22.03.2013 Views

epoxidation of wild safflower (carthamus oxyacantha) - Research ...

epoxidation of wild safflower (carthamus oxyacantha) - Research ...

epoxidation of wild safflower (carthamus oxyacantha) - Research ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

International Journal <strong>of</strong> ChemTech <strong>Research</strong><br />

CODEN( USA): IJCRGG ISSN : 0974-4290<br />

Vol. 3, No.3, pp 1152-1163, July-Sept 2011<br />

Epoxidation <strong>of</strong> Wild Safflower (Carthamus<br />

<strong>oxyacantha</strong>) Oil with Peroxy acid in presence<br />

<strong>of</strong> strongly Acidic Cation Exchange Resin IR-<br />

122 as Catalyst<br />

Pawan D. Meshram*, Ravindra G. Puri, Harshal V. Patil<br />

Department <strong>of</strong> Chemical Technology, North Maharashtra University, Jalgaon,<br />

Maharashtra - 425 001, India;<br />

Tel: +91-0257-2257144; Fax: +91-0257-2258403, 2258406<br />

*Corres.author: pawan.dm@gmail.com<br />

Abstract: Epoxidation brings out the transformations at the unsaturated part <strong>of</strong> oil molecule to introduce reactive<br />

oxirane ring. These plant oil-based epoxides are sustainable, renewable and biodegradable that can replace petroleumderived<br />

materials in numerous industrial applications, such as plasticizer, lubricant and in coatings formulations. In<br />

present work <strong>wild</strong> <strong>safflower</strong> oil (WSO) with an iodine value <strong>of</strong> 155(g I2/100g), and containing 13% oleic acid (C 18:1)<br />

and 72% linoleic acid (C 18:2), was epoxidised using a peroxy acid generated in situ by the reaction <strong>of</strong> aqueous hydrogen<br />

peroxide and carboxylic acid in presence <strong>of</strong> strongly acidic cation exchange resin, Amberlite ® IR-122. Acetic acid was<br />

found to be a better oxygen carrier than formic acid and the use <strong>of</strong> catalyst improved the epoxide yield. The influence <strong>of</strong><br />

various reaction parameters such as temperature, stirring intensity, catalyst loading, and molar reactant ratio exerted on<br />

the <strong>epoxidation</strong> reaction was examined to optimize the condition for achieving maximum epoxy group formation. The<br />

formation <strong>of</strong> the epoxide adduct was confirmed by FTIR spectral analysis.<br />

The rate <strong>of</strong> <strong>epoxidation</strong> was not substantially affected with stirring speed beyond 2000 rpm under the given conditions.<br />

A high temperature <strong>of</strong> above 60 o C is detrimental since at continued reaction it accelerated the rapid destruction <strong>of</strong><br />

oxirane rings. Higher concentrations <strong>of</strong> acetic acid and hydrogen peroxide (above 0.5 & 1.5 mol/mol <strong>of</strong> ethylenic<br />

unsaturation respectively) are also unfavorable as they lead to epoxy ring opening. The study shows that a relative<br />

conversion to epoxide ring moiety <strong>of</strong> 85% can be achieved by using the optimum molar ratio 1:0.5:1.5 (ethylenic<br />

unsaturation: acetic acid: hydrogen peroxide) at temperature 60 o C and 2000rpm.<br />

Key words: Epoxidation, <strong>wild</strong> <strong>safflower</strong> oil, acidic cation exchange resin, reaction parameters.<br />

INTRODUCTION<br />

Vegetable oils and fats, the important renewable<br />

resources, are biodegradable, low in cost and readily<br />

available. A number <strong>of</strong> methods such as chemical and<br />

enzymatic modification <strong>of</strong> oils have been suggested for<br />

improving their properties. The modified oils serve as<br />

feedstock that can replace petroleum-derived materials<br />

in many applications [1] . Thus, there is an immediate<br />

need <strong>of</strong> attention towards effective utilization <strong>of</strong> these<br />

biobased products taking into the consideration <strong>of</strong><br />

current environmental harm and steadily rising price <strong>of</strong><br />

petroleum. Moreover, this trend will reduce the<br />

dependence on imported petroleum and will promote<br />

the sustainable agricultural initiative.<br />

The unsaturation present in fatty acid molecule <strong>of</strong> the<br />

vegetable oil can be used to introduce various<br />

functional groups by carrying out chemical reactions.<br />

Among them, epoxidization is one <strong>of</strong> the most widely


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1153<br />

used reaction. Epoxidation introduces the high strain<br />

energy three-membered oxirane ring into the<br />

unsaturated part <strong>of</strong> the oil molecule, thereby increasing<br />

its complexity and chemical reactivity with a variety <strong>of</strong><br />

compounds such as amines and carboxylic acids<br />

[2, 3]<br />

and serves as a chemical intermediate for the<br />

preparation <strong>of</strong> derivatives that would be difficult to<br />

obtain directly from the unsaturated bonds <strong>of</strong> fatty<br />

acids [4] . This leads to a wide spectrum <strong>of</strong> commercial<br />

uses for epoxides. There are several methods for<br />

producing epoxides from vegetable oils, fatty acids<br />

and methyl esters. These includes <strong>epoxidation</strong> with<br />

percoarboxylic acid generated in situ in the presence <strong>of</strong><br />

acids or enzymes as catalysts [5,6] , <strong>epoxidation</strong> with<br />

organic and inorganic oxidants such as potassium<br />

peroxomonosulphate, meta-chloroperoxybenzoic acid<br />

and ethyl methydioxirane [7,8,9,10] , <strong>epoxidation</strong> with<br />

halohydrins for the <strong>epoxidation</strong> <strong>of</strong> olefins with<br />

electron-deficient double bonds and <strong>epoxidation</strong> with<br />

molecular oxygen [6] . From process, environmental<br />

safety and efficiency point <strong>of</strong> view <strong>epoxidation</strong> <strong>of</strong><br />

vegetable oils in one step, i.e., with peroxy acid<br />

generated in situ from carboxylic acid (formic/acetic<br />

acid) with hydrogen peroxide in the presence <strong>of</strong> acid<br />

catalyst is widely used on an industrial scale.<br />

However, the use mineral acid as catalyst in<br />

<strong>epoxidation</strong> is inefficient because <strong>of</strong> problems<br />

associated with separation <strong>of</strong> the catalyst from the<br />

reaction product. The process can be made competitive<br />

with the use <strong>of</strong> ion-exchange catalyst instead <strong>of</strong><br />

traditional homogeneous one in <strong>epoxidation</strong> <strong>of</strong><br />

unsaturated compounds [11] . Epoxidized vegetable oils<br />

are widely used in polymer chemistry because <strong>of</strong> their<br />

excellent plasticizing action for poly (vinyl chloride)<br />

and as effective component for stabilizing polymer<br />

formulations [12,13] . Epoxides find industrial application<br />

as diluents [14] , lubricants [15] and coatings [16] .<br />

Industrially epoxides have been prepared from castor,<br />

linseed, rapeseed, soybean and sunflower oils<br />

containing high to moderate unsaturation. The progress<br />

on the utilization <strong>of</strong> other seed oils <strong>of</strong> semidrying<br />

nature like flax, canola, karanja, jatropha, rubber seed,<br />

cottonseed, olive, corn etc. for epoxide synthesis at<br />

laboratory level is very encouraging towards<br />

development <strong>of</strong> an efficient and safe <strong>epoxidation</strong><br />

process. Safflower is an important oil seed crop <strong>of</strong> the<br />

tropical countries, which belongs to the family<br />

Compositae or Asteraceae and genus Carthamus. C.<br />

tinctorius is the only cultivated species <strong>of</strong> this genus<br />

which is used primarily for edible oil and the other<br />

sister species are <strong>wild</strong> and weed. Taking into account<br />

<strong>of</strong> higher proportion <strong>of</strong> unsaturation in fatty acid<br />

composition <strong>of</strong> the <strong>safflower</strong> oil, it is characterized as<br />

semidrying and finds its use in paints, textile and<br />

leather industries [17, 18] . India is the largest producer <strong>of</strong><br />

<strong>safflower</strong> in the world with an area <strong>of</strong> 3.63 ×10 6<br />

hectares, production <strong>of</strong> 2.29 ×10 6 tonnes and<br />

productivity <strong>of</strong> 631 kg/ha [19] . However, being almost<br />

completely utilized as edible oil in India, there is an<br />

increased attention <strong>of</strong> researchers to search the low<br />

cost renewable resource as alternate to the cultivated<br />

<strong>safflower</strong> oil. Among the <strong>wild</strong> & weed species, C.<br />

<strong>oxyacantha</strong> and C. palaestinus were reported to be<br />

progenitor <strong>of</strong> C. tinctorius, being closely related to<br />

cultivated species with similarity coefficient more than<br />

0.95 [20] . The <strong>wild</strong> species C. <strong>oxyacantha</strong> is<br />

widespread in Turkey, subtropical regions <strong>of</strong> western<br />

Iraq, Iran, North-west India, throughout kazakistan,<br />

Turkmenistan and Uzbekistan [21] . In India <strong>wild</strong><br />

<strong>safflower</strong> grows in the arid regions <strong>of</strong> North India<br />

particularly in Punjab and Uttar Pradesh. It is winter<br />

growing anuual plant with distribution in the wheat<br />

fields near Bhognipur, Bhigapur, Pokhrayan,<br />

Ghatampur and adjoining areas <strong>of</strong> Kanpur district.<br />

The fatty acid composition <strong>of</strong> raw and refined oil<br />

extracted from the seeds <strong>of</strong> C. <strong>oxyacantha</strong> is reported<br />

by Banerji et.al. [22] . The exploitation <strong>of</strong> <strong>wild</strong> <strong>safflower</strong><br />

seed oil (WSO) in preparation <strong>of</strong> epoxides is still the<br />

area not discussed. This paper investigates the <strong>wild</strong><br />

<strong>safflower</strong> seed oil epoxide (EWSO) synthesis<br />

methodology with the use <strong>of</strong> strongly acidic cation<br />

exchange resin (Amberlite ® IR-122) as catalyst. Both<br />

formic acid and acetic acid were used as carboxylic<br />

acids and a screening <strong>of</strong> four acid–catalyst<br />

combinations were done to find out the most efficient<br />

one. The influence <strong>of</strong> various reaction parameters such<br />

as temperature, stirring speed, catalyst loading, acetic<br />

acid-to-ethylenic unsaturation molar ratio, hydrogen<br />

peroxide-to-ethylenic unsaturation molar ratio on<br />

<strong>epoxidation</strong> reaction is focused with intention to<br />

optimize the condition for maximum epoxy yield with<br />

highest conversion <strong>of</strong> double bonds to oxirane rings.<br />

MATERIALS AND METHODS<br />

Materials<br />

About 20 Kg seeds <strong>of</strong> C. <strong>oxyacantha</strong> were collected<br />

from the adjoining regions <strong>of</strong> Punjab district and<br />

extracted in lab-scale immersion-percolation type<br />

batch extractor (Fig.1) with Rankosolv SG (distillation<br />

range 63.0-70.0°C; water content, as determined by<br />

KF method, 0.1%) solvent obtained from Rankem<br />

RFCL, New Delhi. The five number <strong>of</strong> extraction<br />

stages (20 min for each stage) were evaluated, below<br />

the boiling point <strong>of</strong> solvent with miscella recycle flow<br />

rate <strong>of</strong> 9mL/sec, for oil yield . The seeds yielded 26%<br />

<strong>of</strong> pale yellow oil, which after due analytical study was<br />

used for <strong>epoxidation</strong> as such without any refinement.<br />

The chemicals glacial acetic acid, 98% and formic<br />

acid, 90% were obtained from Qualigens India Ltd.,<br />

Mumbai. AR grade hydrogen peroxide (30 wt.%),


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1154<br />

Iodine monochloride and Wij’s solution were procured from S.D. Fine Chemicals Ltd., Mumbai.<br />

Figure 1: Lab-scale immersion-percolation type batch extractor<br />

HBr in acetic acid (33 wt.%) was obtained from<br />

Fischer Scientific, Mumbai and then diluted with<br />

glacial acetic acid to make 0.1N HBr. Strongly acidic<br />

cation exchange resin (Amberlite ® IR-122, sulfonic<br />

acid functionality) with total cation exchange capacity<br />

<strong>of</strong> 2.1meq/mL was procured from Sigma–Aldrich<br />

(USA).<br />

Epoxidation experiments<br />

The <strong>epoxidation</strong> <strong>of</strong> WSO was carried out in batch<br />

mode in a glass reactor consisting <strong>of</strong> a four-necked<br />

round bottom flask <strong>of</strong> 500ml capacity. A motor driven<br />

speed regulator stirrer was inserted in the reactor<br />

through the central neck while other neck was used for<br />

thermo-electrode. A condenser was fitted to the reactor<br />

through third neck and the fourth neck is used for<br />

dropping the raw materials into the reactor. The reactor<br />

was heated by an electric heating mantle having<br />

special arrangement for smooth and accurate control <strong>of</strong><br />

the temperature within ±1°C <strong>of</strong> the desired<br />

temperatures.<br />

The <strong>epoxidation</strong> method reported by Dalai et.al. [23]<br />

was used, and the same procedure was employed for<br />

all <strong>of</strong> the experimental runs dealing with different<br />

parameters. The <strong>epoxidation</strong> reactions were first<br />

carried out with both formic acid and acetic acid as<br />

carboxylic acids and with & without catalyst (IR-122)<br />

combinations. These initial <strong>epoxidation</strong> experiments<br />

were conducted to select the most active carboxylic<br />

acid-catalyst combination. Further <strong>epoxidation</strong> runs<br />

were carried out with the best reagent–catalyst<br />

combination using the following range <strong>of</strong> parameters:<br />

stirring speeds, 1000-2500 rpm; temperature, 40-80 o C;<br />

hydrogen peroxide-to-ethylenic unsaturation molar<br />

ratio, 0.5-2.0 and acetic acid-to-ethylenic unsaturation<br />

molar ratio, 0.3-1.0; catalyst loading (expressed as the<br />

weight percentage <strong>of</strong> WSO used), 5-20%.<br />

A calculated amount <strong>of</strong> WSO (35 g) was introduced in<br />

the reactor and heated with agitation to a temperature<br />

<strong>of</strong> 40°C for 15 min. The requisite amount <strong>of</strong><br />

acetic/formic acid (carboxylic acid to ethylenic<br />

unsaturation molar ratio, 0.5:1) and catalyst (20 wt.%)<br />

were added, and the mixture was continuously stirred<br />

at 2000 rpm for 30 min. Then, 33.76 g <strong>of</strong> 30% aqueous<br />

hydrogen peroxide (1.5 mole <strong>of</strong> hydrogen peroxide per<br />

mole <strong>of</strong> ethylenic unsaturation) was added dropwise to<br />

the reaction mixture at a rate such that the hydrogen<br />

peroxide addition was completed over a period <strong>of</strong> 30<br />

min with gradual increase in temperature to 60°C.<br />

After complete addition <strong>of</strong> hydrogen peroxide, the<br />

reaction was continued for 8 hours, considering the<br />

completion <strong>of</strong> addition <strong>of</strong> hydrogen peroxide to be the<br />

zero time, at controlled temperature 60±1°C with<br />

continuous stirring to ensure fine dispersion <strong>of</strong> oil.<br />

The course <strong>of</strong> the reaction was followed by<br />

withdrawing samples at regular intervals, the first


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1155<br />

being taken after one hour since zero time. The<br />

collected samples were washed with warm water<br />

successively to make them acid free and extracted with<br />

diethyl ether in a separating funnel to enhance the<br />

separation <strong>of</strong> the oil product from water phase and<br />

dried using a rotary evaporator. The progress <strong>of</strong><br />

<strong>epoxidation</strong> was monitored by determination <strong>of</strong><br />

oxirane content and iodine value <strong>of</strong> the washed<br />

aliquots <strong>of</strong> the reaction mixture. Two replication <strong>of</strong><br />

each experiment were performed alongside to<br />

determine the percentage deviation between two<br />

experimental results and the deviation was found to be<br />

less than 5%.<br />

Analytical methods<br />

Acid value, Saponification value, Iodine value and<br />

Specific gravity <strong>of</strong> the raw WSO were determined<br />

according to AOCS Method Cd 3d-63, AOCS Method<br />

Cd 3-25, AOCS Method Cd 1b-87 and MS 817:1989<br />

respectively. The percentage <strong>of</strong> oxirane oxygen<br />

content was determined using AOCS Official Method<br />

Cd 9-57 (1997) under which the oxygen is titrated<br />

directly with hydrogen bromide solution in acetic acid.<br />

The conversion <strong>of</strong> double bonds to oxirane rings were<br />

observed under FTIR spectra, measured on Shimazdu<br />

FTIR-8400spectrometer equipped with DLATGS<br />

detector and a KBr beam splitter into the range <strong>of</strong> 4000<br />

to 400 cm −1 . The absorbance spectra for each analysis<br />

were averaged over 32 scans with a nominal 4 cm −1<br />

resolution.<br />

From the oxirane content, the percentage relative<br />

conversion to oxirane was determined using the<br />

following formula [3] :<br />

Relative conversion to oxirane (RCO)<br />

= [OOex / OOth] × 100 ----------------------------(1)<br />

where OOex is the experimentally determined content<br />

<strong>of</strong> oxirane oxygen and OOth is the theoretical<br />

maximum oxirane oxygen content in 100 g <strong>of</strong> oil,<br />

which was calculated to be 8.90% using the following<br />

expression [24] :<br />

OOth = {(IVo / 2Ai) / [100 + (IVo / 2Ai) × Ao]}<br />

× Ao × 100 ----------------------------(2)<br />

where Ai (126.9) and Ao (16.0) are the atomic weights<br />

<strong>of</strong> iodine and oxygen respectively and IVo is the initial<br />

iodine value <strong>of</strong> the oil sample.<br />

RESULTS AND DISCUSSION<br />

The properties <strong>of</strong> WSO, with fatty acid composition,<br />

were analyzed as follows: specific gravity = 0.922 (at<br />

25 o C); acid value (mg KOH/g) = 2.2; iodine value (g<br />

I2/100 g) = 155; saponification value (mg KOH/g) =<br />

194; oleic acid, C 18: 1 = 13%; linoleic acid, C18:2 =<br />

73%.<br />

Effect <strong>of</strong> carboxylic acid<br />

In situ <strong>epoxidation</strong> reactions were carried out with<br />

both formic acid and acetic acid as carboxylic acids;<br />

with and without catalyst reagent with 1:0.5:1.5 molar<br />

ratio <strong>of</strong> ethylenic unsaturation: carboxylic acid:<br />

hydrogen peroxide at a temperature 60°C. The<br />

variations <strong>of</strong> iodine values with reaction time for<br />

<strong>epoxidation</strong> <strong>of</strong> WSO with acidic ion exchange resin<br />

IR-122 as catalyst is shown in Fig.2. The mechanism<br />

<strong>of</strong> double bond conversion to epoxy ring in<br />

[25]<br />

<strong>epoxidation</strong> reaction as proposed by Gan et.al. is<br />

as follows:<br />

The excess reagents can react with the epoxy rings to<br />

cause unwanted side reactions as depicted below [1] :<br />

From Fig.2 it can be shown that the relative conversion<br />

<strong>of</strong> ethylenic unsaturation is faster in presence <strong>of</strong> acetic<br />

acid than formic acid and addition <strong>of</strong> catalyst improves<br />

the rate <strong>of</strong> reduction <strong>of</strong> iodine value (IV). With formic<br />

acid/IR-122 combination, a slightly elevated reduction<br />

in IV is achieved as compared with the use <strong>of</strong> acetic<br />

acid with no resin. Fig.3 shows the relative percentage<br />

conversion to oxirane as function <strong>of</strong> time, from which<br />

it is inferred that formic acid/IR-122 combination is<br />

less effective than acetic acid and acetic acid/IR-122<br />

combination towards ethylenic unsaturation<br />

conversion to oxirane. The 65.4% conversion to<br />

oxirane was obtained for formic acid/IR-122 after 8<br />

hours while with acetic acid the conversion was 72%.<br />

Acetic acid/IR-122 combination resulted into 85%<br />

conversion to oxirane with corresponding iodine<br />

conversion <strong>of</strong> 87.2% was considered a better<br />

combination for <strong>epoxidation</strong> <strong>of</strong> WSO with aqueous<br />

hydrogen peroxide. Further experiments were,<br />

therefore, carried out with acetic acid/ IR-122 to study<br />

the effect <strong>of</strong> various parameters on conversions to<br />

oxirane oxygen and the conversion <strong>of</strong> iodine value.


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1156<br />

Figure 2: Variation <strong>of</strong> iodine value with time during <strong>epoxidation</strong> <strong>of</strong> WSO with acidic ion exchange resin IR-<br />

122 as catalyst.<br />

Figure 3: Activities <strong>of</strong> carboxylic acids in the <strong>epoxidation</strong> <strong>of</strong> WSO with acidic ion exchange resin IR-122 as<br />

catalyst. Conditions: Temperature, 60°C; carboxylic acid-to-ethylenic unsaturation molar ratio, 0.5:1;<br />

hydrogen peroxide-to-ethylenic unsaturation molar ratio, 1.5:1; catalyst loading, 20 wt.% <strong>of</strong> oil; stirring<br />

speed, 2000 rpm<br />

Effect <strong>of</strong> stirring speeds<br />

Agitation rate is an important parameter because this<br />

factor substantially affects the conversion to oxirane.<br />

Epoxidation was performed at different speeds varying<br />

from 1000-2500 rpm to study the effect <strong>of</strong> agitation<br />

rate on mass transfer resistance. In reaction, 35 g <strong>of</strong><br />

WSO was treated with 13.76 g <strong>of</strong> hydrogen peroxide,<br />

5.95g <strong>of</strong> acetic acid and 7g resin, as catalyst at 60°C.<br />

The % IV reductions and relative percentage<br />

conversions to oxirane at different speeds are shown<br />

inTable-1 and Fig.4 respectively. The oxirane<br />

formation rate improved with agitation rate upto 2000<br />

rpm but beyond that no significant control was<br />

observed under the given conditions. Thus, all<br />

experiments were performed at 2000 rpm assuming<br />

that the reaction is free from mass transfer resistance<br />

beyond 2000 rpm.


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1157<br />

Table 1: Relative conversion to oxirane at different stirring speeds<br />

Stirring Speeds, rpm<br />

1000 1500 2000 2500<br />

Time (h) 4 8 4 8 4 8 4 8<br />

Oxirane oxygen a (%) 6.42 6.87 7.00 7.23 7.53 7.70 7.62 7.40<br />

Iodine value b,c (g I2/100g) 41.2 32.9 21.2 19.0 22.6 19.8 24.5 20.6<br />

% Iodine value conversion d (A) 73.4 78.8 86.3 87.7 85.4 87.2 84.2 86.7<br />

a Experimentally determined oxirane oxygen content; b IV0,Initial iodine value=155;<br />

c IV Experimentally determined iodine value;<br />

d Conversion <strong>of</strong> ethylenic unsaturated bonds as related to IV0 calculated as A={(IV0 –IV) ×100/ IV0}<br />

Table 2: Effect <strong>of</strong> temperature on iodine value conversion<br />

Temperature ( o C)<br />

40 50 65 80<br />

Time (h) 4 8 4 8 4 8 4 8<br />

Oxirane oxygen a (%) 3.42 5.01 5.63 6.06 7.76 6.44 8.22 6.29<br />

Iodine Value b,c (g I2/100g) 93.6 66.0 55.4 48.0 18.9 16.9 15.2 12.2<br />

% Iodine conversion d (A) 39.6 57.5 64.3 69.1 87.8 89.1 90.19 92.13<br />

Figure 4: Effect <strong>of</strong> stirring speeds on the relative conversion to oxirane. Conditions: Temperature, 60°C; acetic<br />

acid to ethylenic unsaturation molar ratio, 0.5:1; hydrogen peroxide to ethylenic unsaturation molar ratio, 1.5:1;<br />

catalyst loading, 20 wt.% <strong>of</strong> oil<br />

Effect <strong>of</strong> temperature<br />

The effect <strong>of</strong> reaction temperature on the yield <strong>of</strong><br />

oxirane content was investigated (Table-2). Reactions<br />

at the molar ratio <strong>of</strong> 1:0.5:1.5 (ethylenic unsaturation:<br />

acetic acid: H2O2) were performed at five different<br />

temperatures, namely 40, 50, 60, 65 and 80 o C. Being<br />

exothermic in nature the <strong>epoxidation</strong> reaction may<br />

cause an excessive temperature rise and pose serious<br />

reaction control related issues like potential explosion<br />

hazards. In order to have better control over the<br />

reaction, the H2O2 was added to the mixture when the<br />

temperature was about 10 o C below the desired reaction<br />

temperature. The temperature increased thereafter to<br />

actual reaction temperature, and was maintained there<br />

until the completion <strong>of</strong> reaction. From Fig.5, it is<br />

evident that increase in the temperature favored<br />

<strong>epoxidation</strong> reaction and a maximum <strong>of</strong> 92.3%<br />

conversion <strong>of</strong> iodine value was obtained at 80 o C after a<br />

reaction time <strong>of</strong> 4 hours. The relative conversion to<br />

oxirane was found to increase with increase in<br />

temperature but the epoxy functionality was retained<br />

more at lower reaction temperature. The oxirane<br />

content <strong>of</strong> the EWSO, synthesized at 60 o C reached a<br />

maximum <strong>of</strong> 85.6 % after 6 hours <strong>of</strong> reaction and<br />

oxirane cleavage was very mild when the reaction<br />

continued further. At 65 o C, although the <strong>epoxidation</strong><br />

rate not improved much, but a sharp epoxy ring<br />

breaking was observed after 4 hours <strong>of</strong> continuation <strong>of</strong><br />

reaction. A similar trend appeared at higher reaction<br />

temperature and longer reaction durations. Therefore, a<br />

temperature <strong>of</strong> 60 o C was considered optimal for the<br />

<strong>epoxidation</strong> reaction under investigation.


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1158<br />

Table 3: Effect <strong>of</strong> catalyst loading on iodine value conversion<br />

AIER, IR122 loading (%)<br />

5% 10% 15% 20%<br />

Time (h) 1 4 1 4 1 4 1 4<br />

Oxirane oxygen a (%) 4.17 5.74 4.46 6.51 5.17 6.75 5.93 7.53<br />

Iodine Value b,c (g I2/100g) 76.8 51.2 72.1 40.8 62.3 32.5 49.5 22.6<br />

% Iodine value conversion d (A) 50.4 66.9 53.4 73.6 59.8 78.9 68.1 85.4<br />

Figure 5: Effect <strong>of</strong> temperature on the relative conversion to oxirane. Conditions: acetic acid to ethylenic<br />

unsaturation molar ratio, 0.5:1; hydrogen peroxide to ethylenic unsaturation molar ratio, 1.5:1; catalyst<br />

loading, 20 wt.% <strong>of</strong> oil; stirring speed, 2000 rpm<br />

Effect <strong>of</strong> catalyst loading<br />

In general, increasing the solid catalyst loading leads<br />

to increase in both the total active matter and the total<br />

surface area <strong>of</strong> the catalyst. The active moieties present<br />

in catalyst increases the rate <strong>of</strong> <strong>epoxidation</strong> with<br />

increase in the catalyst concentration in the reaction<br />

system. In our experimentation, IR-122 resin (16-50<br />

wet mesh size, macroreticular form) was used as<br />

catalyst and the effect <strong>of</strong> catalyst loading on the<br />

relative conversion to oxirane is presented in Fig.6.<br />

This figure shows that the conversion <strong>of</strong> ethylenic<br />

unsaturation to oxirane increases rapidly with catalyst<br />

dosage in the beginning and very slowly towards the<br />

end <strong>of</strong> the reaction. The relative conversion to oxirane<br />

increased from 64 to 75% with increasing dosage <strong>of</strong><br />

resin from 5 to 15% after 4 hours <strong>of</strong> reaction. With<br />

use <strong>of</strong> 10% resin a slight oxirane cleavage is observed<br />

after 3 hours <strong>of</strong> reaction and maximum conversion is<br />

72.9%, only 2 units less than that obtained with 15%<br />

catalyst loading. However, the maximum conversion<br />

<strong>of</strong> 84% was obtained with 20% catalyst loading under<br />

identical conditions and the product has stable oxirane<br />

ring. Therefore, 20% catalyst loading was considered<br />

to be optimum for the <strong>epoxidation</strong> reaction.<br />

Figure 6: Effect <strong>of</strong> catalyst loading on the relative<br />

conversion to oxirane. Conditions: Temperature,<br />

60°C; acetic acid to ethylenic unsaturation molar<br />

ratio, 0.5:1; hydrogen peroxide to ethylenic<br />

unsaturation molar ratio, 1.5:1; stirring speed, 2000<br />

rpm


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1159<br />

Table 4: Effect <strong>of</strong> acetic acid-to-ethylenic unsaturation molar ratio on iodine value conversion<br />

Acetic acid to ethylenic unsaturation molar ratio<br />

0.3 0.5 0.8 1.0<br />

Time (h) 4 8 4 8 4 8 4 8<br />

Oxirane oxygen a (%) 5.38 6.48 7.53 7.70 7.45 6.69 7.85 7.34<br />

Iodine Value b,c (g I2/100g) 57.5 32.8 22.6 19.8 21.3 15.1 18.6 14.3<br />

% Iodine conversion d (A) 62.9 78.8 85.4 87.2 86.2 90.3 88.0 90.8<br />

Figure 7: Effect <strong>of</strong> acetic acid to ethylenic unsaturation molar ratio on the relative conversion to oxirane.<br />

Conditions: Temperature, 60°C; hydrogen peroxide to ethylenic unsaturation molar ratio, 1.5:1; catalyst<br />

loading, 20 wt. % <strong>of</strong> oil; stirring speed, 2000 rpm<br />

Effect <strong>of</strong> acetic acid to ethylenic unsaturation molar<br />

ratio<br />

Optimization <strong>of</strong> acetic acid concentration was<br />

investigated at 0.3, 0.5, 0.8 and 1.0 mole per mole <strong>of</strong><br />

ethylenic unsaturation for rate conversion to oxirane<br />

with respect to 1.5 mole <strong>of</strong> H2O2 per mole <strong>of</strong><br />

unsaturation. Acetic acid acts as the oxygen carrier and<br />

gets regenerated once <strong>epoxidation</strong> reaction takes place.<br />

However, it can also act as reactant in hydrolysis <strong>of</strong> the<br />

oxirane ring formed. Therefore, an optimum level <strong>of</strong><br />

acetic acid is desired to speed up the conversion to<br />

oxirane ring with minimal epoxy ring breaking. The<br />

influence <strong>of</strong> acetic acid concentration on the relative<br />

conversion to oxirane is shown in Fig.7. Increasing<br />

acid concentration was found to increase the iodine<br />

value conversion (Table-4) and leads to more<br />

production <strong>of</strong> peracetic acid. The higher conversion<br />

rate <strong>of</strong> the double bond to epoxide was observed with<br />

increased mole ratio <strong>of</strong> acetic acid used. The reaction<br />

conversion <strong>of</strong> 73.7% was achieved within first hour<br />

when 1.0 mole acetic acid was used while the same<br />

conversion was obtained after 8 hour <strong>of</strong> reaction for<br />

0.3 mole. However, considering the stability <strong>of</strong> epoxy<br />

rings, the use <strong>of</strong> higher acid concentrations is <strong>of</strong> no<br />

significance as the excess acid promotes the hydrolysis<br />

<strong>of</strong> the epoxide. From Fig.7 it can be seen that an acidto-ethylenic<br />

unsaturation molar ratio <strong>of</strong> 0.5 is optimum<br />

with a moderately high rate <strong>of</strong> <strong>epoxidation</strong> reaction<br />

and negligible ring-opening reaction.<br />

Table 5: Effect <strong>of</strong> hydrogen peroxide-to-ethylenic unsaturation molar ratio on iodine value conversion<br />

Hydrogen peroxide to ethylenic unsaturation molar ratio<br />

0.5 1.0 1.5 2.0<br />

Time (h) 4 8 4 8 4 8 4 8<br />

Oxirane oxygen a (%) 3.89 4.19 5.07 6.16 7.53 7.70 8.03 6.61<br />

Iodine Value b,c (g I2/100g) 85.6 78.6 64.8 45.6 22.6 19.8 16.4 13.6<br />

% Iodine conversion d (A) 44.8 49.3 58.2 70.6 85.4 87.2 89.4 91.2


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1160<br />

Figure 8: Effect <strong>of</strong> the hydrogen peroxide to ethylenic unsaturation molar ratio on the relative conversion to<br />

oxirane. Conditions: Temperature, 60°C; acetic acid to ethylenic unsaturation molar ratio, 0.5:1; catalyst<br />

loading, 20 wt. % <strong>of</strong> oil; stirring speed, 2000 rpm<br />

Effect <strong>of</strong> hydrogen peroxide to ethylenic<br />

unsaturation molar ratio<br />

The effect <strong>of</strong> different mole ratios <strong>of</strong> H2O2-to-ethylenic<br />

unsaturation on rate <strong>of</strong> oxirane formation and iodine<br />

conversion was studied by carrying out a series <strong>of</strong><br />

reactions (Table-5). As seen in Fig.8 the reaction rate<br />

increased with an increase in concentration <strong>of</strong> H2O2 in<br />

the reaction. However; although the maximum<br />

conversion (91.2 %) <strong>of</strong> double bonds to epoxide rings<br />

was achieved with 2.0 mole <strong>of</strong> H2O2 concentration<br />

after 5 hours, it also tend to destroy the epoxy rings<br />

when reaction was continued further. For a lower H2O2<br />

dosing <strong>of</strong> 0.5 and 1.0 mole, the maximum oxirane<br />

level and initial <strong>epoxidation</strong> rates were significantly<br />

lower as compared to that obtained with use <strong>of</strong> 1.5<br />

mole H2O2. Moreover, the epoxy rings formed for<br />

1.5:1 mole ratio <strong>of</strong> H2O2 -to-ethylenic unsaturation<br />

shows excellent stability. Therefore, the optimal<br />

hydrogen peroxide-to-ethylenic unsaturation molar<br />

ratio was found to be 1.5 mol/mol, with a relative<br />

conversion to oxirane <strong>of</strong> 86.5% and iodine value<br />

conversion <strong>of</strong> 87.2% after 8 hours <strong>of</strong> reaction.<br />

Fourier transform infrared (FTIR) spectroscopy<br />

analysis<br />

Shimazdu FTIR-8400 spectrometer was used for<br />

monitoring the disappearance <strong>of</strong> double bonds and<br />

formation <strong>of</strong> epoxy groups during the reaction by<br />

qualitative identification <strong>of</strong> main signals. Only two<br />

spectra are analyzed: the raw oil, WSO and epoxide<br />

adduct, EWSO (Fig.9A & 9B, respectively), to show<br />

the changes in their functional groups. For WSO the<br />

characteristic peaks at 3008 cm -1 , 1650 cm -1 and<br />

721cm -1 are attributed to the stretching vibration <strong>of</strong> the<br />

double bonds: =C-H, C=C, cis-CH=CH, respectively.<br />

S. Hernandez-Lopez et.al. [26] reported the diminution<br />

<strong>of</strong> peak, C=C-H stretching at 3020 cm −1 after<br />

<strong>epoxidation</strong> reaction, which supports our study where<br />

the almost complete disappearance <strong>of</strong> double bonds<br />

band at 3008 cm −1 at 60 0 C after 8 h was observed. This<br />

confirms that almost all the C=C-H had taken part in<br />

the <strong>epoxidation</strong> reaction. Also, there is decrease in the<br />

intensity <strong>of</strong> the other important unsaturated bond<br />

signals in comparison with the unreacted oil, giving<br />

reliable support <strong>of</strong> its chemical transformation to an<br />

oxirane ring. The presence <strong>of</strong> new peaks in the FTIR<br />

spectrum <strong>of</strong> WSFO at 831 cm −1 , attributed to the<br />

epoxy group, confirmed the success <strong>of</strong> the <strong>epoxidation</strong><br />

reaction <strong>of</strong> WSO. Vleck and Petrovic [27] reported the<br />

presence <strong>of</strong> epoxy groups at 822–833 cm −1 , which<br />

agrees well with this study. The other new peak at the<br />

3470 cm −1 was attributed to the hydroxyl functional<br />

group, derived from the epoxy functional group via<br />

partial epoxy ring opening reaction. The intensity <strong>of</strong><br />

the 3470 cm −1 band indicated the extent <strong>of</strong> hydrolysis<br />

<strong>of</strong> EWSO. The epoxy ring opening reaction could<br />

occur either by acid catalysis in the presence <strong>of</strong> water<br />

associated with aqueous solution <strong>of</strong> H2O2 used [28] .<br />

The hydrolysis <strong>of</strong> the ester groups during <strong>epoxidation</strong><br />

reaction in oils is the main side reaction. The band<br />

corresponding to carboxylic acid group is located at<br />

1650 cm -1 and is usually very intense even at low<br />

carboxylic group concentration [29, 30] . In case <strong>of</strong><br />

hydrolysis, a carboxylic acid functional group is<br />

formed and this carbonyl group will appear near but<br />

differentiable <strong>of</strong> the ester carbonyl stretching C=O in<br />

the glyceride moiety at 1744 cm -1 . However; in the<br />

course <strong>of</strong> the <strong>epoxidation</strong> reaction carried out in our


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1161<br />

study, no evidence <strong>of</strong> the carbonyl from carboxylic acid group signal was observed.<br />

Figure 9: FTIR spectra <strong>of</strong> (A) <strong>wild</strong> <strong>safflower</strong> oil and (B) and epoxidised <strong>wild</strong> <strong>safflower</strong> oil. Epoxidation<br />

Conditions: Temperature, 60 o C, carboxylic acid (acetic acid)-to-ethylenic unsaturation molar ratio, 0.5 :<br />

1;hydrogen peroxide-to-ethylenic unsaturation molar ratio, 1.5 : 1; catalyst loading, 20 wt. % <strong>of</strong> oil; reaction<br />

time, 8 h; stirring speed, 2000 rpm<br />

CONCLUSION<br />

The results <strong>of</strong> the present investigation shows that<br />

WSO can be successfully utilized for <strong>epoxidation</strong><br />

using peroxy acid generated in situ. Acetic acid–IR-<br />

122 combination was found to be most effective for<br />

higher epoxide yield at shorter reaction time. The<br />

optimized parameters to get higher degree <strong>of</strong><br />

<strong>epoxidation</strong> with minimum epoxy ring breaking were<br />

noted as temperature <strong>of</strong> 60°C, stirring speed 2000 rpm<br />

ensuring kinetically control <strong>of</strong> reaction, an acetic acid<br />

to ethylenic unsaturation molar ratio <strong>of</strong> 0.5:1, a<br />

hydrogen peroxide to ethylenic unsaturation molar<br />

ratio <strong>of</strong> 1.5:1, and a catalyst (IR-122) loading <strong>of</strong> 20<br />

wt% <strong>of</strong> WSO. The completion <strong>of</strong> reaction was<br />

supported by FTIR spectral analysis. The epoxy<br />

signals were well identified and intensity <strong>of</strong> the signals<br />

due to the double bonds vibrations was evidenced by<br />

the spectroscopic techniques used. Under these<br />

optimum conditions, 7.87% oxirane oxygen content in<br />

synthesized ESWO was obtained. The synthesized<br />

epoxide WSO is an attractive intermediate for the<br />

preparation <strong>of</strong> various derivatives <strong>of</strong> industrial<br />

importance.


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1162<br />

REFERENCES<br />

1. Le P. L., Wan Yunus W.M.Z., Yeong S. K.,<br />

Dezulkelfly K. A., Lim W. H., “Optimization <strong>of</strong><br />

the <strong>epoxidation</strong> <strong>of</strong> methyl ester <strong>of</strong> palm fatty acid<br />

distillate” J. Oil Palm Res., 2009, 21, 675-682.<br />

2. Holser R. A., “Transesterification <strong>of</strong> epoxidized<br />

soybean oil to prepare epoxy methyl esters”, Ind.<br />

Crops and Products, 2008, 27, 130-132.<br />

3. Goud V.V., Mungroo R., Pradhan N.C., and Dalai<br />

A.K., “Modification <strong>of</strong> epoxidised canola oil”,<br />

Asia-Pacific J. Chem. Eng., 2011, 6, 14–22.<br />

4. Dahlke B., Hellbardt S., Paetow M., Zech W.H.,<br />

“Polyhydroxy fatty acids and their derivatives<br />

from plant oils”, J. Am. Oil Chem. Soc., 1995, 72,<br />

349–353.<br />

5. Okieimen F.E, Pavithran C., Bakare I.O.,<br />

“Epoxidation and hydroxylation <strong>of</strong> rubber seed oil:<br />

one-pot multi step reactions”, Euro. J. Lipid Sci. &<br />

Tech., 2005, 107, 864-870.<br />

6. Patwardhan A.V., Goud V.V., Pradhan N.C.,<br />

“Epoxidation <strong>of</strong> Karanja (Pongamia glabra) Oil by<br />

H2O2”, J. Am. Oil Chem. Soc., 2006, 83, 247-252.<br />

7. Carlson K.D., Kleiman R., Bagby M.O,<br />

“Epoxidation <strong>of</strong> lesquerella and limmanthes<br />

(meadowfoam ) oil” , J. Am. Oil Chem. Soc.,<br />

1994, 71, 175-182.<br />

8. Marcel S.F., Lei K. J. and Mohammad K.P.,<br />

“Epoxidation reactions <strong>of</strong> unsaturated fatty esters<br />

with peroxomonosulphate”, Lipids, 1998, 33, 633-<br />

637.<br />

9. Sonnet P. E. and Foglia T., “Epoxidation <strong>of</strong> natural<br />

triglycerides with ethyl methydioxyrane”,<br />

J. Am. Oil Chem. Soc., 83, 835-840.<br />

10. Aerts A. J., Jacob P.A., “Epoxide yield<br />

determination <strong>of</strong> oils and fatty acid methyl esters<br />

using 1 H NMR”, J. Am. Oil Chem. Soc., 2004, 81,<br />

841-846<br />

11. Gurbanov M. S., and Mamedov B. A.,<br />

“Epoxidation <strong>of</strong> Flax Oil with Hydrogen Peroxide<br />

in a Conjugate System in the presence <strong>of</strong> Acetic<br />

Acid and Chlorinated Cation Exchanger KU-2×8<br />

as Catalyst”, Russian J. App. Chem., 2009, 82,<br />

1483-1487.<br />

12. Khot S. N., Lascala J. J., Can E., Moyre S. S.,<br />

Williams, G. I., Palmese ,G. R., Kusefoglu, S. H.<br />

and Wool, R.P., “Development and Application <strong>of</strong><br />

Triglyceride-Based Polymers and Composites”, J.<br />

Appl. Polym. Sci., 2001, 82, 703-723.<br />

13. Biermann U., Friedt W., Lang S., Luhs W.,<br />

Machmuller G., Metzger J. O., Klaas M. R.,<br />

Schafer H.J. and Schneider M.P., “New Synthesis<br />

with Oils and Fats as Renewable Materials for the<br />

Chemical Industry”, Agew. Chem. Int. Ed., 2000,<br />

39, 2206-2224.<br />

14. Muturi P., Wang D., and Dirlikov S., “Epoxidized<br />

vegetable oils as reactive diluents. I. Comparison<br />

<strong>of</strong> vernonia, epoxidized soybean and epoxidized<br />

linseed oil”, Progress in Organic Coatings, 1994,<br />

25, 85-94.<br />

15. Adhvaryu A., Erhan S.Z., “Epoxidized soybean oil<br />

as a potential source <strong>of</strong> high-temperature<br />

lubricants”, Industrial Crops and Products, 2002,<br />

15, 247-254.<br />

16. Soucek M. D., Johnson A. H., Wegner J. M.,<br />

“Ternary evaluation <strong>of</strong> UV-curable seed oil<br />

inorganic/organic hybrid coatings using<br />

experimental design” Progress in Organic<br />

Coatings, 2004, 51, 300-311.<br />

17. Zhang L.P., “Safflower: a versatile plant”, IV<br />

International Safflower Conf., Bari, 2-7 June,<br />

1997, 311-329.<br />

18. Hilker D., Bothe J. P. and Warnecke H.J.,<br />

“Chemo-Enzymatic Epoxidation <strong>of</strong> Unsaturated<br />

Plants Oils”, Chemical Engineering Science, 2001,<br />

56, 427-432<br />

19. Anjani K., Mukta N. and Lakshmamma P., “Crop<br />

improvement: In <strong>Research</strong> Achievements in<br />

Safflower”, Directorate <strong>of</strong> Oilseeds <strong>Research</strong>,<br />

Hyderabad, India, 11.<br />

20. Ravikumar R.L., Soregaon C.D. and Satish D.,<br />

“Molecular diversity analysis <strong>of</strong> five different<br />

species <strong>of</strong> genus Carthamus”, National Seminar on<br />

Changing Global Vegetable Oils Scenario: Issues<br />

and Challenges before India”, Directorate <strong>of</strong><br />

Oilseeds <strong>Research</strong>, Hyderabad, India, 29-31<br />

January, 2007, 2-4.<br />

21. Ashri A., Zimmer D.E., Urie A.L., Cahaner A. and<br />

Marani A., “Evaluation <strong>of</strong> world collection <strong>of</strong><br />

<strong>safflower</strong> Carthamus tinctorius L. yield and yield<br />

components and their relationships”, Crop Sci.,<br />

1974, 14, 799-802.<br />

22. Banergy R., Pandey V. and Dixit B. S., “Wild<br />

Safflower: An alternative source <strong>of</strong> <strong>safflower</strong> oil”,<br />

J. Oil Tech. Assoc. Ind., 1999, 31, 59-60.<br />

23. Dalai A.K., Mungroo R., Pradhan N.C., Goud<br />

V.V. ,“ Epoxidation <strong>of</strong> canola oil with hydrogen<br />

peroxide catalyzed by acidic ion exchange resin”,<br />

J. Am. Oil Chem. Soc.,2008, 85, 887-896.<br />

24. Petrovic Z. S., Zlatanic A., Lava C.C.,<br />

Sinadinovic-Fiser S., Epoxidation <strong>of</strong> soybean oil in<br />

toluene with peroxoacetic and perox<strong>of</strong>ormic acids.<br />

Kinetics and side reactions”, Eur. J. Lipid Sci. &<br />

Tech., 2002, 104, 293–299.


Pawan D. Meshram et al /Int.J. ChemTech Res.2011,3(3) 1163<br />

25. Gan L.H., Goh S.H., and Ooi K.S., “Kinetics<br />

studies <strong>of</strong> <strong>epoxidation</strong> and oxirane cleavage <strong>of</strong><br />

palmolein methyl esters”, J. Am. Oil Chem. Soc.,<br />

1992, 69, 347-351.<br />

26. S. Hernandez-Lopez, E. Martin del Campo-Lopez,<br />

V. Sanchez-Mendieta, F. Urena-Nunez and E.<br />

Vigueras- Santiago, “Gamma Irradiation on<br />

Acrylated-Expoxidized Soybean Oil:<br />

Polymerization and Characterization”, Adv. in<br />

Tech. <strong>of</strong> Mat. and Mat. Proc. J., 2006, 8, 220-225.<br />

27. Vlcek T., Petrovic Z.S., “Optimization <strong>of</strong> the<br />

chemoenzymatic <strong>epoxidation</strong> <strong>of</strong> soybean oil”, J.<br />

Am. Oil Chem. Soc., 2006, 83,247–252<br />

28. Pares X.P.X., Bonnet C., and Morin O., “Synthesis<br />

<strong>of</strong> New Derivatives from Vegetable Oil Methyl<br />

*****<br />

Esters via Epoxidation and Oxirane Opening”, in<br />

Recent Developments in the Synthesis <strong>of</strong> Fatty<br />

Acid Derivative (Knothe, G. and Derksen, J.T.P.,<br />

Eds.) AOCS Press, 1999, Champaign, IL, 141-156.<br />

29. G. Lopez Tellez, E. Vigueras-Santiago, S.<br />

Hernandez-Lopez, “ Characterization <strong>of</strong><br />

epoxidized linseed oil at different percentage”,<br />

Surface and Emptiness, 2009, 22, 5-10.<br />

30. G. Lopez-Tellez, E. Vigueras-Santiago, S.<br />

Hernandez-Lopez, B. Bilyeu, “Synthesis and<br />

thermal cross linking study <strong>of</strong> partially aminated<br />

epoxidized linseed oil”, Designed Monomers and<br />

Polymers, 2008, 11, 435-445.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!