benzoyl)pyridinium cations, BP-X, exist in equilibrium with their hydrated forms (gem-diols), whose concentrations depend on the para substituent (-X). In the presence of cucurbit[7]uril (CB[7]), the benzoyl group shows a preference for the CB[7] cavity, and the ketone to gem-diol equilibrium is shifted toward the keto form, meaning that the stabilization realized through hydrophobic interactions of the benzoyl group in the CB[7] cavity exceeds the hydrogen-bonding stabilization of the gem-diols in the aqueous environment.

16.
J Am Chem Soc ; 126(13): 4094-5, 2004 Apr 07.
Article in English | MEDLINE | ID: mdl-15053584

ABSTRACT

We report the redox properties of four star systems incorporating the 4-benzoyl-N-alkylpyridinium cation; the redox potential varies along the branches but remains constant at fixed radii. Bulk electrolysis shows that at a semi-infinite time scale all redox centers are electrochemically accessible. However, voltammetric analysis (cyclic voltammetry and differential pulse voltammetry) shows that only two of the three redox-active centers in the perimeter are electrochemically accessible during potential sweeps as slow as 20 mV s-1 and as fast as 10 V s-1. On the contrary, both redox centers along branches are accessible electrochemically within the same time frame. These results are explained in terms of slow through-space charge transfer and the globular 3-D folding of the molecules and are discussed in terms of their implications on the design of efficient redox functional dendrimers.

17.
Org Lett ; 4(23): 4113-6, 2002 Nov 14.
Article in English | MEDLINE | ID: mdl-12423099

ABSTRACT

The 3a,9a-diazaperylenium dication (1) was synthesized for the first time in two steps from p-phenylene diamine. Ab initio calculations show a twisted ground state with a 6.4 degrees tilt between the two quinolizinium building blocks. Dication 1 is photoluminescent in fluid solutions of H(2)O, CH(3)CN, and CH(3)NO(2), but not in rigid matrices of the same solvents. This phenomenon has been attributed to a geometric relaxation of the tilted ground state into an emitting planar lowest singlet excited state. [structure: see text]


Subject(s)
Aza Compounds/chemical synthesis , Cations, Divalent , Perylene/analogs & derivatives , Perylene/chemical synthesis , Indicators and Reagents , Models, Molecular , Molecular Conformation
18.
J Org Chem ; 67(21): 7501-10, 2002 Oct 18.
Article in English | MEDLINE | ID: mdl-12375985

ABSTRACT

In anhydrous CH3CN a series of nine 4-(4-substituted-benzoyl)-N-methylpyridinium cations (substituent: -OCH3, -CH3, -H, -SCH3, -Br, -Ctbd1;CH, -CHO, -NO2, and -(+)S(CH3)2) demonstrate two chemically reversible, well-separated one-electron (1-e) reductions in the same potential range as other main stream redox catalysts such as quinones and viologens. Hammett linear free energy plots yield excellent correlation between the E(1/2) values of both waves and the substituent constants sigma(p)(-)(X). The reaction constants for the two 1-e reductions are rho(1) = 2.60 and rho(2) = 3.31. The lower rho(1) value is associated with neutralization of the pyridinium ring, and the higher rho(2) value with the negative charge developing during the 2nd-e reduction. Structure-function correlations point to a purely inductive role for substitution in both 1-e reductions. The case of the 4-(4-nitrobenzoyl)-N-methylpyridinium cation is particularly noteworthy, because the 4-nitrobenzoyl moiety undergoes reduction before the 2nd reduction of the 4-benzoyl-N-methylpyridinium system. Correlation of the third wave of this compound with the 2nd-e reduction of the others yields sigma(p)(-NO)2*- = -0.97 +/- 0.02, thus placing the -NO2-* group among the strongest electron donors. Solvent deuterium isotope effects and maps of the electrostatic potential (via PM3 calculations) as a function of substitution support that 2-e reduced forms develop H-bonding with proton donors (e.g., CH3OH) via the O-atom. The average number of CH3OH molecules entering the H-bonding association increases with e-donating substituents. H-bonding shifts the 2nd reduction wave closer to the first one. This has important practical implications, because it increases the equilibrium concentration of the 2-e reduced form from disproportionation of the 1-e reduced form.


Subject(s)
Pyridinium Compounds/chemistry , Pyridinium Compounds/chemical synthesis , Cations , Electrochemistry/methods , Electrons , Hydrogen Bonding , Oxidation-Reduction , Thermodynamics
SELECTION OF CITATIONS
SEARCH DETAIL
...