Add to Mendeley

Dichloroethane, 1,2-

I. Syed, S.D. Ray, in Encyclopedia of Toxicology (Third Edition), 2014

Environmental Fate and Behavior

1,2-Dichloroethane can enter the environment when it is made, packaged, shipped, or used. Most 1,2-dichloroethane is released to the air, although some is released to rivers or lakes. 1,2-Dichloroethane could also enter soil, water, or air in large amounts in an accidental spill (evaporates into the air very fast from soil and water). If released to air, a vapor pressure of 78.9 mm Hg at 25 °C indicates that 1,2-dichloroethane will exist solely as a vapor in the ambient atmosphere. Vapor-phase 1,2-dichloroethane will be degraded in the atmosphere by reaction with photochemically produced hydroxyl radicals; the half-life for this reaction in air is estimated to be 63 days. Indirect evidence for photooxidation of 1,2-dichloroethane comes from the observation that monitoring levels are highest during the night and early morning. It may also be removed from air in rain or snow. Since it stays in the air for a while, the wind may carry it over large distances. In water, 1,2-dichloroethane breaks down very slowly and most of it will evaporate to the air. Only very small amounts are taken up by plants and fish. Exact longevity of 1,2-dichloroethane in water remains unknown. It is thought that it remains longer in lakes than in rivers.

In soil, it either evaporates into the air or travels down through soil and enters underground water. 1,2-Dichloroethane has been found in the United States drinking water at levels ranging from 0.05 to 64 ppb. An average amount of 175 ppb has been found in 12% of the surface water and groundwater samples taken at 2783 hazardous wastes sites. 1,2-Dichloroethane has also been found in the air near urban areas at levels of 0.10–1.50 ppb and near hazardous waste sites at levels of 0.01–0.003 ppb. Small amounts of 1,2-dichloroethane have also been found in foods. Small organisms living in soil and groundwater may transform it into other, less harmful compounds, although this happens slowly. Large amounts from an accident, hazardous waste site, or landfill may likely reach the underground and contaminate drinking water wells. Biodegradation occurs slowly in water and soil surfaces. It is not expected to undergo hydrolysis and photolysis. Humans may get exposed to very low levels of 1,2-dichloroethane through its use as a gasoline additive (leaded gasoline is no longer used in the United States).

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780123864543012471

Renal Toxicology

H.M. Mehendale, in Comprehensive Toxicology, 2010

7.19.2.3 1,2-Dichloroethane (Ethylene Dichloride)

1,2-Dichloroethane is one of the most abundant synthetic chemicals in the United States and is used mainly as an intermediate in the production of vinyl chloride. Human exposure to 1,2-dichloroethane occurs mostly via inhalation of vapors at industrial sites. However, drinking water may also be a source of human exposure (Morgan et al. 1990). 1,2-Dichloroethane intoxication in humans can produce acute renal failure that is associated with oliguria, albuminuria, and increased blood urea nitrogen concentrations. The renal lesion, characterized by accumulation of fat and degeneration of tubular epithelium, appears to be selectively localized to the proximal tubular epithelium (Yodaiken and Babcock 1973). Single and repeated inhalation exposure of rats, guinea pigs, and monkeys to 1,2-dichloroethane at concentrations of 400–2400 ppm resulted in fatty degeneration, cast formation, and proliferation of the tubular epithelium (Spencer et al. 1951; Yodaiken and Babcock 1973). However, when 1,2-dichloroethane was given in drinking water at 8000 ppm for 13 weeks to both sexes of F344/N rats, Sprague–Dawley rats, and Osborne–Mendel rats, only female F344/N rats exhibited renal lesions (Morgan et al. 1990). These results suggest that animals may be more susceptible to 1,2-dichloroethane-induced nephrotoxicity after inhalation exposure compared with exposure by drinking contaminated water. While the above results also suggest that female F344/N rats are unusually sensitive to 1,2-dichloroethane nephrotoxicity when administered orally, the biochemical basis for this sensitivity remains unclear. While in vitro studies with 1,2-dichloropropane, a nephrotoxic analogue of 1,2-dichloroethane, show that renal cortical slices from male Wistar rats are more sensitive to 1,2-dichloropropane-induced toxicity than females, possibly because of higher expression of CYP2E1 in the male rat kidney (Odinecs et al. 1995), the effect of gender on 1,2-dichloropropane in vivo nephrotoxicity has not been investigated.

1,2-Dichloroethane bioactivation can occur via two major metabolic pathways catalyzed by CYPs and glutathione S-transferases (Figures 3 and 4) (Elfarra et al. 1985; Guengerich et al. 1980; Jean and Reed 1992; Marchand and Reed 1989; Reed and Foureman 1986; Schasteen and Reed 1983; Webb et al. 1987). 1,2-Dichloroethane is a known substrate for human CYP2E1 (Gonzalez and Gelboin 1994).

Figure 3. Oxidative metabolism of 1,2-dichloroethane (1) to 1,2-dichloroethanol (2), 2-chloroacetaldehyde (3), and 2-chloroacetic acid (4) by cytochrome P450s (CYPs); 2-(S-glutathionyl) acetaldehyde (5); 2-hydroxyethylglutathione (6).

Figure 4. Glutathione S-transferase-dependent metabolism of 1,2-dichloroethane (1) to S-(2-chloroethyl)glutathione (2); episulfonium ion derived from S-(2-chloroethyl)glutathione (3); ethylene (4); S-(2-hydroxyethyl)glutathione (5); S,S-(1,2-ethanediyl)bisglutathione (6); S-(2-hydroxyethyl)-l-cysteine (7).

CYP-mediated oxidation at one of the carbon atoms forms a highly unstable gem-chlorohydrin that spontaneously decomposes to yield the electrophilic metabolite, 2-chloroacetaldehyde. 2-Chloroacetaldehyde may react with GSH to yield 2-(S-glutathionyl) acetaldehyde, which can then be reduced by alcohol dehydrogenase to yield S-(2-hydroxyethyl)glutathione. 2-Chloroacetaldehyde can also be oxidized by aldehyde dehydrogenase to chloroacetic acid, a major urinary metabolite of 1,2-dichloroethane. The alternative metabolic pathway involves GSH conjugation by cytosolic glutathione S-transferases to yield S-(2-chloroethyl)glutathione, half-sulfur mustard, which may rearrange to an electrophilic episulfonium ion. This episulfonium ion can either be hydrolyzed to yield S-(2-hydroxyethyl)glutathione or react with GSH to yield S,S′-(1,2-ethanediyl)bisglutathione or ethylene. An in vitro study in rat hepatocytes suggests that glutathione S-transferase-catalyzed reactions are responsible for the majority of GSH depletion induced by 1,2-dichloroethane, 1,2-dibromoethane, and 1-bromo-2-chloroethane (Jean and Reed 1992). Because chloroacetaldehyde and S-(2-chloroethyl)glutathione, and its putative metabolite S-(2-chloroethyl) cysteine, have been shown to cause nephrotoxicity in vivo and cytotoxicity in vitro, and/or can act as alkylating agents (Elfarra et al. 1985; Erve et al. 1995; Guo et al. 1990; Kramer et al. 1987; Meyer et al. 1994; Sood and O’Brien 1994; Zamlauski-Tucker et al. 1994), the formation of these metabolites is implicated in 1,2-dichloroethane-induced nephrotoxicity.

1,2-Dichloroethane metabolism by glutathione S-transferases and CYP to yield nephrotoxic metabolites may occur in both liver and kidney. The relative roles of hepatic and renal metabolism by either pathway are likely to be species-, sex-, and age-dependent (Gonzalez and Gelboin 1994; Hinchman et al. 1991; Hu et al. 1993; Rozell et al. 1993; Singhal et al. 1992; Vos and Van Bladeren 1990). For S-(2-chloroethyl)glutathione and S-(2-chloroethyl) cysteine, renal transport mechanisms may play a major role in determining the target organ of toxicity caused by these S-conjugates. Thus, conjugates formed in the liver are likely to be translocated via the systemic circulation and/or the enterohepatic circulation to the kidney, where they can be actively transported and concentrated within the tubular epithelium (Dantzler et al. 1995; Elfarra et al. 1985; Guo et al. 1990; Kramer et al. 1987; Lash and Anders 1989; Monks and Lau 1987; Schaeffer and Stevens 1987; Silbernagl and Heuner 1990).

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780080468846008241

Structure of Organic Compounds

Robert J. Ouellette, J. David Rawn, in Principles of Organic Chemistry, 2015

1.10 Isomers

Compounds that have the same molecular formula but different structures are isomers. Structure refers to the linkage of the atoms. As we examine the structure of organic compounds in increasing detail, you will learn how subtle structural differences in isomers affect the physical and chemical properties of compounds. There are several types of isomers. Isomers that differ in their bonded connectivity are skeletal isomers. Consider the structural differences in the two isomers of C4H10, butane and isobutane. Butane has an uninterrupted chain of four carbon atoms (Figure 1.10), but isobutane has only three carbon atoms connected in sequence and a fourth carbon atom appended to the chain. The boiling points (bp) of butane and isobutane are −1°C and −12°C, respectively; the chemical properties of the two compounds are similar but different.

Figure 1.10. Structure of Isomers

Isomers that have different functional groups are functional group isomers. The molecular formula for both ethyl alcohol and dimethyl ether is C2H6O (Figure 1.10). Although the compositions of the two compounds are identical, their functional groups differ. The atomic sequence is C—C—O in ethyl alcohol, and the oxygen atom is present as an alcohol. The C—O—C sequence in the isomer corresponds to an ether.

The physical properties of these two functional group isomers, as exemplified by their boiling points, are very different. These substances also have different chemical properties because their functional groups differ.

Positional isomers are compounds that have the same functional groups in different positions on the carbon skeleton. For example, the isomeric alcohols 1-propanol and 2-propanol differ in the location of the hydroxyl group. The chemical properties of these two compounds are similar because they both contain the same type of functional group and have identical molecular weights.

Isomerism is not always immediately obvious. Sometimes two structures appear to be isomers when in fact the structures are the same compound written in slightly different ways. It is important to be able to recognize isomers and distinguish them from equivalent representations of the same compound. For example, 1,2-dichloroethane can be written in several ways. In each formula, the bonding sequence is Cl—C—C—Cl.

The isomer of 1,2-dichloroethane is 1,1-dichloroethane. In 1,1-dichloroethane, the two chlorine atoms are bonded to the same carbon atom, but in 1,2-dichloroethane, the two chlorine atoms are bonded to different carbon atoms. The different condensed structural formulas, CHCl2CH3 and CH2ClCH2Cl, also tell us that in the first case two chlorine atoms are bound to the same carbon and that in the second case the two chlorine atoms are bound to adjacent carbons.

Problem 1.9

The structural formulas for two compounds used as general anesthetics are shown below. Are they isomers? How do they differ?

Solution

The atomic compositions of these structural formulas are identical; the molecular formula is C3H2F5ClO. Therefore, the compounds are isomers. The carbon skeletons are identical and the compounds are both ethers.

Both isomers have a CHF2 unit on the right side of the ether oxygen atom in spite of the different ways in which the fluorine and hydrogen are written—this is not the basis for isomerism. The two-carbon unit on the left of the oxygen atom has the halogen atoms distributed in two different ways. That is, they are positional isomers. The structure on the left has two fluorine atoms bonded to the carbon atom bonded to the oxygen atom. The carbon atom on the left has a fluorine and a chlorine atom bonded to it. The structure on the right has one chlorine atom bonded to the carbon atom appended to the oxygen atom. The carbon atom on the left has three fluorine atoms bonded to it.

Problem 1.10

Compare the following structures of two intermediates in the metabolism of glucose. Are they isomers? How do they differ?

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B978012802444700001X

The chemistry of thieno[b]pyrrolones, dihydrothieno[b]pyrrolones, and their fused derivatives

Andrew D. Harper, R. Alan Aitken, in Advances in Heterocyclic Chemistry, 2020

6.3.2.3.5 By electrophilic aromatic substitution of thiophenes

When heated in DCE with aluminum bromide, the N-aryl-2-bromo-N-(thiophen-3-yl)acetamide derivatives 401 and 402 undergo intramolecular Friedel–Crafts alkylation reactions to give the 4-aryl-4,6-dihydro-5H-thieno[3,2-b]pyrrol-5-ones 305 and 306, respectively—in the case of substrate 402, which has the more electron-rich N-aryl group, the 1-(thiophen-3-yl)indolin-2-one 404 was also produced in 18% yield. When the 2-bromo-N-(4-methoxyphenyl)-N-(thiophen-3-yl)acetamide derivative 403 was subjected to these conditions, cleavage of the methyl ether also occurred, resulting in a mixture of phenolic products 307 and 405 being obtained (Scheme 113) (2003RCB1873).

Scheme 113.

In the presence of catalytic quantities of triethylamine, the 3-aminothiophenes 406 and 407 react with chloroacetyl chloride to give the 4,6-dihydro-5H-thieno[3,2-b]pyrrol-5-ones 333 (2008PS(183)1679) and 408 (2011MI360), respectively (Scheme 114).

Scheme 114.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/S0065272519300534

Part I: Functional Groups and Their Properties

Robert J. Ouellette, J. David Rawn, in Organic Chemistry, 2014

2.7 Isomers

Compounds that have the same molecular formula whose atoms are linked in different ways are called isomers. As we examine the structure of organic compounds in detail, we will find that subtle structural differences profoundly affect the physical and chemical properties of isomers.

We can divide isomers into two broad classes. Substances that differ in their connectivity are constitutional isomers. Isomers that have the same connectivity but differ in the arrangement of the atoms in space are stereoisomers. We will consider stereoisomers in greater detail in Chapter 8 and thereafter.

Constitutional isomers can differ in their carbon backbones. Consider the structural differences in the two isomers of C4H10, butane and isobutane. Butane has an uninterrupted chain of carbon atoms (Figure 2.5), but isobutane has only three carbon atoms connected in sequence. The fourth carbon atom is bonded to the chain as a "branch".

Figure 2.5. Structures of Constitutional Isomers

Constitutional isomers have the same molecular formulas, but they have different connectivities. n-Butane and isobutane are examples of constitutional isomers, as are ethanol and dimethyl ether.

Constitutional isomers can also have different functional groups. For example, both ethyl alcohol and dimethyl ether have the same molecular formula: C2H6O. Although the molecular formulas of the two compounds are identical, their functional groups differ (Figure 2.5). The atomic connectivity is C─C─O in ethyl alcohol and the oxygen atom is part of an alcohol. In contrast, the C─O─C connectivity in the isomer forms an ether.

Constitutional isomers can have the same functional groups, but they are located at different points on the carbon skeleton. For example, the isomers 1-propanol and 2-propanol have hydroxyl group on different carbon atoms.

Sometimes two structural formulas appear to be isomers, but represent the same compound. For example, 1,2-dichloroethane can be written in several ways. But, the bonding sequence is Cl─C-C─Cl in each formula, so all three structural formulas represent the same molecule.

The isomer of 1,2-dichloroethane is 1,1-dichloroethane. In 1,1-dichloroethane, the two chlorine atoms are bonded to the same carbon atom, but in 1,2-dichloroethane, the two chlorine atoms are bonded to different carbon atoms. The different condensed structural formulas, CHCl2CH3 and CH2ClCH2Cl, also tell us that in the first case two chlorine atoms are bound to the same carbon and that in the second case the two chlorine atoms are bound to adjacent carbons.

Problem 2.16

The structural formulas of two compounds used as general anesthetics are shown below. Are they isomers? How do they differ?

Problem 2.17

Compare the following structures of two intermediates in the metabolism of glucose. Are they isomers? How do they differ?

Nomenclature

As the number of atoms represented in a molecular formula increases, the number of isomers increases exponentially. Each isomer must be uniquely identified by a name. Nomenclature in organic chemistry is the systematic method of naming compounds. The method was devised at a meeting in Geneva, Switzerland, in 1892. Compounds are now named by rules developed by the International Union of Pure and Applied Chemistry (IUPAC). The rules generate a single definitive name for each compound. A universal, systematic method for naming organic compounds is essential to avoid confusion. In the past, different names have often been given to the same compound. For example, CH3CH2OH has been called alcohol, spirit, grain alcohol, ethyl alcohol, and ethanol. For a small molecule like ethane this variety of names presents no problem, but for larger molecules a systematic name is essential.

A chemical name consists of three parts: prefix, parent, and suffix. The parent indicates how many carbon atoms are in the main carbon backbone. The suffix identifies most of the functional groups present in the molecule, for example -ol for alcohols, -al for aldehydes, and -one for ketones. The prefix specifies the location of the functional group designated in the suffix and any other groups on the parent chain.

Once the rules are applied, there is only one name for each structure and one structure for each name. For example, the compound partly responsible for the odor of skunk is 3-methyl-1-butanethiol.

Butane is the parent name of the four-carbon unit written horizontally. The prefix "3-methyl" identifies and locates the CH3 written above the chain of carbon atoms. The prefix "l-" and the suffix "thiol" specify the position and identity of the sulfhydryl (SH) group. This method of assigning numbers to the carbon chain and other features of the IUPAC system will be discussed in greater detail in subsequent chapters.

In spite of the IUPAC system, many common names are so well established that both common and IUPAC names are accepted. The IUPAC name for CH3CH2OH is ethanol, but the common name ethyl alcohol is often used.

As we introduce the nomenclature of each class of organic compounds, we will see that determining a systematic name is straightforward when the rules are followed. In this text, we will often give common names within parentheses after the IUPAC name.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780128007808000024

The Integration of Process Design and Control

A.C. Dimian, C.S. Bildea, in Computer Aided Chemical Engineering, 2004

4.2.2 Plantwide control problem

The quality of the intermediate DCE must fulfil strict purity specifications. Low impurity levels imply high energetic consumption, but higher impurity amounts are not desired for operation. The intermediate DCE is conditioned mainly in the distillation column S2. In the bottom product the concentration of the two "bad impurities" I1 and I2 must not exceed an upper limit, of 100 and 600 ppm, respectively, while the concentration of the good impurity I3 must be kept around optimal value of 2000 ppm.

Because the key impurities are implied in all-three reaction systems through recycles that cross in the separation system, their inventory is a plantwide control problem. This is completed by technological and environmental constraints, as mentioned.

The advanced removal of I1 and I2 must find a compromise with an optimal concentration of I3 in the bottom product of column S2. It is worth to mention that these contradictory requirements cannot be fulfilled by any stand-alone design of S2. Thus, the effective control of impurities becomes possible only by exploiting the positive feedback effects of the recycle loops that are balanced by the negative feedback effects of chemical conversion and exit streams.

Hence, the plantwide control objective is the quality of DCE sent to the cracking section, for which three specifications regarding key impurities are required. These are the outputs of the plantwide control problem, available by direct concentration measurements, as by IR spectroscopy or on-line chromatography.

An analysis of the degrees of freedom indicates as first choice manipulated variables belonging to the column S2, used for quality control: D2 - distillate flow rate, SS2 - side stream flow rate, and Q2 - reboiler duty. We may also consider manipulated variables belonging to the column S4, adjacent and connected with S2 by a recycle, but dynamically much faster. Thus, supplementary inputs are: D4 - distillate flow rate, and Q4 - reboiler duty. Hence, the inputs are the variables D2, SS2, Q2, D4 and Q4.

A major disturbance of the material balance is simulated here by a step variation in the external feed (FDCE). A second significant disturbance is X13, the fraction of impurity I3 introduced by the external DCE feed. The most probable range of frequencies for disturbance rejection is 0.1-1 rad/h for throughput, and 0.1-10 rad/h for impurities.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/S1570794604800683

Scientific Bases for the Preparation of Heterogeneous Catalysts

J. Blanco, ... J.A. Martin, in Studies in Surface Science and Catalysis, 2006

2.3 Catalytic measurements

Catalytic combustion of SO2, toluene and 1,2-dichloroethane were conducted at atmospheric pressure in a tubular flow reactor with an inner diameter of 18 mm. Catalyst extrudates or monoliths were packed into the reactor with glass wool plugs at each end. The reactor was placed in a furnace equipped with a temperature control to maintain a constant reactor temperature, and two thermocouples to measure the inlet and outlet reactor temperatures. Gas compositions and flow rates were set by mass flow controllers. The following reaction conditions were used to test the catalytic activity of the 0.2 wt. % Pt supported samples:

toluene oxidation conditions: gas inlet: 250 ppm toluene in air, GHSV = 16000 h−1 (NTP), vL= 0.24 Nm s−1.

1,2-dichloroethane oxidation conditions: gas inlet: 250 ppm 1,2-dichloroethane in air, GHSV = 9500 h−1 (NTP), vL = 0.98 Nm s−1.

SO2 oxidation conditions: gas inlet: 1000 ppm SO2 and 7 % O2 with nitrogen balance, GHSV = 6000 h−1 (NTP), vL = 0.60 Nms−1.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/S0167299106809031

Volume 1

In Bretherick's Handbook of Reactive Chemical Hazards (Seventh Edition), 2007

4742. Dinitrogen tetraoxide (Nitrogen dioxide)

[10544-72-6]

N2O4

HCS 1980, 675 (cylinder)

The equilibrium mixture of nitrogen dioxide and dinitrogen tetraoxide is completely associated at —9°C to the latter form which is marginally endothermic (ΔH°f (g) +9.7 kJ/mol, 0.10 kJ/g). Above 140°C it is completely dissociated to nitrogen dioxide, which is moderately endothermic (ΔH°f (g) +33.8 kJ/mol, 0.74 kJ/g).

Acetonitrile, Indium

MRH Acetonitrile 7.87/25

Addison, C. C. et al., Chem. & Ind., 1958, 1004

Shaking a slow-reacting mixture caused detonation, attributed to indium-catalysed oxidation of acetonitrile.

Alcohols

Daniels, F., Chem. Eng. News, 1955, 33, 2372

A violent explosion occurred during the ready interaction to produce alkyl nitrates.

Ammonia

MRH 6.61/33

Mellor, 1940, Vol. 8, 541

Liquid ammonia reacts explosively with the solid tetraoxide at —80°C, while aqueous ammonia reacts vigorously with the gas at ambient temperature.

Barium oxide

Mellor, 1940, Vol. 8, 545

In contact with the gas at 200°C the oxide suddenly reacts, reaches red heat and melts.

Boron trichloride

Mellor, 1946, Vol. 5, 132

Interaction is energetic.

Carbon disulfide

1.

Mellor, 1940, Vol. 8, 543

2.

Sorbe, 1968, 132

Liquid mixtures proposed for use as explosives are stable up to 200°C [1], but can be detonated by mercury fulminate, and the vapours by sparking [2].

Carbonylmetals

Cloyd, 1965, 74

Combination is hypergolic.

Cellulose, Magnesium perchlorate

See Magnesium perchlorate: Cellulose, etc.

Cycloalkenes, Oxygen

Lachowicz, D. R. et al., US Pat. 3 621 050, 1971

Contact of cycloalkenes with a mixture of dinitrogen tetraoxide and excess oxygen at temperatures of 0°C or below produces nitroperoxonitrates of the general formula — CHNO2—CH(OONO2)—which appear to be unstable at temperatures above 0°C, owing to the presence of the peroxonitrate group.

See Hydrocarbons, below

Difluorotrifluoromethylphosphine

Mahler, W., Inorg. Chem., 1979, 18, 352

A reaction, to produce the phosphine oxide on 12 mmol scale, ignited.

Dimethyl sulfoxide

MRH 6.99/36

See Dimethyl sulfoxide: Dinitrogen tetraoxide

Formaldehyde

1.

Pollard, F. H. et al., Trans. Faraday Soc., 1949, 45, 767—770

2.

Rastogi, R. P. et al., Chem. Abs., 1975, 83, 12936

The slow (redox) reaction becomes explosive around 180°C [1], or even lower [2].

See other REDOX REACTIONS

Halocarbons

MRH Chloroform 2.38/67, 1,2-dichloroethane 5.06/42, 1,1-dichloroethylene 5.06/46, trichloroethylene 3.97/56

1.

Turley, R. E., Chem. Eng. News, 1964, 42(47), 53

2.

Benson, S. W., Chem. Eng. News, 1964, 42(51), 4

3.

Shanley, E. S., Chem. Eng. News, 1964, 42(52), 5

4.

Kuchta, J. M. et al., J. Chem. Eng. Data, 1968, 13, 421—428

Mixtures of the tetraoxide with dichloromethane, chloroform, carbon tetrachloride, 1,2-dichloroethane, trichloroethylene and tetrachloroethylene are explosive when subjected to shock of 25 g TNT equivalent or less [1]. Mixtures with trichloroethylene react violently on heating to 150°C [2]. Partially fluorinated chloroalkanes were more stable to shock. Theoretical aspects are discussed in the later reference [2,3]. The effect of pressure on flammability limits has been studied [4].

See Uranium: Nitric acid

See Vinyl chloride: Oxides of nitrogen

Heterocyclic bases

MRH Pyridine 7.82/22, quinoline 7.87/22

Mellor, 1940, Vol. 8, 543

Pyridine and quinoline are attacked violently by the liquid oxide.

Hydrazine derivatives

1.

Cloyd, 1965, 74

2.

Miyajima, H. et al., Combust. Sci. Technol., 1973, 8, 199—200

Combinations with hydrazine, methylhydrazine, 1,1-dimethylhydrazine or mixtures thereof are hypergolic and used in rocketry [1]. The hypergolic gas-phase ignition of hydrazine at 70—160°C/53—120 mbar has been studied [2].

See ROCKET PROPELLANTS

Hydrocarbons

MRH values below references

1.

Mellor, 1967, Vol. 8, Suppl. 2.2, 264

2.

Fierz, H. E., J. Soc. Chem. Ind., 1922, 41, 114R

3.

Raschig, F., Z. Angew. Chem., 1922, 35, 117—119

4.

Berl, E. Z. Angew. Chem., 1923, 36, 87—91

5.

Schaarschmidt, A., Z. Angew. Chem., 1923, 36, 533—536

6.

Berl, E., Z. Angew. Chem., 1924, 37, 164—165

7.

Schaarschmidt, A., Z. Angew. Chem., 1925, 38, 537—541

8.

MCA Case History No. 128

9.

Folecki, J. et al., Chem. & Ind., 1967, 1424

10.

Cloyd, 1965, 74

11.

Urbanski, 1967, Vol. 3, 289

12.

Biasutti, 1981, 50

13.

Biasutti, 1981, 53—54

MRH Benzene 7.99/19, hexane 7.91/17, isoprene 8.28/18, methylcyclohexane 7.87/17

A mixture of the tetraoxide and toluene exploded, possibly initiated by unsaturated impurities [1]. During attempted separation by low temperature distillation of an accidental mixture of light petroleum and the oxide, a large bulk of material awaiting distillation became heated by unusual climatic conditions to 50°C and exploded violently [2]. Subsequently, discussion of possible alternative causes involving either unsaturated or aromatic compounds was published [3,4,5,6,7]. Erroneous addition of liquid in place of gaseous nitrogen tetraoxide to hot cyclohexane caused an explosion [8]. During kinetic studies, one sample of a 1:1 molar solution of tetraoxide in hexane exploded during (normally slow) decomposition at 28°C [9]. Cyclopentadiene is hypergolic with the oxide [10]. These incidents are understandable because of their similarity to rocket propellant systems and liquid mixtures previously used as bomb fillings [11]. The liquid oxide leaking from a ruptured 6 t storage tank ran into a gutter containing toluene and a violent explosion ensued [12]. An alternative account describes the hydrocarbon as benzene [13].

See Cycloalkenes, above; Unsaturated hydrocarbons, below

Hydrogen, Oxygen

Lewis, B., Chem. Rev., 1932, 10, 60

The presence of small amounts of the oxide in non-explosive mixtures of hydrogen and oxygen renders them explosive.

Isopropyl nitrite, Propyl nitrite

Safety in the Chemical Laboratory, Vol. 1, 121, Steere, N. V. (Ed.), Easton (Pa.) J. Ch. Ed., 1967

A pressurised mixture of the cold components exploded very violently during a combustion test run. The mixture was known to be autoexplosive at ambient temperature, and both of the organic components are capable of violent decomposition in absence of added oxidant.

Laboratory grease

Arapava, L. D. et al., Chem. Abs., 1985, 102, 169310

Contact of the lubricating grease Litol-24 with the oxidant at below 80°C led to explosion on subsequent impact. This involved nitration products of the antioxidant present, 4-hydroxydiphenylamine. Above 80°C decomposition superceded nitration, and no explosion occurred.

See other NITRATION INCIDENTS

Metal acetylides or carbides

MRH values show % of oxidant

Mellor, 1946, Vol. 5, 849

Caesium acetylide ignites at 100°C in the gas.

See Tungsten carbide: Nitrogen oxides

MRH 4.02/63

Ditungsten carbide: Oxidants

MRH 3.85/67

Metals

MRH Magnesium 12.97/50, potassium 3.72/46

1.

Mellor, 1940, Vol. 8, 544—545; 1942, Vol. 13, 342

2.

Pascal, 1956, Vol. 10, 382; 1958, Vol. 4, 291

Reduced iron, potassium and pyrophoric manganese all ignite in the gas at ambient temperature. Magnesium filings burn vigorously when heated in the gas [1]. Slightly warm sodium ignites in contact with the gas, and interaction with calcium is explosive [2].

See Aluminium: Oxidants

Nitroaniline

Anon., CISHC Chem. Safety Summ., 1978, 49, 3—4

Process errors led to discharge of copious amounts of nitrous fumes into the glass reinforced plastic ventilation duct above a diazotisation vessel. On two occasions fires were caused in the duct by vigorous reaction of the dinitrogen tetraoxide with nitroaniline dusts in the duct. Laboratory tests confirmed this to be the cause of the fires, and precautions are detailed.

Nitroaromatics

1.

Urbanski, 1967, Vol. 3, 288

2.

Kristoff, F. T. et al., J. Haz. Mat., 1983, 7, 199—210

Mixtures with nitrobenzene were formerly used as liquid high explosives, with addition of carbon disulfide to lower the freezing point, but high sensitivity to mechanical stimulus was disadvantageous [1]. During the recovery of acids from nitration of toluene, mixtures of the oxide with nitrotoluene or dinitrotoluene may be isolated under certain process conditions. While such mixtures are not unduly sensitive to impact, friction or thermal initiation, when oxygen-balanced they are extremely sensitive to induced shock and are capable of explosive propagation at film thicknesses below 0.5 mm. It is suspected that many explosions in TNT acid recovery operations, previously attributed to tetranitromethane, may have been caused by such mixtures [2].

Nitrogen trichloride

See Nitrogen trichloride: Initiators

Organic compounds

Riebsomer, J. L., Chem. Rev., 1945, 36, 158

In a review of the interaction of the oxidant with organic compounds, attention is drawn to the possibility of formation of unstable or explosive products.

Other reactants

Yoshida, 1980, 269

MRH values calculated for 18 combinations with oxidisable materials are given.

Ozone

See Ozone: Nitrogen oxide

Phospham

See Phospham: Oxidants

Phosphorus

MRH 9.12/35

See Phosphorus: Non-metal oxides

Sodium amide

Beck, G., Z. Anorg. Chem., 1937, 233, 158

Interaction with the oxide in carbon tetrachloride is vigorous, producing sparks.

Steel, Water

U.S. National Transportation Safety Board, Hazardous Materials Accident Brief,

Jan. 1998

A carbon steel tank for rail transportation of the tetroxide became contaminated with water, probably when a leaking valve, later replaced, was hosed down. After repair, the tank was charged with 50 tons of the oxide. This was later found to be wet, attempts were than made to empty the tanker. Acording to the single meter used to measure the transfer, this was accomplished (subsequent investigation suggested that only about 3 tons had been transferred because the dip pipes had corroded away). Water was charged to wash out the tank. The sequence of supposed emptying and washing was repeated and more water was added. It was noticed that pressure and fumes were excessive, atempts to deal with this continued some days. About a month after initial loading, and ten days after first washing, one of the heads blew off, throwing cladding about 100 m. Inspection of the remains showed several bands of corrosion, caused by nitric acid, produced from the oxide and water, reacting with steel to produce hydrogen and/or lower oxides of nitrogen which pressurised the weakened tank. Large tank cars are no longer used.

Tetracarbonylnickel

Bailar, 1973, Vol. 3, 1130

Interaction of the liquids is rather violent.

See Carbonylmetals, above

Tetramethyltin

Bailar, 1973, Vol. 2, 355

Interaction is explosively violent even at —80°C, and dilution with with inert solvents is required for moderation.

2-Toluidinium nitrate

Rastogi, R. P. et al., Indian J. Chem., Sect. A, 1980, 19A, 317—321

Reaction in this hybrid rocket propellant system is enhanced by presence of ammonium vanadate.

Triethylamine

Davenport, D. A. et al., J. Amer. Chem. Soc., 1953, 75, 4175

The complex, containing excess oxide over amine, exploded at below 0°C when free of solvent.

Triethylammonium nitrate

Addison, C. C. et al., Chem. & Ind., 1953, 1315

The two component form an addition complex with diethyl ether, which exploded violently after partial desiccation: an ether-free complex is also unstable.

See Triethylamine, above

Unsaturated hydrocarbons

MRH Isoprene 8.28/18

1.

Sergeev, G. P. et al., Chem. Abs., 1966, 65, 3659g

2.

Biasutti, 1981, 123

Dinitrogen tetraoxide reacts explosively between —32° and —90°C with propene, 1-butene, isobutene, 1,3-butadiene, cyclopentadiene and 1-hexene, but 6 other unsaturates failed to react [1]. Reaction of propene with the oxide at 2 bar/30°C to give lactic acid nitrate was proceeding in a pump-fed tubular reactor pilot plant. A violent explosion after several hours of steady operation was later ascribed to an overheated pump gland which recently had been tightened. A similar pump with a tight gland created a hot-spot at 200°C [2].

See Nitrogen dioxide: Alkenes

Vinyl chloride

See Vinyl chloride: Oxides of nitrogen

Xenon tetrafluoride oxide

Christe, K. O., Inorg. Chem., 1988, 27, 3764

In the reaction of the pentaoxide with xenon tetrafluoride oxide to give xenon difluoride dioxide and nitryl fluoride, the xenon tetrafluoride oxide must be used in excess to avoid formation of xenon trioxide, which forms a sensitive explosive mixture with xenon difluoride dioxide.

See Xenon tetrafluoride oxide: Caesium nitrate

See other ENDOTHERMIC COMPOUNDS, NON-METAL OXIDES, OXIDANTS

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780123725639500215

POLYMER CHARACTERIZATION BY FLUORESCENCE SPECTROSCOPY

JOHANNES W. HOFSTRAAT, in Applied Polymer Science: 21st Century, 2000

Preparation of Labeled Samples

The latex was loaded with Fluoroprobe using 1,2-dichloroethane as carrier. The Fluoroprobe has been added to the aqueous latex as an emulsion in water by making use of Aerosol MA-80 surfactant (sodium dihexylsulfosuccinate, from American Cyanamid). Jeffamine T403, a multifunctional crosslinker containing primary amines, was labeled with modified Fluoroprobe, containing a maleimide functionality. The reaction occurred readily in ethanol; 1 % of the amino-functionality on Jeffamine was labeled. The Jeffamine was subsequently used in a curing reaction with a polyacrylate dispersion, equipped with acetoacetate groups. Aminogroups on functionalized silica spheres were labeled In 0.1 M sodiumhydrogencarbonate at room temperature during 2 h. In heterogeneous labeling reactions the excess label can be easily removed by application of a filtration procedure. The silica spheres have been functionalized with primary aminogroups, using γ-aminopropyltrlethoxysilane [12]. Aldehyde functionalities on latices were labeled by reaction in water at 40 °C. A reaction time of 3 hours is sufficient for quantitative conversion of the functional groups. Excess label was removed by extensive washing with methanol. The methanol was checked for any residual probe by measuring the fluorescence. The nylons have been labeled in m-cresol. Reaction is complete after 4 h at 50 °C.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780080434179500416

Reduction

M. Zaidlewicz, ... M. Budny, in Comprehensive Organic Synthesis (Second Edition), 2014

8.24.1.6.2.3 Reaction media and catalyst recycling

Catalytic hydroboration is usually carried out in DCM, DCE, THF, benzene, and toluene, and at low temperatures in DME or DME/THF. Catalysts bearing fluorinated ligands exhibit high affinity for fluorous media, making possible the reaction under biphasic or monophasic conditions. On workup, the catalyst remains in the fluorous layer and can be recycled. Thus, Rh(PPh3)3Cl, soluble only slightly in CF3C6H5, has been transformed into ClRh(PCH2CH2C6F13)3 29, soluble in CF3C6H5 and CF3C6F11. Hydroboration of various aliphatic olefins with CatBH at 40 °C in a biphasic system toluene/CF3C6F11/29 affords the anti-Markovnikov product in 77–90% yield.376 The regioselectivity of the reaction with styrene and p-methoxystyrene is low, α/β 55:45 and 39:61, respectively. The catalyst has been recycled five times, TON 2411.377 The reaction of p-methoxystyrene with CatBH in supercritical carbon dioxide (scCO2), catalyzed by rhodium complexes with fluorinated phosphines, shows higher regioselectivity than with triphenylphosphine and is influenced by solvents378 (Table 10).

Table 10. Hydroboration of p-methoxystyrene with CatBH catalyzed by Rh complexes with fluorinated phosphines

Ligand Solvent Conversion (%) B L DB ArEt
No ligand scCO2 89 14 14 31 41
PPh3 scCO2 92 71 13 13 3
P(RF)3 scCO2 94 82 17 1
Cy2PRF THF 100 32 34 17 17
CF3C6F11 91 25 41 17 17
scCO2 100 100

Ar=MeOC6H4; coe=cyclooctene; scCO2 2800 psig; RF=CH2CH2C6F13.

The reaction carried out in THF or CF3C6F11 in the presence of P(RF)3 or Cy2PRF gives a mixture of the four products, whereas the α-substituted product is cleanly formed in scCO2/Cy2P(RF). Hydroboration of 2-vinylnaphthalene with PinBH catalyzed by [Rh(diphenylphosphinoethane (dppe))(cod)]BF4 in the ionic liquid [Bu4N]BF4/scCO2 is less selective, α/β 70:30.363

An alternative way of recycling the catalyst is to immobilize it on a solid support. A polymer-supported rhodium catalyst has been prepared by the copolymerization of ethylene glycol dimethyl acrylate and the p-isopropenylphenyl analog of [(nbd)Rh(dppe)]BF4 2% in the presence of dimethylformamide (DMF) used as a porogen.363 The polymer catalyst, 3 mol%, has been recycled three times in the hydroboration of styrene in THF, achieving 100% conversion and α/β 92:8. However, the combined crude materials from the three runs contain 4–5% of the Rh initially present in the polymeric catalyst.

In another approach, natural bentonites have been used as support of the asymmetric rhodium complex. The highest activities and enantioselectivities in the hydroboration of vinylarenes, competitive with the homogenous counterpart, have been obtained in the presence of (S)-QUINAP/Rh+X, immobilized in predried MK-10 when X=BF4, PF6, and SO3CF3. The catalyst can be recycled and its important feature is the activity remaining even when manipulated in air between the runs.379,380

The hydroboration of simple vinylarenes with CatBH catalyzed by rhodium complexes bound to the SiO2-immobilized CH2CHPPh2 group (Si-DPP) affords preferentially the branched product.381 In the presence of Rh(acac)(cod)2 and Si-DPP (two equivalents), p-methoxy and p-fluorostyrene are transformed into the branched products with 100% regioselectivity. Styrene and 2-vinylnaphthalene react with the same excellent selectivity. The immobilized catalyst can be recycled with no significant loss of activity or selectivity. However, the reaction lacks selectivity with more hindered vinylarenes or when PinBH is used instead of CatBH.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780080977423008260