Showing posts with label anthony crasto. Show all posts
Showing posts with label anthony crasto. Show all posts

Sunday 21 July 2013

Enantioselective Michael Addition


Coming on the heels of the very nice combined computational/experimental study of the enantioselective Strecker reaction by Jacobsen , there’s this JACS communication that really disappoints in its use of computational chemistry. Cobb uses yet another chiral thiourea to produce the enantioselective intramolecular Michael addition of nitronoates (Reaction1).1 The reaction goes with excellent diastereoselectivity and eneatioselectivity, and can even be done with a substrate to produce three chiral centers. This is very nice synthetic chemistry.
Reaction 1.
dr >19:1
ee 95%
The lack of reactivity of the Z ester suggested that the thiourea must associate with both the nitro group and the ester carbonyl. The authors provide a B3LYP/3-21G complex of thiourea with a simple nitroester (once again without providing coordinates in the supporting materials!) to demonstrate this sort of association. But this single structure, at this very low computational level, with these simplified reagents, and lacking solvent (see Rzepa’s comment) really makes one wonder just what value this computation provides. It also goes to demonstrate just how much effort Jacobsen went through to provide substantive computational support for his proposed mechanism of action.

References

(1) Nodes, W. J.; Nutt, D. R.; Chippindale, A. M.; Cobb, A. J. A., "Enantioselective Intramolecular Michael Addition of Nitronates onto Conjugated Esters: Access to Cyclic γ-Amino Acids with up to Three Stereocenters," J. Am. Chem. Soc. 2009, 131, 16016-16017, DOI: 10.1021/ja9070915

Novel cyclophanes: Out-of-Plane Bending and Aromaticity


The novel cyclophanes 1 and 2 have now been synthesized.1 An interesting question is whether the bent pyrenes portion of the two molecules remains aromatic. The bending angles is 93.8° in 1and 95.8° in 2. This distortion is readily apparent in Figure 1, which presents their B3LYP/6-311G(d,p) optimized geometries. NICS computations were used to assess the aromaticity of the pyrene portion. The central rings of pyrene have NICS(0) = -4.4 ppm. The corresponding values in1 and 2 are -4.5 ppm. The apical rings of pyrene have NICS(0)= -11.9 ppm, while the value is -11.1 ppm in 1 and -11.0 ppm in 2. These calculations indicate that the molecule retains much of the aromaticity of the parent pyrene despite the significant out-of-plane distortions.
Figure 1. B3LYP/6-311G(d,p) optimized geometries of 1 and 2.1

1

2

References

(1) Zhang, B.; Manning, G. P.; Dobrowolski, M. A.; Cyranski, M. K.; Bodwell, G. J., "Nonplanar Aromatic Compounds. 9. Synthesis, Structure, and Aromaticity of 1:2,13:14-Dibenzo[2]paracyclo[2](2,7)-pyrenophane-1,13-diene," Org. Lett., 2008, 10, 273-276, DOI: 10.1021/ol702703b.

Friday 19 July 2013

Do You Remember Darzen condensation Reaction




The Darzens reaction (also known as the Darzens condensation or glycidic ester condensation) is the chemical reaction of a ketone or aldehyde with an α-haloester in the presence of base to form an α,β-epoxy ester, also called a "glycidic ester".[1][2] This reaction was discovered by the organic chemist Auguste George Darzens in 1904.[3]
Darzens reaction overview.png


The reaction process begins when a strong base is used to form a carbanion at the halogenated position. Because of the ester, this carbanion is a resonance-stabilized enolate, which makes it relatively easy to form. This nucleophilic structure attacks another carbonyl component, forming a new carbon–carbon bond. These first two steps are similar to a base-catalyzed aldol reaction. The oxygen anion in this aldol-like product then does an intramolecular SN2 attack on the formerly-nucleophilic halide-bearing position, displacing the halide to form an epoxide.[4] This reaction sequence is thus a condensation reaction, since there is a net loss of HCl when the two reactant molecules join.Reaction mechanism

Darzensepoxide.png
The primary role of the ester is to enable the initial deprotonation to occur, and other carbonyl functional groups can be used instead. If the starting material is an α-halo amide, the product is an α,β-epoxy amide.[5] If an α-halo ketone is used, the product is an α,β-epoxy ketone.[4]
Any sufficiently strong base can be used for the initial deprotonation. However, if the starting material is an ester, the alkoxide corresponding to the ester side-chain is commonly in order to prevent complications due to potential acyl exchange side reactions.

Stereochemistry

Depending on the specific structures involved, the epoxide may exist in cis and trans forms. A specific reaction may give only cis, only trans, or a mixture of the two. The specific stereochemical outcome of the reaction is affected by several aspects of the intermediate steps in the sequence.
The initial stereochemistry of the reaction sequence is established in the step where the carbanion attacks the carbonyl. Two sp3 (tetrahedral) carbons are created at this stage, which allows two different diastereomeric possibilities of the halohydrin intermediate. The most likely result is due to chemical kinetics: whichever product is easier and faster to form will be the major product of this reaction. The subsequent SN2 reaction step proceeds with stereochemical inversion, so the cis or trans form of the epoxide is controlled by the kinetics of an intermediate step. Alternately, the halohydrin can epimerize due to the basic nature of the reaction conditions prior to the SN2 reaction. In this case, the initially formed diastereomer can convert to a different one. This is an equilibrium process, so the cis or trans form of the epoxide is controlled by chemical thermodynamics--the product resulting from the more stable diastereomer, regardless of which one was the kinetic result.[5]

Alternative reactions

Glycidic esters can also be obtained via nucleophilic epoxidation of an α,β-unsaturated ester, but that approach requires synthesis of the alkene substrate first whereas the Darzens condensation allows formation of the carbon–carbon connectivity and epoxide ring in a single reaction.

Subsequent reactions

The product of the Darzens reaction can be reacted further to form various types of compounds. Hydrolysis of the ester can lead to decarboxylation, which triggers a rearrangement of the epoxide into a carbonyl (4). Alternately, other epoxide rearrangements can be induced to form other structures.
The Darzens reaction

References

  1.  Darzens, G. (1905). Compt. Rend. 141: 766.
  2.  Darzens, G. (1906). Compt. Rend. 142: 214.
  3.  Darzens, G. (1904). Compt. Rend. 139: 1214.
  4. Jie Jack Li (2006). "Darzens glycidic ester condensation". Name Reactions (3rd. expanded ed.). Springer-Verlag. pp. 183–184. doi:10.1007/3-540-30031-7.
  5. Tung, C. C.; Speziale, A. J.; Frazier, H. W. (1963). "The Darzens Condensation. II. Reaction of Chloroacetamides with Aromatic Aldehydes". The Journal of Organic Chemistry 28 (6): 1514. doi:10.1021/jo01041a018. edit
Review articles

  • Newman, M. S.; Magerlein, B. J. (1949). Org. React. 5: 413.
  • Ballester, M. (1955). "Mechanisms of The Darzens and Related Condensations Manuel Ballester". Chem. Rev. 55 (2): 283. doi:10.1021/cr50002a002.
  • Rosen, T. (1991). Comp. Org. Syn. 2: 409–439.




Enantioselective Carbene-Catalyzed Annulations

 

2-benzoxopinones were synthesized by a dual activation strategy, involving N-heterocyclic carbene catalysis and a second Lewis base
Read more


http://www.chemistryviews.org/details/news/5017851/Enantioselective_Carbene-Catalyzed_Annulations.html

Wednesday 17 July 2013

Camphor and Rivanol (Ethacridine Lactate) - An interesting reaction


Camphor and Rivanol (Ethacridine Lactate) - An interesting reaction



Two days ago, while trying to clean a pimple with an antiseptic camphor solution, I accidentally used ethacridine lactate (Rivanol) instead. After realizing this, instead of first cleaning the rivanol stain and then applying the camphor, I directly cleaned it with camphor using a cotton pad. Suddenly a new white layer formed at the contact surface between the two layers of substance. The occurrence of a chemical reaction was obvious, and after repeating the experiment in lab conditions, I tried to understand the formation of this new product.


Camphor structure


Rivanol structure
 READ ALL THIS AT
http://dailychem.blogspot.in/2011/11/camphor-and-rivanol-ethacridine-lactate.html

Quindine catalyzed synthesis of spirooxindoles through sequential Michael-Henry reactions.

Abstract Image
A novel organocatalytic strategy for the synthesis of highly substituted spirocyclopentaneoxindoles was developed employing simple nitrostyrenes and 3-substituted oxindoles as starting materials. Michael–Henry cascade reactions, enabled through cinchona alkaloid organocatalysis, provided products in high yield and excellent enantioselectivity in a single step.

Spirooxindoles in one step


If you just leaf through an organic chemistry journal, from JOC to Organic Letters, from EJOC to OBC, there is a huge number of papers describing synthesis of oxindoles! Indeed, the interest in the oxindoles scaffold rapidly increased in the last years and several methodologies have been published for the asymmetric synthesis of 3,3’-disubstituted oxindoles.
Cascade organacatalysis is a powerful tool for the for the asymmetric synthesis of very complex structures containing several stereogenic centers and oxindoles are optimal templates for the applications of these wonderful catalytic methodologies. For example, prof. Paolo Melchiorre, one the most expert in the organocatalysis field, published several works in the application of organocascade catalysis for the synthesis of quaternary oxindoles.
In this communication, Carlos Barbas, another pioneer of organocatalysis, reported a quindine-catalyzed synthesis of spirooxindoles through sequential Michael-Henry reactions.
They first performed a preliminary methodological study aimed at finding the most efficient catalyst, which turned out to be quindine derivative depicted below.
Subsequently the scope of the reaction has been expanded, by employing a wide range of substituted oxindoles, as well as different nitrostyrens. Performing the reaction at 0°C in CH2Cl2 as solvent in presence of quinidine derivative 1 permitted the isolation of the products in very high yield and optimal enantiomeric excess and diastereoisomeric ratio.
It is noteworthy to mention that this simple methodology permits the straightforward construction of the spirooxindole moiety containing four stereogenic centers,, with full control of the stereochemical outcome. This is another demonstration of the powerful of organocatalysis. 
http://pubs.acs.org/doi/abs/10.1021/ol300441z

Assembly of Spirooxindole Derivatives Containing Four Consecutive Stereocenters via Organocatalytic Michael–Henry Cascade Reactions

The Skaggs Institute for Chemical Biology and the Departments of Chemistry and Molecular Biology, The Scripps Research Institute, 10550 North Torrey Pines Road, La Jolla, California 92037, United States
Org. Lett.201214 (7), pp 1834–1837
DOI: 10.1021/ol300441

Tuesday 16 July 2013

Efficient Odorant Synthesis

Efficient Odorant Synthesis
Efficient Odorant Synthesis

2-Nonen-4-olide occurs naturally in the volatile components of various foods and plant extracts, such as stewed beef/vegetable gravy,  sunflower oil,  Chinese white salted noodles watermelon, potato crisps,  lard cured with spices and aromatic herbs,  mushrooms, French fries, and flue-cured tobacco. It has an oily, coconut-like, and rancid odour. A simple, efficient and low cost synthesis of this odorant is reported in the August issue of Journal of Chemical Research.

Sunday 30 June 2013

Beckmann rearrangement




The Beckmann rearrangement, named after the German chemist Ernst Otto Beckmann (1853–1923), is an acid-catalyzed rearrangement of an oxime to an amide. Cyclic oximes yield lactams.
The Beckmann Rearrangement

 
This example reaction starting with cyclohexanone, forming the reaction intermediate cyclohexanone oxime and resulting in caprolactam is one of the most important applications of the Beckmann rearrangement, as caprolactam is the feedstock in the production of Nylon 6.
The Beckmann solution consists of acetic acid, hydrochloric acid and acetic anhydride, and was widely used to catalyze the rearrangement. Other acids, such as sulfuric acid or polyphosphoric acid, can also be used. sulfuric acid is the most commonly used acid for commercial lactam production due to its formation of an ammonium sulfate by-product when neutralized with ammonia. Ammonium sulfate is a common agricultural fertilizer providing nitrogen and sulfur.

The Beckmann rearrangement is an organic reaction used to convert an oxime to an amide under acidic conditions. The reaction begins by protonation of the alcohol group forming a better leaving group. The R group trans to the leaving group then migrates to the nitrogen, resulting in a carbocation and the release of a water molecule. This trans [1-2]-shift allows for the prediction of the regiochemistry of this reaction. The water molecule then attacks the carbocation and after deprotonation and tautomerization results in the final amide product.





Beckmann rearrangement mechanism



lactam, a monomer for the production of Nylon 12.



Beckmann reaction

SEE BELOW ILLUSTRATION IN

http://www.mdpi.com/1420-3049/17/11/13662

 Upon treatment with phenyl dichlorophosphate (PhOP=OCl2) in acetonitrile at ambient temperature, a variety of ketoximes underwent a Beckmann rearrangement in an effective manner to afford the corresponding amides in moderate to high yields.

Saturday 15 June 2013

REAGENTS FOR ORGANIC SYNTHESIS, my WEBSITE


SEE MY FULL WEBSITE
https://sites.google.com/site/anthonycrastoreagents/

CONTENTS

Monday 27 May 2013

Friday 24 May 2013

A new method for producing clean hydrogen



TEM_image_Au-a-Fe2O3_catalyst






Duke University engineers have developed a novel method for producing clean hydrogen, which could prove essential to weaning society off of fossil fuels and their environmental implications. While hydrogen is ubiquitous in the environment, producing and collecting molecular hydrogen for transportation and industrial uses is expensive and complicated. Just as importantly, a byproduct of most … more…

Thursday 23 May 2013

One-Pot Method for Regioselective Bromin­ation and Sequential Carbon–Carbon Bond-Forming Reactions of Allylic Alcohol Derivatives

One-Pot Method for Regioselective Bromin­ation and Sequential Carbon–Carbon Bond-Forming Reactions of Allylic Alcohol Derivatives
European Journal of Organic Chemistry Noriki Kutsumura, Yusuke Matsubara, Kentaro Niwa, Ai Ito and Takao Saito
DOI: 10.1002/ejoc.201300173

Thumbnail image of graphical abstract



Di- or trisubstituted olefins were synthesized in high yields with excellent regio- and cis–trans selectivities in one-pot reactions, including a regioselective DBU-promoted trans HBr elimination. This one-pot methodology could become a straightforward transformation of “straight” alkenes into “Y-shaped” alkenes.


An efficient one-pot method for the regioselective bromination of allylic alcohol derivatives (two-step reaction sequence) followed by Sonogashira, Negishi, or Suzuki–Miyaura coupling reactions in the same reaction vessel (three-step reaction sequence) has been developed. The key reaction in these one-pot systems is the regioselective DBU-promoted trans HBr elimination of vicinal dibromides bearing an adjacent O-functional group.

Synthetic Uses of Ammonia in Transition-Metal Catalysis

European Journal of Organic Chemistry


Jinho Kim, Hyun Jin Kim and Sukbok Chang
DOI: 10.1002/ejoc.201300164
Thumbnail image of graphical abstract

Although ammonia (NH3) is a cheap, abundant, and readily available nitrogen source, it has rarely been used in transition-metal catalysis, due to several obstacles. However, significant advances in the metal-mediated utilization of ammonia have been made recently. This review presents the most recent examples in metal-mediated amination and other relevant reactions with ammonia or ammonium salts.
http://onlinelibrary.wiley.com/doi/10.1002/ejoc.201300164/abstract

Ammonia (NH3) is a cheap, abundant, and readily available nitrogen source, being one of the chemicals produced in the greatest quantities. Whereas ammonia is utilized mainly as a feedstock for the production of fertilizers, it is also employed in industry as a component of various nitrogen-containing compounds. In metal catalysis, in contrast, ammonia has been used only with limited success, due to several difficulties such as generation of stable Lewis acid-base adducts, facile ligand exchange for ammonia in active metal complexes, a propensity towards undesired second transformations of initially formed species, and the requirement for special equipment to run the reactions. Despite these obstacles, the direct use of ammonia in catalysis has continuously attracted great interest, leading recently to significant progress. Whereas liquid or gaseous ammonia were most commonly employed in the past, under harsh conditions, notable catalytic reactions using easy-to-handle ammonium salts under milder and more convenient conditions have now been developed. In this review we briefly describe the most recent examples of transition-metal-catalyzed reactions using ammonia or ammonium salts.

Friday 17 May 2013