International Journal of Organic Chemistry
Vol.3 No.3(2013), Article ID:37186,4 pages DOI:10.4236/ijoc.2013.33028

Regioselective Direct Carboxylation of 2-Naphthol with Supercritical CO2 in the Presence of K2CO3

Shigeki Furukawa1, Kensuke Otokawa1, Osamu Sasaki2, Ken Nakayasu2,3, Tatsuaki Yamaguchi2*

1College of Industrial Technology, Nihon University, Chiba, Japan

2Faculty of Engineering, Chiba Institute of Technology, Chiba, Japan

3Instrumental Analysis Division, Center for Supports to Research and Education Activities, Kobe University, Kobe, Japan

Email: *tatsuaki.yamaguchi@it-chiba.ac.jp

Copyright © 2013 Shigeki Furukawa et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Received July 5, 2013; revised August 15, 2013; accepted September 1, 2013

Keywords: Supercritical CO2; CO2 Utilization; Kolbe-Schmitt Reaction; Carboxylation; 2-Naphthol; Potassium Carbonate; 2-Hydroxynaphthalene-6-carboxylic Acid; DFT Calculation

ABSTRACT

A direct regioselective preparation of 2-hydroxynaphthalene-6-carboxylic acid, a useful industrial intermediate of aromatic polyester from 2-naphthol was conducted by use of excess amount of K2CO3 (10-fold molar to 2-naphthol) under supercritical CO2 at 10 MPa and 473 K. The obtained yield under this condition was ca. 20 mol% to 2-naphthol. The further investigations may provide an alternative process to the conventional Kolbe-Schmitt reaction, because of no use of strong alkali and recoverability of K2CO3. Theoretical explanation about the regioselectivity was achieved by means of DFT calculations.

1. Introduction

Aromatic hydroxyl carboxylic acids have been prepared over a long period of time via Kolbe-Schmitt reaction [1], utilizing CO2 as a carbon resource. However, the conventional process, which is a heterogeneous reaction in the gas-solid phase, requires a high temperature and a long reaction time. The process consists of three steps: 1) the preparation of alkali metal salts of hydroxylarenes with strong alkali; 2) their desiccation; and 3) the carboxylation with CO2 in a pressure device. A number of ameliorating trials for its process have been examined up to the present, because aromatic hydroxyl carboxylic acids have been widely utilized for various purposes by now, for example, as valuable raw materials or intermediates in the production of pharmaceutical and fine chemicals.  

It is well known that the carboxylation can be carried out under moderate temperatures and pressures when an organic solvent is used [2-6]. One of the authors (TY) found that the reaction readily occurred at atmospheric pressure of CO2 at 333 K in benzene with small amount of crown ether or polar solvent, which created a homogenous solution with alkali metal 2-naphtholate [7]. However, in view of the contemporary concept of socalled “green chemistry” or “sustainable society”, it is required to get rid of organic solvent from the chemical process as much as possible. The development of utilization of CO2 is commendable under the same sort of intentions. This prompts us to make use of supercritical CO2 as a medium and a reagent of the carboxylation. As a result, comparative studies of the reaction between sodium phenolate and CO2 under the gaseous and supercritical conditions have demonstrated the preferential effect of the supercritical condition for promoting the reactivity of CO2 [8]. Furthermore, the direct carboxylation of phenol to salicylic acid, namely without the preparation of the alkali metal salt, has been observed regioselectively with AlBr3 under supercritical CO2 condition [9]. The direct selective synthesis of salicylic acid from phenol with K2CO3 has been performed in a similar way [10].

In this paper, the examination of the similar direct carboxylation of 2-naphthol with K2CO3 under supercritical CO2 conditions will be reported shortly with respect to the regioselective synthesis of 2-hydroxy-naththalene-6-carboxylic acid (2-H-6-NA), which is a useful chemical intermediate of aromatic polyesters.

2. Experimental

2.1. Material

2-Naphthol (Kanto Chemical Co., Inc., Tokyo, Japan) and K2CO3 (Kanto Chemical Co., Inc., Tokyo, Japan) of reagent grade were used without further purifications. The standard 2-hydroxynaphthalene-6-carboxylic acid (2-H-6-NA), 2-hydroxynaphthalene-1-carboxylic acid (2-H-1-NA) and 2-hydroxynaphthalene-3-carboxylic acid (2-H-3-NA) were supplied from Tokyo Chemical Industry Co., Ltd and employed as gas-chromatography standards.  

2.2. Carboxylation

The experimental system was basically the same with the previous report [10], using a stainless steel autoclave reactor (AKICO Corporation; Tokyo, Japan). After CO2 was introduced to the reactor, the reaction started under the prescribed conditions. After the reaction, the resulting product was dissolved in tetrahydrofuran. The solution was filtered and removed the solvent by a rotary evaporator.  

2.3. Analysis

The reaction product of 0.10 g was collected and silylated with 500 μL of N-trimethylsilylimidazole solvation (GL Sciences Inc., Tokyo, Japan) at 313 K for 3 h. Then the pyridine solution was analyzed by gas-chromatogramphy, using dodecane as an internal standard.  

2.4. Computational Method

A quantum chemical calculation was performed by means of the Gausian 03 program package in this study. The DFT method was selected, using the function commonly called as B3LYP. The geometrical optimization was carried out using the LANZL2DZ basis set.

3. Results & Discussion

The topicality of this communication is an inexplicable effect of the coexistent K2CO3 on the carboxylation of 2-naphthol in supercritical condition of CO2. The increase of the molar ratio of K2CO3 to 2-naphthol resulted in the increase of the yield and hence the regioselectivity to 2-hydroxynaphthalene-6-carboxylic acid (2-H-6NA). Its maximum yield of ca. 20% was obtained currently with the ratio of 10 at 10MPa and 473 K. The increase of product yields with the ratio is presented in Figure 1. The reduction of pressure and temperature from this attained conditions retarded significantly the yields as shown in Table 1.

As it has been reported in our previous paper [10], the similar effect of the molar ratio of K2CO3 has been ob-

Figure 1. Effect of molar ratio of K2CO3 to 2-naphthol on the yield of 2-hydroxynaphthalene carboxylic acid under the pressure of 10 MPa, 473 K, duration 8 h.

Table 1. Effect of reaction pressure and temperature on the carboxylation of 2-naphthol with carbon dioxide at molar ratio 10 of K2CO3, duration 8 h.

served with respect to the carboxylation of phenol in the supercritical CO2 condition. In this case, the salicylic acid yield increases with an increase of K2CO3 up to the equimolecular amount with phenol. The catalytic action of K2CO3, formerly proposed, should be withdrawn, because its performances are not catalytic in the strict meanings. Of course, K2CO3 may form a potassium salt with phenol to some extent to activate the carboxylation. But the formation would be a little because of its lower basicity than potassium hydroxide, conventionally used for Kolbe-Schmitt reaction.

The requirement of the presence of excess K2CO3 under supercritical CO2 condition might be matter of solvation. It has been pointed out [11,12] that the reaction rates are maximal near the critical pressure of CO2 when CO2 is used as a reactant and a solvent. This is in accordance with our former results [8,10]. These phenomena may be explained by the local augment effect, namely the reaction rate increases because of the clustering solvent molecule around the solute species [13]. On the contrary, excessively high CO2 pressures may retard the interaction between reactants because of increasing the density by clustering of CO2 molecules [14-16]. As CO2 has a lower dielectric constant, K2CO3 may behave as a substitute to cluster around 2-naphtholate in the present case, separating each 2-naphthol molecule from others as a strainer does in the extraction with supercritical fluids. The aggregation of reactants would become a serious problem in the supercritical condition.

Previously, it has been shown that the carboxylation of 2-naphtholate readily occurs through the addition of crown ethers or polar solvents in a nonpolar solvent, affording the selective production of 2-hydroxynaphthalene-1-carboxylic acid (2-H-1-NA) under ordinary conditions [7]. This is because the separation of alkali cation from 2-naphtholate by the solvation enhances greatly its reactivity toward CO2. As shown in Figure 2(a), the HOMO distribution of the naked 2-naphtholate is localized on 1-position of naphthalene ring, making it possible to attack by CO2 directly to this position. On the other hands, a carbonate like complex formation between potassium phenolate and CO2 has been proposed for the conventional Kolbe-Schmitt reaction [17,18]. The reaction mechanism is considered as the migration of carbonate group to the ortho-position. The HOMO distribution of potassium 2-naphtholate (Figure 2(b)) is decreased by the direct binding of K+, which causes the requirement for the higher reaction temperature in the conventional methods.

The possibility of complex formation of 2-naphthol with K2CO3, as speculated above, was examined by DFT calculations. The resulting optimized geometry is shown in Figure 2(c), being K+ found at the region between O and C of 1-position of 2-naphthol molecule. A moiety is located nearby. Obviously, these situations might prohibit the attack of CO2 molecule toward C of 1-position. The additional calculations on the complex formation with CO2 molecule gave no geometry optimization. This suggests the difficulty of salvation by CO2 in the supercritical state.

(a)(b)(c)

Figure 2. The results of DFT calculations. (a) 2-naphtholate (HOMO); (b) potassium 2-naphtholate (HOMO); (c) 2-naphthol-K+ complex (total density).

The increase of the product yield and selectivity to 2-H-6-NA by the increase of the molar ratio of K2CO3 could be explained by the solvation (or clustering) effect by some additional K2CO3 to the first one, blocking the most reactive carbon position-1 of 2-naphthol.

REFERENCES

  1. S. A. Lindsey and H. Jesky, “The Kolbe-Schmitt Reaction,” Chemical Reviews, Vol. 57, No. 4, 1957, pp. 583- 620. doi:10.1021/cr50016a001
  2. F. Seidel, L. Wolf and H. Krause, “Über die Bildungsweise der 2-Oxynaphthoesäure-(3),” Journal für Praktische Chemie, Vol. 91, No. 1-2, 1955, pp. 53-83. doi:10.1002/prac.19550020105
  3. W. H. Meek and C. H. Fuchsmann, “Carboxylation of Substituted Phenols in N,N-Dimethylamide Solvents at Atmospheric Pressure,” Journal of Chemical & Engineering Data, Vol. 14, No. 4, 1968, pp. 388-391. doi:10.1021/je60042a005
  4. T. Yamaguchi, N. Nagaoka and K. Takahashi, “The Carboxylation of Naphthols by Kolbe-Schmitt Reaction in the Homogeneous Solution,” The Chemical Society of Japan, No. 7, 1989, pp. 1164-1165.
  5. T. Iijima, T. Iwase and T. Yamaguchi, “Carboxylation of 2-Naphthol with Carbon Dioxide in Anisole,” Journal of the Japan Petroleum Institute, Vol. 49, No. 4, 2006, pp. 206-209. doi:10.1627/jpi.49.206
  6. T. Iijima, D. Takagi and T. Yamaguchi, “Carboxylation of 2-Naphthol in Kerosene,” Journal of the Japan Petroleum Institute, Vol. 31, No. 1, 2008, pp. 65-69. doi:10.1627/jpi.51.65
  7. J. Baxter and T. Yamaguhi, “Effect of Cation Capture by Crown Ether and Polar Solvent in the Carboxylation with CO2 of Alkali Metal 2-Naphtholate under Ordinary Conditions,” Journal of Chemical Research (S), 1997, pp. 374-375. doi:10.1039/A703432F
  8. T. Iijima and T. Yamaguchi, “The Improved Kolbe— Schmitt Reaction Using Supercritical Carbon Dioxide,” Tetahedron Letters, Vol. 48, No. 30, 2007, pp. 5309-5311. doi:10.1016/j.tetlet.2007.05.132
  9. T. Iijima and T. Yamaguchi, “Efficient Regioselective Carboxylation of Phenol to Salicylic Acid with Supercritical CO2 in the Presence of Aluminum Bromide,” Journal of Molecular Catalysis A: Chemical, Vol. 295, 2008, pp. 52-56. doi:10.1016/j.molcata.2008.07.017
  10. T. Iijima and T. Yamaguchi, “K2CO3-Catalyzed Direct Synthesis of Salicylic Acid from Phenol and Supercritical CO2,” Applied Catalysis A: General, Vol. 345, No. 1, 2008, pp. 12-17. doi:10.1016/j.apcata.2008.03.037
  11. S. Fujita, B. M. Bhanage, Y. Ikushima and M. Arai, “Synthesis of Dimethyl Carbonate from Carbon Dioxide and Methanol in the Presence of Methyl Iodide and Base Catalysts under Mild Conditions: Effect of Reaction Conditions and Reaction Mechanism,” Green Chemistry, Vol. 3, No. 2, 2001, pp. 87-91. doi:10.1039/b100363l
  12. T. Sakakura, Y. Saito, M. Okano, J. Choi and C. T. Sako, “Selective Conversion of Carbon Dioxide to Dimethyl Carbonate by Molecular Catalysis,” The Journal of Organic Chemistry, Vol. 63, No. 20, 1998, pp. 7095-7096. doi:10.1021/jo980460z
  13. J. F. Brennkeand and J. E. Chateauneuf, “Homogeneous Organic Reactions as Mechanistic Probes in Supercritical Fluids,” Chemical Reviews, Vol. 99, No. 2, 1999, pp. 433-452. doi:10.1021/cr970035q
  14. Y. Du, F.Cai, D.-L.Kong and L.-N. He, “Organic Solvent-Free Process for the Synthesis of Propylene Carbonate from Supercritical Carbon Dioxide and Propylene Oxide Catalyzed by Insoluble Ion Exchange Resins,” Green Chemistry, Vol. 7, No. 7, 2005, pp. 518-523. doi:10.1039/b500074b
  15. G. Steinebrunner, A. J. Dyson, B. Kirchner and H. Hanspeter, “Structural and Thermodynamic Properties of Fluid Carbon Dioxide from a New ab initio Potential Energy Surface,” Journal of Chemical Physics, Vol. 109, 1998, Article ID. 3153. doi:10.1063/1.476922
  16. K. Rah and B. Eu, “Density Fluctuations and Shear Viscosity of Molecular Liquids: Carbon Dioxide and Nitrogen,” Journal of Chemical Physics, Vol. 112, No. 16, 2000, pp. 7118-7138. doi: 0.1063/1.481306
  17. Z. Markovic, J. P. Engelbrecht and S. Markovic and Z. Naturforch, “Theoretical Study of the Kolbe-Schmitt Reaction Mechanism,” Zeitschrift für Naturforchung, Vol. 57a, 2002, pp. 812-818.
  18. Y. Kosugi, Y. Imaoka, F. Gotoh, M. A. Rahim, Y. Matsui and K. Sakanishi, “Carboxylations of Alkali Metal Phenoxides with Carbon Dioxide,” Organic & Biomolecular Chemistry, Vol. 1, No. 5, 2003, pp. 817-821. doi:10.1039/b210793g

NOTES

*Corresponding author.