Welcome to LookChem.com Sign In|Join Free

CAS

  • or

1194-65-6

Post Buying Request

1194-65-6 Suppliers

Recommended suppliersmore

  • Product
  • FOB Price
  • Min.Order
  • Supply Ability
  • Supplier
  • Contact Supplier

1194-65-6 Usage

Chemical Properties

white powder

Uses

Different sources of media describe the Uses of 1194-65-6 differently. You can refer to the following data:
1. Herbicide.
2. Soil-applied herbicide used to control many annual and perennial broad-leaved weeds.
3. 2,6-dichlorobenzonitrile in which Q-amino-G- chlorobenzonitrile is used as starting material. And it is used for the synthesis of diclofenac sodium oxacillin and guanabenz acetate.

Definition

ChEBI: A nitrile that is benzonitrile which is substituted by chlorines at positions 2 and 6. A cellulose synthesis inhibitor, it is used as a pre-emergent and early post-emergent herbicide.

Synthesis Reference(s)

Synthetic Communications, 20, p. 2785, 1990 DOI: 10.1080/00397919008051490Synthesis, p. 943, 1992 DOI: 10.1055/s-1992-26271

General Description

2,6-Dichlorobenzonitrile is a white solid dissolved or suspended in a water-emulsifiable liquid carrier. The primary hazard is the threat to the environment. Immediate steps should be taken to limit spread to the environment. Can easily penetrate the soil and contaminate groundwater and nearby streams. Can cause illness by inhalation, skin absorption and/or ingestion. Used as a herbicide.

Air & Water Reactions

Not soluble in water.

Reactivity Profile

A halogenated nitrile. Nitriles may polymerize in the presence of metals and some metal compounds. They are incompatible with acids; mixing nitriles with strong oxidizing acids can lead to extremely violent reactions. Nitriles are generally incompatible with other oxidizing agents such as peroxides and epoxides. The combination of bases and nitriles can produce hydrogen cyanide. Nitriles are hydrolyzed in both aqueous acid and base to give carboxylic acids (or salts of carboxylic acids). These reactions generate heat. Peroxides convert nitriles to amides. Nitriles can react vigorously with reducing agents. Acetonitrile and propionitrile are soluble in water, but nitriles higher than propionitrile have low aqueous solubility. They are also insoluble in aqueous acids.

Health Hazard

SOLID: Harmful if swallowed.

Flammability and Explosibility

Nonflammable

Agricultural Uses

Herbicide: Dichlobenil is a herbicide used on cranberry bogs, dichondra, ornamentals, blackberry, raspberry, and blueberry fields, apple, pear, filbert and cherry orchards, vineyards, hybrid poplar-cottonwood plantations, and rights-of-way to control weeds; and sewers to remove roots. It acts on dandelion, prickly oxtongue (pre-emergence), and tree roots. Not approved for use in EU countries. Actively registered in the U.S.

Trade name

BARRIER?; BH Prefix D?; CARSORON?; CASORON? 133; CARSORON? G; CARSORON? G4; CARSORON? G20-SR; CODE H 133?; DECABANE?; DU-SPREX?; DYCLOMEC?; FYDULAN; FYDUMAS; FYDUSIT; H 133?; H 1313?; NIA 5996?; NIAGARA? 5006; NIAGARA 5,996; NOROSAC?; PREFIX D?

Environmental Fate

Biological. A cell suspension of Arthrobacter sp., isolated from a hydrosol, degraded dichlobenil to 2,6-dichlorobenzamide (71% yield) and several unidentified water soluble metabolites (Miyazaki et al., 1975). This microorganism was capable of rapidly degrading dichlobenil in aerobic sediment-water suspensions and in enrichment cultures (Miyazaki et al., 1975). Soil. The major soil metabolite is 2,6-dichlorobenzamide which undergoes further degradation to form 2,6-dichlorobenzoic acid. The estimated half-lives ranged from 1 to 12 months (Hartley and Kidd, 1987). Under field conditions, dichlobenil persists from 2 to 12 months (Ashton and Monaco, 1991). The disappearance of dichlobenil from a hydrosol and pond water was primarily due to volatilization and biodegradation. The time required for 50 and 90% dissipation of the herbicide from a hydrosol were approximately 20 and 50 days, respectively (Rice et al., 1974). Dichlobenil has a high vapor pressure and volatilization should be an important process. Williams and Eagle (1979) found that the half-life of dichlobenil was 4 weeks in soil 4–8 weeks after application. After 1 year following application, the half-life increased to 1 year. Plant. In plants, dichlobenil is transformed into glucose conjugates, insoluble residues and hydroxy products that are phytotoxic (Ashton and Monaco, 1991). These include three phytotoxic compounds, namely 2,6-dichlorobenzonitrile, 3-hydroxy-2,6-dichlorobenzonitrile and 4-hydroxy-2,6-dichlorobenzonitrile (Duke et al., 1991). Massini (1961) provided some evidence that dichlobenil is metabolized by plants. French dwarf beans, tomatoes, gherkin and oat plants were all exposed to a saturated atmosphere of dichlobenil at room temperature for 4 days. Most of the herbicide was absorbed and translocated by the plants in 3 days. After 6 days of exposure, bean seedlings were analyzed for residues using thinlayer plate chromatography. In addition to dichlobenil, another compound was found but it was not 2,6-dichlorobenzoic acid (Massini, 1963). Surface Water. The time required for 50 and 90% dissipation of the herbicide from New York pond water was approximately 21 and 60 days, respectively (Rice et al., 1974). Photolytic. When dichlobenil was irradiated in methanol with a 450-W mercury lamp and a Corex filter for 8 hours, o-chlorobenzonitrile and benzonitrile formed as the major and minor products, respectively (Plimmer, 1970). Chemical/Physical. Dichlobenil is hydrolyzed, especially in the presence of alkali, to 2,6-dichlorobenzamide (Briggs and Dawson, 1970; Worthing and Hance, 1991). Emits toxic fumes of nitrogen oxides and chlorine when heated to decomposition (Sax and Lewis, 1987).

Metabolic pathway

Twelve metabolites are isolated from either urine or bile from either rats (11 metabolites) or goats (seven metabolites) given single oral doses of 14C-labeled 2,6-dichlorobenzonitrile (DCBN). Five of these metabolites are also excreted in urine from rats dosed orally with 2,6-dichlorothiobenzamide (DCTBA) which is an acid amide analog. All metabolites from either DCBN or DCTBA are benzonitriles with the following ring substituents: Cl2, OH (three isomers); Cl2, (OH)2; Cl, (OH)2; Cl, OH, SH; Cl, OH, SCH3; SOCH3, OH; Cl2, S-(N-acetyl)cysteine; Cl, S-(N-acetyl)cysteine; Cl, OH, S-(N-acetyl)cysteine. The thiobenzamide moiety of DCTBA is converted to the nitrile in all extracted urinary metabolites. No hydrolysis of the nitrile in DCBN to either amide or an acid is detected. Urine is the major route for excretion; however, enterohepatic circulation occurs.

Purification Methods

Crystallise the nitrile from acetone. [Beilstein 9 IV 1006.]

Check Digit Verification of cas no

The CAS Registry Mumber 1194-65-6 includes 7 digits separated into 3 groups by hyphens. The first part of the number,starting from the left, has 4 digits, 1,1,9 and 4 respectively; the second part has 2 digits, 6 and 5 respectively.
Calculate Digit Verification of CAS Registry Number 1194-65:
(6*1)+(5*1)+(4*9)+(3*4)+(2*6)+(1*5)=76
76 % 10 = 6
So 1194-65-6 is a valid CAS Registry Number.
InChI:InChI=1/C7H3Cl2N/c8-6-2-1-3-7(9)5(6)4-10/h1-3H

1194-65-6 Well-known Company Product Price

  • Brand
  • (Code)Product description
  • CAS number
  • Packaging
  • Price
  • Detail
  • Alfa Aesar

  • (A14986)  2,6-Dichlorobenzonitrile, 98%   

  • 1194-65-6

  • 10g

  • 273.0CNY

  • Detail
  • Alfa Aesar

  • (A14986)  2,6-Dichlorobenzonitrile, 98%   

  • 1194-65-6

  • 25g

  • 337.0CNY

  • Detail
  • Alfa Aesar

  • (A14986)  2,6-Dichlorobenzonitrile, 98%   

  • 1194-65-6

  • 100g

  • 979.0CNY

  • Detail

1194-65-6SDS

SAFETY DATA SHEETS

According to Globally Harmonized System of Classification and Labelling of Chemicals (GHS) - Sixth revised edition

Version: 1.0

Creation Date: Aug 10, 2017

Revision Date: Aug 10, 2017

1.Identification

1.1 GHS Product identifier

Product name 2,6-dichlorobenzonitrile

1.2 Other means of identification

Product number -
Other names 2,6-Dichlorobenzonitrile

1.3 Recommended use of the chemical and restrictions on use

Identified uses For industry use only. Herbicide
Uses advised against no data available

1.4 Supplier's details

1.5 Emergency phone number

Emergency phone number -
Service hours Monday to Friday, 9am-5pm (Standard time zone: UTC/GMT +8 hours).

More Details:1194-65-6 SDS

1194-65-6Relevant articles and documents

Investigation of BiVO4 structure variations on the dichlorotoluene ammoxidation performance

Li, Xiongjian,Huang, Chi

, p. 866 - 870 (2021)

In this study, BiVO4 synthesized by hydrothermal and calcination methods were explored as catalysts in the ammoxidation of dichlorotoluenes to shed light on the structure-reactivity correlations. The BiVO4 samples were characterized by X-ray diffraction (XRD), temperature-programmed reduction (TPR), Brunauer–Emmett–Teller (BET), and UV–Vis spectrophotometry. The results showed that the catalytic activity of BiVO4 greatly relied on the structure variations. The hydrothermal prepared BiVO4 exhibited better catalytic activities as a consequence of its greater structure deformation, with maximum yields of 73.1, 72.2, and 70.8% for 3,4-, 2,4- and 2,6- dichlorobenzonitrles, respectively.

Hydrothermal Synthesis of Urchin-like W-V-O Nanostructures with Excellent Catalytic Performance

Li, Xiongjian,Sun, Li,Hu, Mingjie,Huang, Ronghua,Huang, Chi

, p. 14758 - 14763 (2018)

Urchinlike W-V-O microspheres have been successfully synthesized for the first time by a one-pot hydrothermal approach. The as-synthesized W-V-O material was characterized by several techniques such as XRD, SEM, TEM, FTIR, EDS, BET, and Raman spectroscopy. The characterization results have revealed that the W-V-O microspheres consist of numerous one-dimensional nanobelts radially grown from the center. The typical nanobelts display rectangular cross sections with lengths of several micrometers, widths of about 50 nm, and thicknesses of approximately 10-20 nm. Vanadium oxides are dispersed highly either on the external surface or inside the channel surface of the hexagonal WO3 structure. In addition, the as-obtained urchin-like W-V-O material was explored as a catalyst for the ammoxidation of 2,4- and 2,6-dichlorotoluene to the corresponding nitriles. The catalytic results have indicated that the W-V-O nanostructures show excellent performance with yields of 2,4- and 2,6-dichlorobenzonitrile respectively reaching up to 77.3 and 75.1%, which are the highest among the previously reported catalysts with two components. The formation process of the urchinlike W-V-O microspheres was simply investigated.

Effect of nitrogen-containing compounds on polychlorinated dibenzo-p-dioxin/dibenzofuran formation through de novo synthesis

Kuzuhara, Shunsuke,Sato, Hiroshi,Tsubouchi, Naoto,Ohtsuka, Yasuo,Kasai, Eiki

, p. 795 - 799 (2005)

An experimental study was conducted to clarify the suppression effect of nitrogen-containing compounds, that is, ammonia and urea, on the formation of polychlorinated dibenzo-p-dioxins (PCDDs) and polychlorinated dibenzofurans (PCDFs) through the de novo synthesis reaction. In the experiment, graphite and copper chloride contained in a mixture were used as sources of carbon and chlorine, respectively. The granulated sample mixture was charged as a packed-bed in the glass tube and heated at 300 °C in the flow of Ar-O 2 gas mixture. In some cases, urea was added as aqueous solution to the sample, while ammonia was added to the gas flowed through the sample bed. The amount of PCDD/Fs formed decreases significantly by the addition of both ammonia and urea. Particularly, the addition of urea reduces the amount of PCDD/Fs discharged in the outlet gas by approximately 90%. The oxidation rate of carbon in the early stage of the experiment, that is, the heating period, is promoted by the addition of nitrogen-containing compounds. However, soon after the temperature reaches 300 °C, the formation rate becomes lower than that of the case without the addition of nitrogen-containing compounds. On the other hand, organic compounds containing amino (-NH2) or cyanide (-CN) groups and those containing nitrogen within the carbon ring frame were detected in the outlet gas in the case of urea addition. Typically observed aromatic compounds are chlorobenzonitriles, chlorobenzeneamines, and chloropyridines. This suggests a possibility that hydrogen and/or chlorine combined with PCDD/Fs are also substituted by such nitrogen-containing groups, and this decreases the formation rate of their frame of carbon rings. This phenomenon was also consistent with the fact that a significant reduction was observed in the amount of PCDD/Fs released to the outlet gas when urea was added.

Ammoxidation of 2,6-dichloro toluene - From first trials to pilot plant studies

Martin,Kalevaru,Smejkal

, p. 275 - 279 (2010)

The scaling-up of the gas phase catalytic ammoxidation of 2,6-dichloro toluene (DCT) to 2,6-dichloro benzonitrile (DCBN) over a promoted vanadium phosphate (VPO) catalyst from first lab-scale experiments to pilot plant runs is reported. First experiments in a row of conversions of isomeric dichloro toluenes using simple, non-promoted VPO catalysts only show poor yield and selectivity. In particular, DCT ammoxidation is hindered due to bulky chlorine substituents probably preventing a sufficient interaction of the methyl group and lattice oxygen and/or N-containing surface species. Improved synthesis of VPO catalyst with the addition of promoters and γ-alumina or titania leads to significant increase in DCT conversion and DCBN yield. A Cr containing vanadyl pyrophosphate catalyst admixed with titania (anatase) showed conversion up to 97% with DCBN yields of ca. 80%. The same catalyst was also used for pilot plant runs, usually in the form of 5 mm × 3.5 mm shaped tablets that were prepared from a larger batch of solid synthesis. The scaling-up of the process using 100 ml of catalyst was investigated both by catalytic experiments and reactor simulations. The results showed that the temperature control will be crucial in scaling-up. Validation of simulation results with that of experimental results was also checked and a good agreement between measured and simulated results is observed.

Highly Modular Flow Cell for Electroorganic Synthesis

Gütz, Christoph,Stenglein, Andreas,Waldvogel, Siegfried R.

, p. 771 - 778 (2017)

A highly modular electrochemical flow cell and its application in electroorganic synthesis is reported. This innovative setup facilitates many aspects: an easy adjustment of electrode distance, quick exchange of electrode material, and the possibility to easily switch between a divided or undivided cell. However, the major benefit of the cell is the exact thermal positioning of the electrode material into a Teflon piece. Thereby, the application of expensive and nonmachinable electrode materials like boron-doped diamond or glassy carbon can easily be realized in flow cells. By this geometry, the maximum surface of such valuable electrode materials is exploited. The cell size can compete with classical preparative approaches in terms of performance and productivity. The optimization of reaction parameters and an easy up-scaling to larger flow cells is possible. By using this cell, the starting material can be saved in the development of the electroorganic transformations. To demonstrate the utility of this particular cell, two transformations of important building blocks for the fine chemical and pharmaceutical industry were established including an efficient and simple workup protocol.

Synthesis of nanoparticle-built FeVO4 microspheres with improved low temperature catalytic activity in ammoxidation

Chen, Fang,Chen, Piao,Hu, Hao,Li, Xiongjian,Yang, Shuijin

, (2021/10/12)

FeVO4 microspheres were successfully fabricated by a hydrothermal approach. Characterization results revealed that the FeVO4 microspheres are composed of numerous nanoparticles. In the ammoxidation of 2,6-dichlorotoluene (DCT) to 2,6-dichlorobenzonitrile (DCBN) catalyzed by FeVO4 microspheres, the conversion of DCT reaches up to 87% at a much lower temperature of 320 °C in comparison to previously reported catalysts (380–420 °C). The reason for this is that irons greatly improve the electron transport, leading to well reduction redox properties of vanadium species which are important to enhance the initial activation efficiency of C-H bond in methyl group.

Post a RFQ

Enter 15 to 2000 letters.Word count: 0 letters

Attach files(File Format: Jpeg, Jpg, Gif, Png, PDF, PPT, Zip, Rar,Word or Excel Maximum File Size: 3MB)

1

What can I do for you?
Get Best Price

Get Best Price for 1194-65-6