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










Publication year range
1.
Eur J Orthop Surg Traumatol ; 34(1): 489-497, 2024 Jan.
Article in English | MEDLINE | ID: mdl-37632546

ABSTRACT

PURPOSE: The modified Kapandji technique has been proposed for fracture reduction in pediatric displaced distal radius fractures (DDRFs), but evidence is sparse. The purpose of this study was to evaluate our outcomes and complications, critically and systematically, when performing the modified Kapandji technique in pediatric DDRFs. Using this technique since 2011, we asked: (1) What is the quality of fracture reduction using this technique? (2) How stable is fracture alignment with this technique? (3) What are the postoperative complications and complication rates? METHODS: Retrospective observational study of 195 pediatric patients treated with the modified Kapandji technique. Quality of fracture reduction, fixation type (intrafocal, combined, or extrafocal), and coronal/sagittal angulation were recorded at surgery and healing. Perioperative complications were graded. Patients were stratified by fracture (metaphyseal or Salter-Harris) and fixation type, as well as age (≤ 6 years; 6 to 10 years; > 10 years). RESULTS: Fracture reduction was 'good' to 'anatomical' in 85% of patients. 'Anatomical' fracture reduction was less frequent in metaphyseal fractures (21% vs. 51%; p < .001). Mean angulation change was higher in metaphyseal fractures in both the sagittal (p = .011) and coronal (p = .021) planes. Metaphyseal fractures showed a higher mean change in sagittal angulation during fracture healing for the 'intrafocal' group. We observed a 15% overall complication rate with 1% being modified Sink Grade 3. CONCLUSION: The modified Kapandji technique for pediatric DDRFs is a safe and effective treatment option. Metaphyseal fractures that do not involve the physis should be treated with extrafocal or combined wire fixation. Complications that require additional surgical treatment are rare. LEVEL OF EVIDENCE: Level of evidence IV.


Subject(s)
Radius Fractures , Wrist Fractures , Humans , Child , Radius Fractures/diagnostic imaging , Radius Fractures/surgery , Radius Fractures/etiology , Fracture Fixation, Internal/adverse effects , Fracture Fixation, Internal/methods , Bone Wires , Fracture Fixation/methods
2.
Trauma Case Rep ; 46: 100866, 2023 Aug.
Article in English | MEDLINE | ID: mdl-37396120

ABSTRACT

Arthrofibrosis is defined as an excessive fibrotic tissue response within a joint leading to a painful loss of motion. This pathological scar formation process with dysregulated, inordinate extracellular matrix formation, especially collagen, may occur in any joints, although is frequently localized in the knee. Different etiologies have been described and most of them are related to trauma, infection or recent surgical procedure. Although arthrofibrosis affects people of all ages, it is unusual in pediatric population. We present a case report of an uncommon, foreign body induced knee arthrofibrosis in a 14-year old boy. We also review the current literature regarding diagnostic procedures and treatment rationale for arthrofibrosis of the knee.

4.
ACS Nano ; 4(6): 3374-80, 2010 Jun 22.
Article in English | MEDLINE | ID: mdl-20496882

ABSTRACT

Colloidal quantum dot (CQD) photovoltaics combine low-cost solution processability with quantum size-effect tunability to match absorption with the solar spectrum. Rapid recent advances in CQD photovoltaics have led to impressive 3.6% AM1.5 solar power conversion efficiencies. Two distinct device architectures and operating mechanisms have been advanced. The first-the Schottky device-was optimized and explained in terms of a depletion region driving electron-hole pair separation on the semiconductor side of a junction between an opaque low-work-function metal and a p-type CQD film. The second-the excitonic device-employed a CQD layer atop a transparent conductive oxide (TCO) and was explained in terms of diffusive exciton transport via energy transfer followed by exciton separation at the type-II heterointerface between the CQD film and the TCO. Here we fabricate CQD photovoltaic devices on TCOs and show that our devices rely on the establishment of a depletion region for field-driven charge transport and separation, and that they also exploit the large bandgap of the TCO to improve rectification and block undesired hole extraction. The resultant depleted-heterojunction solar cells provide a 5.1% AM1.5 power conversion efficiency. The devices employ infrared-bandgap size-effect-tuned PbS CQDs, enabling broadband harvesting of the solar spectrum. We report the highest open-circuit voltages observed in solid-state CQD solar cells to date, as well as fill factors approaching 60%, through the combination of efficient hole blocking (heterojunction) and very small minority carrier density (depletion) in the large-bandgap moiety.


Subject(s)
Electric Power Supplies , Electrodes , Nanotechnology/instrumentation , Quantum Dots , Solar Energy , Colloids/chemistry , Electric Conductivity , Energy Transfer , Equipment Design , Equipment Failure Analysis
5.
Chemistry ; 16(4): 1193-201, 2010 Jan 25.
Article in English | MEDLINE | ID: mdl-19998435

ABSTRACT

We have designed and synthesized highly efficient organic sensitizers with a planar thienothiophene-vinylene-thienothiophene linker. Under standard global AM 1.5 solar conditions, the JK-113-sensitized cell gave a short circuit photocurrent density (J(sc)) of 17.61 mA cm(-2), an open-circuit voltage (V(oc)) of 0.71 V, and a fill factor (FF) of 72%, corresponding to an overall conversion efficiency (eta) of 9.1%. The incident monochromatic photo-to-current conversion efficiency (IPCE) of JK-113 exceeds 80% over the spectral region from 400 to 640 nm, reaching its maximum of 93% at 475 nm. The band tails off toward 770 nm, contributing to the broad spectral light harvesting. Solar-cell devices based on the sensitizer JK-113 in conjunction with a volatile electrolyte and a solvent-free ionic liquid electrolyte gave high conversion efficiencies of 9.1% and 7.9%, respectively. The JK-113-based solar cell fabricated using a solvent-free ionic liquid electrolyte showed excellent stability under light soaking at 60 degrees C for 1000 h.

6.
Dalton Trans ; (26): 5137-47, 2009 Jul 14.
Article in English | MEDLINE | ID: mdl-19562174

ABSTRACT

Total halide abstraction from LnCl3 by Ag[Al(OC(CF3)3)4]/CH3CN has been confirmed for a series of lanthanide metal ions by the structural characterization of [Ln(CH3CN)n][Al(OC(CF3)3)4]3 (n=9, Ln3+=Nd, Eu, Gd, Dy; n=8, Ln3+=Tm) complexes. Evidence for the very low coordinating ability of the [Al(OC(CF3)3)4]- anion towards Ln3+ ions is provided in the solid state (X-ray, IR and Raman spectroscopy) and in anhydrous acetonitrile solution (conductivity, EPR and NMR measurements). In the solid state homoleptic nine-coordinated acetonitrile species are characteristic for lanthanides for the beginning (Nd) and the middle (Eu, Gd, Dy) of the Ln series, with a mono-capped square antiprismatic arrangement of the N donor atoms around the metal centres; while for those from the end of the series (Tm) eight-coordinated species are representative with a square antiprismatic arrangement. In anhydrous acetonitrile solution, conductivity measurements revealed 3:1 electrolyte types for all compounds. EPR and 19F NMR line broadening measurements attest lanthanide complexes free of any coordinating [Al(OC(CF3)3)4]- anion.

7.
Chemistry ; 15(8): 1966-76, 2009.
Article in English | MEDLINE | ID: mdl-19132696

ABSTRACT

In this study, we investigated the tetraalkylammonium salts of the weakly coordinating fluorinated alkoxyaluminates [pftb](-) ([Al(O(C(CF(3))(3))(4)](-)), [hfip](-) ([Al(OC(H)(CF(3))(2))(4)](-)) and [hftb](-) ([Al(OC(CH(3))(CF(3))(2))(4)](-)) in order to obtain information on their undisturbed spectral and structural properties, as well as to study their electrochemical behavior (i.e., conductivities in non-polar solvents and electrochemical windows). Several of the compounds qualify as ionic liquids with melting points as low as 42 degrees C for [NBu(4)](+)[hfip](-). Simple and almost quantitative metathesis reactions yielding these materials in high purity were developed. These [NR(4)](+) salts serve as model compounds for undisturbed anions and their vibrational spectra--together with simulated spectra based on quantum chemical DFT calculations--were used for the clear assignment of the anion bands. Besides, the ion volumes of the anions (V(ion)([pftb](-)) = 0.736 nm(3), V(ion)([hftb](-)) = 0.658 nm(3), V(ion)([hfip](-)) = 0.577 nm(3)) and their decomposition pathways in the mass spectrometric measurements have been established. The salts are highly soluble in non-polar solvents (up to 1.09 mol L(-1) are possible for [NBu(4)](+)[hftb](-) in CH(2)Cl(2) and 0.41 mol L(-1) for [NBu(4)](+)[hfip](-) in CHCl(3)) and show higher molar conductivities if compared to [NBu(4)](+)[PF(6)](-). The electrochemical windows of CH(2)Cl(2), CH(3)CN and 1,2-F(2)C(6)H(4) using the [NBu(4)](+) aluminate electrolytes are up to +0.5 V/-0.7 V larger than those using the standard [NBu(4)](+)[PF(6)](-).

8.
Dalton Trans ; (7): 946-56, 2008 Feb 21.
Article in English | MEDLINE | ID: mdl-18259629

ABSTRACT

The room-temperature stable CI3+ salts [CI3+[pftb](-)1 and [CI3]+[al-f-al](-) 2([pftb](-) = [Al(OC(CF3)3)4](-); [al-f-al](-) = [((CF3)3CO)3Al-F-Al(OC(CF3)3)3](-)) were prepared in quantitative yields from purified CI4 and the corresponding silver aluminates with total exclusion of light (NMR, IR, UV-VIS, X-ray diffraction). The isolated CI(3)(+) cation is trigonal planar with a sum of <(I-C-I) = 360.0 degrees (1) and 359.9 degrees (2). Attempts to prepare CHI2+ and CH2I+ salts from CHI3 or CH2I2/Ag[pftb] mixtures remained unsuccessful; the reaction with CH2I2 leads to the formation of the adduct [Ag(CH2I2)3]+[pftb](-)3, while for HCI3, dismutation with formation of 1 as well as 3 was observed. All particles were also calculated at the MP2/TZVPP level to predict the vibrational and electronic spectra as well as to calculate the Gibbs free energies of all reactions (DeltaG degrees , gas phase and CH2Cl2 solution). Quantum chemical calculations were also used to investigate the stability of the [pftb](-) anion against the electrophilic attack of the CX3+ and CHnX3-n+ cations (X = F-I, n = 1-3). The strength of the Lewis acidity of these cations and of the isoelectronic boron halides BX()and BHnX3-n have been established on the basis of their fluoride ion affinities (FIAs). The FIAs of the carbon and the boron containing compounds show opposite trends, with fluorinated halomethyl cations being stronger acids than their heavier congeners but iodinated holoboranes being stronger acids than their lighter homologues.

9.
Dalton Trans ; (46): 5376-86, 2007 Dec 14.
Article in English | MEDLINE | ID: mdl-18026586

ABSTRACT

What is the preferred coordination site of CI3+? Recent computational work suggested the iodine atoms of the Lewis acid CI3+ to be more electrophilic than the classically expected carbon atom, e.g. the complex with water is of type I2C-I...OH2+ and not the classically expected I3C-OH2+. If this structure is correct, one may also anticipate reactions of CI3+ as an I+ donor. Thus, we were interested in investigating the chemistry of CI3+ in the room-temperature stable salt [CI3]+[pftb](-) ([pftb](-) = [Al(OC(CF3)3)4]-) with weak nucleophiles that i) mimic water (OEt2) or ii) are electronically deactivated weak nucleophiles (PX3, X = Cl-I; AsI3). One question was: is it possible to obtain iodine-coordinated Lewis acid-base adducts of the CI3+ cation? With Et2O as a base, the cation behaves as a strong Lewis acid and cleaves the ether to give I3C-OEt, C2H4 and [H(Et2O)2]+. By contrast PX3 and AsI3 coordinate to the CI3+ cations and the adducts have classical, carbon-bound ethane-like structures, as proven by X-ray single-crystal diffraction, IR, UV-Vis and NMR spectroscopy. From variable temperature 13C NMR studies, it followed for the I3C-AsI3+ salt that the equilibrium between CI3+ and AsI3 is reversible and temperature dependent in solution. The I3C-PI3+ salt decomposes at room temperature giving PI4+ and C2I4, likely through an iodine coordinated I2C-I[dot dot dot]PI3+ intermediate. Thus CI3+ may also act as an I+ donor. All reactions are in agreement with ab initio quantum chemical calculations at the MP2/TZVPP level and assignments of experimental spectra were aided by quantum chemistry.

10.
J Phys Chem A ; 111(50): 13209-17, 2007 Dec 20.
Article in English | MEDLINE | ID: mdl-18044849

ABSTRACT

To compensate for lacking experimental standard enthalpies of formation DeltafH degrees of haloboranes/-boranates as well as the isoelectronic halocarbenium ions and halomethanes, high-level quantum chemical calculations up to the ccsd(t)/(SDB-)aug-cc-pVQZ level have been performed to establish these values. Very reliable experimental data (e.g., DeltafH degrees of HCl, F, and CH4) or at the G3 level established values (e.g., DeltafH degrees of CF3+=410 kJ mol(-1)) were used as anchor points to obtain accurate absolute DeltafH degrees and mean bond enthalpy (mBE) values. To further minimize systematic errors of the protocol, all derived quantities were assessed in isodesmic reactions at the G3 and ccsd(t) level using the (SDB-)aug-cc-PVTZ basis set. The obtained DeltafH degrees values are in very good agreement to (scarcely available) accurate experimental and computational data. Almost all B-containing compounds have been assessed for the first time. We derived "best" DeltafH degrees values and used them to determine the mean E-X bond enthalpies in H4-nEXn-/0 and H3-nEXn0/+ (with n=1-3, E=B, C, and X=F-I). In each of the series, the DeltafH degrees values increase from fluorine to iodine, and except for the iodine-containing carbenium ions and the bromo- and iodomethanes, the DeltafH degrees values become lower with the more halogen atoms that are present in the particle. The boron containing species always have a lower DeltafH degrees than the isoelectronic carbenium ions and methanes, and the H4-nEXn-/0 are lower in energy than the parent H3-nEXn+/0. This reflects the greater average B-X bond strengths.

11.
Chemistry ; 13(26): 7510-22, 2007.
Article in English | MEDLINE | ID: mdl-17616955

ABSTRACT

In an earlier publication (J. Am. Chem. Soc. 2002, 124, 7111) we showed that polymeric cationic [Ag(P(4)S(3))(n)](+) complexes (n=1, 2) are accessible if partnered with a suitable weakly coordinating counterion of the type [Al(OR(F))(4)](-) (OR(F): poly- or perfluorinated alkoxide). The present work addresses the following questions that could not be answered in the initial report: How many P(4)S(3) cages can be bound to a Ag(+) ion? Why are these complexes completely dynamic in solution in the (31)P NMR experiments? Can these dynamics be frozen out in a low-temperature (31)P MAS NMR experiment? What are the principal binding sites of the P(4)S(3) cage towards the Ag(+) ion? What are likely other isomers on the [Ag(P(4)S(3))(n)](+) potential energy surface? Counterion influence: Reactions of P(4)S(3) with Ag[Al{OC(CH(3))(CF(3))(2)}(4)] (Ag[hftb]) and Ag[{(CF(3))(3)CO}(3)Al-F-Al{OC(CF(3))(3))}(3)] (Ag[al-f-al]) gave [(P(4)S(3))Ag[hftb]](infinity) (7) as a molecular species, whereas [Ag(2)(P(4)S(3))(6)](2+)[al-f-al](-) (2) (8) is an isolated 2:1 salt. We suggest that a maximum of three P(4)S(3) cages may be bound on average to an Ag(+) ion. Only isolated dimeric dications are formed with the largest cation, but polymeric species are obtained with all other smaller aluminates. Thermodynamic Born-Haber cycles, DFT calculations, as well as solution NMR and ESI mass spectrometry indicate that 8 exhibits an equilibrium between the dication [Ag(2)(P(4)S(3))(6)](2+) (in the solid state) and two [Ag(P(4)S(3))(3)](+) monocations (in the gas phase and in solution). Dynamics: (31)P MAS NMR spectroscopy showed these solid adducts to be highly dynamic, to an extent that the (2)J(P,P) coupling within the cages could be resolved (J-res experiment). This is supported by DFT calculations, which show that the extended PES of [Ag(P(4)S(3))(n)](+) (n=1-3) and [Ag(2)(P(4)S(3))(2)](+) is very flat. The structures of alpha- and gamma-P(4)S(3) were redetermined. Their variable-temperature (31)P MAS NMR spectra are discussed jointly with those of all four currently known [Ag(P(4)S(3))(n)](+) adducts with n=1, 2, and 3.

12.
Chemistry ; 10(20): 5041-51, 2004 Oct 11.
Article in English | MEDLINE | ID: mdl-15372574

ABSTRACT

The unexpected but facile preparation of the silver salt of the least coordinating [(RO)3Al-F-Al(OR)3]- anion (R=C(CF3)3) by reaction of Ag[Al(OR)4] with one equivalent of PCl3 is described. The mechanism of the formation of Ag[(RO)3Al-F-Al(OR)3] is explained based on the available experimental data as well as on quantum chemical calculations with the inclusion of entropy and COSMO solvation enthalpies. The crystal structures of (RO)3Al<--OC4H8, Cs+[(RO)2(Me)Al-F-Al(Me)(OR)2]-, Ag(CH2Cl2)3+[(RO)3Al-F-Al(OR)3]- and Ag(eta2-P4)2+[(RO)3Al-F-Al(OR)3]- are described. From the collected data it will be shown that the [(RO)3Al-F-Al(OR)3]- anion is the least coordinating anion currently known. With respect to the fluoride ion affinity of two parent Lewis acids Al(OR)3 of 685 kJ mol(-1), the ligand affinity (441 kJ mol(-1)), the proton and copper decomposition reactions (-983 and -297 kJ mol(-1)) as well as HOMO level and HOMO-LUMO gap and in comparison with [Sb4F21]-, [Sb(OTeF5)6]-, [Al(OR)4]- as well as [B(R(F))4]- (R(F)=CF3 or C6F5) the [(RO)3Al-F-Al(OR)3]- anion is among the best weakly coordinating anions (WCAs) according to each value. In contrast to most of the other cited anions, the [(RO)3Al-F-Al(OR)3] anion is available by a simple preparation in conventional inorganic laboratories. The least coordinating character of this anion was employed to clarify the question of the ground state geometry of the Ag(eta2-P4)2+ cation (D(2h), D(2) or D(2d)?). In agreement with computational data and NMR spectra it could be shown that the rotation along the Ag-(P-P-centroid) vector has no barrier and that the structure adopted in the solid state depends on packing effects which lead to an almost D(2h) symmetric Ag(eta2-P4)2+ cation (0 to 10.6 degrees torsion) for the more symmetrical [Al(OR)4]- anion, but to a D2 symmetric Ag(eta2-P4)2+ cation with a 44 degrees twist angle of the two AgP2 planes for the less symmetrical [(RO)3Al-F-Al(OR)3]- anion. This implies that silver back bonding, suggested by quantum chemical population analyses to be of importance, is only weak.

13.
Chemistry ; 10(20): 5017-30, 2004 Oct 11.
Article in English | MEDLINE | ID: mdl-15372577

ABSTRACT

This article describes BP86/SV(P) (DFT) calculations on a representative set of weakly coordinating anions (WCAs) of type [M(L)n]-, their parent neutral Lewis acids M(L)(n-1) and their ate complexes with fluoride, that is, [FM(L)(n)](n-1) (M=B, L=F, OTeF5, C6H5, C6F5, C6H3(CF3)2, CF3; M=P, As, Sb, L=F, OTeF5; M=Al, L=OC(CF3)3). Compounds with fluoride bridges, that is, Sb(n)F(5n) and [Sb(n)F(5n+1)]- (n=2, 3, 4), Al2(L)5F and [(L)3Al-F-Al(L)3]- (L=OC(CF3)3), (F4C6[1,2-B(L)2]2, [F4C6[1,2-B(L)2]2F]-, [F4C6[1,2-B(L)2]2OMe]- (L=C6F5) were also calculated. Based on these BP86/SV(P) and auxiliary MP2/TZVPP, G2, and CBS-Q calculations the relative stabilities and coordinating abilities of these WCAs were established with regard to the fluoride ion affinities (FIA) of the parent Lewis acids, the ligand affinity (LA) of the WCAs, the decomposition of a given WCA in the presence of a hard (H+, proton decomposition PD) and a soft electrophile (Cu+, copper decomposition CuD), the position of the HOMO, the HOMO-LUMO gap, and population analyses of the anions providing partial charges for all atoms. To obtain data that is more reliable, the assessed quantities were calculated through isodesmic reactions. If parts of the calculations could not be done isodesmically, higher levels such as MP2/TZVPP, G2, and CBS-Q were used to obtain reliable values for these reactions. Although the obtained results can not be taken as absolute, the relative ordering of the stabilities of all WCAs will undoubtedly be correct, since a single methodology was chosen for the investigation. To include media effects the decomposition reactions of a subset of 14 WCAs with the SiMe3+ and [Cp2ZrMe]+ ions were also calculated in PhCl and 1,2-F2C6H4 (COSMO solvation model). We found that in most cases gas-phase calculations and solution calculations give comparable results for the stability of the anion. Applications of the LA and FIA that allow one to decide, on thermodynamic grounds, which WCA or Lewis acid is the most suitable for a given problem are sketched.

14.
J Am Chem Soc ; 126(24): 7571-7, 2004 Jun 23.
Article in English | MEDLINE | ID: mdl-15198604

ABSTRACT

Good quality ab initio calculations (MP2) show that the water adducts of BX(3) and CX(3)(+) have totally different structures (X = F-I). While all H(2)O-BX(3) complexes have classical C(s) symmetric structures with strong B-O bonds and additional H-bonding, the heavier CX(3)(+) cations (X = Cl-I) form weakly bonded "non-classical" water adducts that maximize C-X pi-bonding rather than C-O sigma-bonding. The delocalization of the positive charge as the driving force for pi-bond formation is absent in BX(3), and therefore, pi-bonding is only weak and not structure determining in H(2)O-BX(3). Since the PES of all H(2)O --> EX(3)(0/+1) particles (E = B, C) is very flat, flexible basis sets (like TZVPP) are required to rigorously characterize the adducts. In earlier calculations (J. Am. Chem. Soc. 1997, 119, 6648), classical structures were reported for all H(2)O --> EX(3)(0/+1) (E = B, C) complexes, likely resulting from the insufficient quality of the basis sets employed. By introducing a positive charge to three coordinate boron-halogen cations Do --> BX(2)(+) (Do = NH(3), OH(2), X-H), also the B-X bonds shrink due to the stronger pi-bonding induced by the positive charge delocalization and if compared to the respective neutral compounds like H(2)N-BX(2) or BX(3). The "non-classical" water adducts also suggest that the mechanism of organic reactions involving carbenium ion intermediates with alpha-bromine or -iodine substituents and a nucleophile may proceed through halogen- rather than carbon coordination.


Subject(s)
Boron Compounds/chemistry , Carbon/chemistry , Halogens/chemistry , Water/chemistry , Cations/chemistry , Hydrogen Bonding , Models, Molecular , Molecular Structure
15.
Angew Chem Int Ed Engl ; 43(16): 2066-90, 2004 Apr 13.
Article in English | MEDLINE | ID: mdl-15083452

ABSTRACT

Is there anything resembling a truly noncoordinating anion? Would it not be great to be able to prepare any crazy, beautiful, or simply useful cationic species that one has in mind, or has detected by mass spectroscopy? In condensed phases the target cation has to be partnered with a suitable counteranion. This is the moment when difficulties arise and many wonderful ideas end in the sink owing to coordination or decomposition of the anion. However, maybe these counteranion problems can be overcome by one of the new weakly coordinating anions (WCAs). Herein is an overview on the available candidates in the quest for the least coordinating anion and a summary of new applications, available starting materials, and general strategies to introduce a WCA into a system. Some of the unusual properties of WCA salts such as high solubility in low dielectric media, pseudo gas-phase conditions in condensed phases, and the stabilization of weakly bound and low-charged complexes are rationalized on thermodynamic grounds. Limits of the WCAs, that is, anion coordination and decomposition, are shown and a quantum chemical analysis of all types of WCAs is presented which allows the choice of a particular WCA to be based on quantitative data from a wide range of different anions.

17.
Chemistry ; 8(19): 4475-92, 2002 Oct 04.
Article in English | MEDLINE | ID: mdl-12355536

ABSTRACT

PX(4) (+)[Al(OR)(4)](-) (X=I: 1 a, X=Br: 1 b) was prepared from X(2), PX(3), and Ag[Al(OR)(4)] [R=C(CF(3))(3)] in CH(2)Cl(2) at -30 degrees C in 69-86 % yield. P(2)X(5) (+) salts were prepared from 2 PX(3) and Ag[Al(OR)(4)] in CH(2)Cl(2) at -30 degrees C yielding almost quantitatively P(2)X(5) (+)[Al(OR)(4)](-) (X=I: 3 a, X=Br: 3 b). The phosphorus-rich P(5)X(2) (+) salts arose from the reaction of cold (-78 degrees C) mixtures of PX(3), P(4), and Ag[Al(OR)(4)] giving P(5)X(2) (+)[Al(OR)(4)](-) (X=I: 4 a, X=Br: 4 b) with a C(2v)-symmetric P(5) cage. Silver salt metathesis presumably generated unstable PX(2) (+) cations from PX(3) and Ag[Al(OR)(4)] (X=Br, I) that acted as electrophilic carbene analogues and inserted into the Xbond;X (Pbond;X/Pbond;P) bond of X(2) (PX(3)/P(4)) leading to the highly electrophilic and CH(2)Cl(2)-soluble PX(4) (+) (P(2)X(5) (+)/P(5)X(2) (+)) salts. Reactions that aimed to synthesize P(2)I(3) (+) from P(2)I(4) and Ag[Al(OR)(4)] instead led to anion decomposition and the formation of P(2)I(5)(CS(2))(+)[(RO)(3)Al-F-Al(OR)(3)](-) (5). All salts were characterized by variable-temperature solution NMR studies (3 b also by (31)P MAS NMR), Raman and/or IR spectroscopy as well as X-ray crystallography (with the exception of 4 a). The thermochemical volumes of the Pbond;X cations are 121 (PBr(4) (+)), 161 (PI(4) (+)), 194 (P(2)Br(5) (+)), 271 (P(2)I(5) (+)), and 180 A(3) (P(5)Br(2) (+)). The observed reactions were fully accounted for by thermochemical calculations based on (RI-)MP2/TZVPP ab initio results and COSMO solvation enthalpy calculations (CH(2)Cl(2) solution). The enthalpies of formation of the gaseous Pbond;X cations were derived as +764 (PI(4) (+)), +617 (PBr(4) (+)), +749 (P(2)I(5) (+)), +579 (P(2)Br(5) (+)), +762 (P(5)I(2) (+)), and +705 kJ mol(-1) (P(5)Br(2) (+)). The insertion of the intermediately prepared carbene analogue PX(2) (+) cations into the respective bonds were calculated, at the (RI-)MP2/TZVPP level, to be exergonic at 298 K in CH(2)Cl(2) by Delta(r)G(CH(2)Cl(2))=-133.5 (PI(4) (+)), -183.9 (PBr(4) (+)), -106.5 (P(2)I(5) (+)), -81.5 (P(2)Br(5) (+)), -113.2 (P(5)I(2) (+)), and -114.5 kJ mol(-1) (P(5)Br(2) (+)).

SELECTION OF CITATIONS
SEARCH DETAIL
...