rosin: chemistry, derivatives, and applications: a review · of terpene and terpenoid derivatives...

16
BAOJ Chemistry Manar El-Sayed Abdel-Raouf, et al. BAOJ Chem 2018, 4: 1 4: 039 BAOJ Chem, an open access journal Volume 4; Issue 1; 039 Research Rosin: Chemistry, Derivatives, and Applications: a review Manar El-Sayed Abdel-Raouf 1* and Abdul-Raheim Mahmoud Abdul-Raheim 2 1 Egypan Petroleum research instute, Nasr city, Cairo, Egypt *Corresponding Author: Manar El-Sayed Abdel-Raouf, Egypan Petroleum research instute, Nasr city, Cairo, Egypt, E-mail: drmanar770 @yahoo.com Sub Date: August 09 th 2018, Acc Date: August 09 th 2018, Pub Date: August 29 th 2018 Citaon: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul- Raheim (2018) Rosin: Chemistry, Derivaves, and Applicaons: a review. BAOJ Chem 4: 039 Copyright: © 2018 Manar El-Sayed Abdel-Raouf, et al. This is an open access arcle distributed under the terms of the Creave Commons Aribuon License, which permits unrestricted use, distribuon, and reproducon in any medium, provided the original author and source are credited. Abstract Rosin (colophony) is a thermoplastic solid resin extracted from the trees of Pinus species as clear, pale yellow to dark amber. e most important ingredients of rosin include resin acids. ey have the gen- eral formula C 20 H 30 O 2 which represents free acids or dimers or anhy- drides. Abietic acid constitutes the major resin acid. e chemistry, grades, and chemical reactions of rosin are reviewed in some details. e toxicity studies on rosin acids or their derivatives have proved them be practically non-toxic. Rosin derivatives have been investigat- ed for their value as pharmaceutical purposes such as in film coating, matrix formation, drug delivery systems and microencapsulation. ey find great importance in many industrial fields such as ink in- dustry, coating industry, epoxy curing agents and as adhesives. Fur- thermore, they can be utilized as reactive monomer or co-monomers in a diversity of polymerization systems. Rosin and its derivatives are oſten used as polymer additives in different capacities. ey play an important role as plasticizers and having specific importance in pa- per industry. In petroleum field, rosin derivatives find applications as emulsifiers due to their amphiphilic nature and as corrosion inhibi- tors. ey can be also used as oil spill dispersants. Some future trends for rosin derivatives were also discussed. Keywords: Rosin; Paper Industry; Surfactants; Drug Delivery Introduction Rosin is extracted from various wood species in very small percent- age, i.e. only about 2-5%. e other constituents are generally: 35- 45% cellulose, 25-35% lignin, and 20-30% hemicellulose. is gener- al composition varies widely among different wood species and also within different parts of the same tree. Rosin extractives consist pri- marily of rosin, fatty acids, phenolic compounds and a large variation of terpene and terpenoid derivatives [1,2]. Moreover, the extractives content changes aſter the death of a living tree due to chemical chang- es [3]. e non-volatile residue is known as Rosin (colophony). e major source of rosin is pine trees (Pinus genus), not only because they are well-known in the Northern Hemisphere, but also because of their rigorous use in the timber and pulp industries. In the past, Pine resins were used in the waterproofing of wooden ships thus it was known as “Naval Stores” [4]. Generally, there are three main methods for extracting rosin commer- cially [4]: (1) Distillation of volatile turpentine from oleoresin exuded from the wound of living pine trees to obtain gum rosin (gum naval stores); (2) Solvent extraction of pine stump wood along with the removal of the turpentine by steam distillation to obtain wood rosin (Wood rosin); and (3) Separation of tall oil to get tall oil rosin. As tall oil rosin is the major source of rosin, different new facilities were planned to recover and fractionate practically all of the crude tall oil in the pulp industry since 1949 [3-6]. e name ‘tall oil’ is originat- ed from Swedish word ‘tall’ which means pine. e total world rosin production obviously declined between the 1960s and the 1980s. Fi- nally, the rosin production stabilized at around 1.0–1.2 million tons per year [7,8]. Based on the 1994 survey [8], the rosin production is divided as following: 60% gum rosin, 35% tall oil rosin and wood rosin to only a few percent. e progress of gum rosin production has been governed by two distinctive factors:

Upload: vanhanh

Post on 26-Apr-2019

440 views

Category:

Documents


18 download

TRANSCRIPT

Page 1: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chemistry

Manar El-Sayed Abdel-Raouf, et al. BAOJ Chem 2018, 4: 1

4: 039

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Research

Rosin: Chemistry, Derivatives, and Applications: a reviewManar El-Sayed Abdel-Raouf1* and Abdul-Raheim Mahmoud Abdul-Raheim2

1Egyptian Petroleum research institute, Nasr city, Cairo, Egypt

*Corresponding Author: Manar El-Sayed Abdel-Raouf, Egyptian Petroleum

research institute, Nasr city, Cairo, Egypt, E-mail: drmanar770 @yahoo.com

Sub Date: August 09th 2018, Acc Date: August 09th 2018, Pub Date:

August 29th 2018

Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-

Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a review.

BAOJ Chem 4: 039

Copyright: © 2018 Manar El-Sayed Abdel-Raouf, et al. This is an open

access article distributed under the terms of the Creative Commons

Attribution License, which permits unrestricted use, distribution, and

reproduction in any medium, provided the original author and source

are credited.

Abstract

Rosin (colophony) is a thermoplastic solid resin extracted from the trees of Pinus species as clear, pale yellow to dark amber. The most important ingredients of rosin include resin acids. They have the gen-eral formula C20H30O2 which represents free acids or dimers or anhy-drides. Abietic acid constitutes the major resin acid. The chemistry, grades, and chemical reactions of rosin are reviewed in some details. The toxicity studies on rosin acids or their derivatives have proved them be practically non-toxic. Rosin derivatives have been investigat-ed for their value as pharmaceutical purposes such as in film coating, matrix formation, drug delivery systems and microencapsulation. They find great importance in many industrial fields such as ink in-dustry, coating industry, epoxy curing agents and as adhesives. Fur-thermore, they can be utilized as reactive monomer or co-monomers in a diversity of polymerization systems. Rosin and its derivatives are often used as polymer additives in different capacities. They play an important role as plasticizers and having specific importance in pa-per industry. In petroleum field, rosin derivatives find applications as emulsifiers due to their amphiphilic nature and as corrosion inhibi-tors. They can be also used as oil spill dispersants. Some future trends for rosin derivatives were also discussed.

Keywords: Rosin; Paper Industry; Surfactants; Drug Delivery

Introduction

Rosin is extracted from various wood species in very small percent-age, i.e. only about 2-5%. The other constituents are generally: 35-45% cellulose, 25-35% lignin, and 20-30% hemicellulose. This gener-al composition varies widely among different wood species and also within different parts of the same tree. Rosin extractives consist pri-marily of rosin, fatty acids, phenolic compounds and a large variation of terpene and terpenoid derivatives [1,2]. Moreover, the extractives content changes after the death of a living tree due to chemical chang-es [3].

The non-volatile residue is known as Rosin (colophony). The major source of rosin is pine trees (Pinus genus), not only because they are well-known in the Northern Hemisphere, but also because of their

rigorous use in the timber and pulp industries. In the past, Pine resins were used in the waterproofing of wooden ships thus it was known as “Naval Stores” [4].

Generally, there are three main methods for extracting rosin commer-cially [4]:

(1) Distillation of volatile turpentine from oleoresin exuded from the wound of living pine trees to obtain gum rosin (gum naval stores);

(2) Solvent extraction of pine stump wood along with the removal of the turpentine by steam distillation to obtain wood rosin (Wood rosin); and

(3) Separation of tall oil to get tall oil rosin.

As tall oil rosin is the major source of rosin, different new facilities were planned to recover and fractionate practically all of the crude tall oil in the pulp industry since 1949 [3-6]. The name ‘tall oil’ is originat-ed from Swedish word ‘tall’ which means pine. The total world rosin production obviously declined between the 1960s and the 1980s. Fi-nally, the rosin production stabilized at around 1.0–1.2 million tons per year [7,8]. Based on the 1994 survey [8], the rosin production is divided as following: 60% gum rosin, 35% tall oil rosin and wood rosin to only a few percent. The progress of gum rosin production has been governed by two distinctive factors:

Page 2: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 2 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

• Theintensecompetitionofpetroleum-basedcounterpartsand

• Thegrowingcostsofman-powerinvolvedintreetappingandresinrecovery.

Due to great industrial revolution, the competition between petro-leum-based products with natural products was quite strong in the second half of the twentieth century. There are many challenges we face in the twenty-first century, such as that we are becoming increas-ingly dependent on the use of fossil fuels despite the continued de-crease [9] in global oil reserves. On the other hand, the uncontrolled utilization of fossil fuels is a major contributor to global warming [10]. Furthermore, It has been estimated that if present trends in green-house-gas emissions continue, the earth could be subjected to an in-crease earth’s temperature by about 1.5-4.50C by the middle of the next century [10]. Reduction of world consumption of fossil fuels by 50% or more over the next several decades may prevent the global warming [10]. Thus, the growing costs of fossil resources will encourage nov-el fundamental and applied research on the development of rosin as a source of chemicals and materials. Elevated man-power costs have contributed to the turn down of gum rosin production in the indus-trialized North-American and European countries while shifting it to other areas of the world such as China and Indonesia, which in the mid-1990s were already responsible for 60 and 10 per cent of world production, respectively [4,5].

Rosin Resins

Three categories of rosin are utilized for resin manufacture according to the method of extraction, namely, gum rosin, wood rosin and tall oil rosin, all generated from the pine tree.

a- Gum rosin was once the lone commercial source of rosin. It is the oleoresin (pine gum) of the living pine tree. The harvesting of the oleoresin is simple, involving only periodic cutting of the tree and collecting of the exudates into cups.

b- Wood rosin: After harvesting pine trees the base is allowed to stay in the ground for about ten years so that its bark and sapwood may decay and slough off to leave the heartwood rich in resin.

c- Resinous material is extracted from the stub and known as tall oil

rosin. It is obtained by distillation of crude tall oil (CTO). CTO contains about 70-90% acidic material, which is composed basical-ly of fatty acid and tall oil rosin. Highly distilled Tall oil rosin can produce esters which are competitive with gum and wood rosin derivatives.

Rosin Chemistry

In spite of its origin, rosin is mainly composed 90–95 % of resin acids. They are diterpenic monocarboxylic acids having the general formula C19H29COOH. The residual components are mainly made up of neu-tral compounds, the nature of which depends on the specific origin of the rosin [5]. The most common resin acids found in pine rosin are derived from the three basic tricyclic carbon skeletons abietane, pima-rane and isopimarane and the less common bicyclic labdane skeleton (Figure 1).

Figure 1: Terpenoids found in most common resin acids

Four resin acids commonly have the abietane skeleton, thus they are known as abietadienoic acids (Figure2). These are abietic, neoabietic, palustric and levopimaric acids. They differ only in the position of the conjugated double bond system, which is an important characteristic of this group of resin acids, because it affects their chemical reactivity and consequently the applications of the resultant products, as dis-cussed below. The aromatic dehydroabietic acid is also found in small and variable amounts in various rosin species. The most common pimarane-type acids are pimaric, isopimaric and sandaracopimaric acids (Figure3).

Upon comparison between the two types, two major differences are noticed. The basic skeleton different and more importantly, the double bond system is now not conjugated (Table 1), a fact which reduces significantly the possible chemical management of these compounds. Although abietic-type acids are the principal structures in most ros-in, their relative abundance is quite variable, depending on the pine species and geographic origin. Furthermore, processing and handling conditions (e.g. temperature and pH) can induce the isomerization of the double bond system or result in Disproportionation process, lead-ing to equilibrium mixtures, as will be discussed later (Figure4). From an industrial perspective, the quality of rosin and rosin derivatives is assessed on the basis of four basic parameters [4,6], namely:

Page 3: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 3 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

(1) The acid number, which is a measure of the amount of the free carboxylic groups; a decrease in this value is indicative of decarboxyl-ation and/or functionalization of the carboxylic moieties.

(2) The saponification number, which is a measure of the total amount of carboxylic groups; a decrease in this value is indicative of resin acid decarboxylation.

(3) The color, whose intensity is a key detrimental factor in many ap-plications, is a measure of rosin oxidation; an increase in color inten-sity is therefore an indication of decreasing quality.

(4) The softening point, which is in fact a measure of the glass tran-sition temperature associated with these complex mixtures of glassy materials; its value strongly influences the possible applications of these resins, as discussed below.

Isomers Double bond position

Abietic acid C-7, C-8 and C-9, C-14

Levopimaric acid C-6, C-7 and C-8, C-14

Palustric acid C-8, C-9 and C-13, C-14

Neoabietic acid C-8, C-14 and C-7, C-18

Dehydroabietic acid Aromatic in the ring bearing isopropyl group.

Dihydroabietic acid One double bond among C-9, C-14, C-13 and C-8

Tetrahydrobeitic acid None

Table 1: Structure of major rosin acid isomers

Figure 2: Chemical Structures of the most common abietane-type resin acids

Figure 3: Chemical Structures of the most common pimarane-type resin

Figure 4: Disproportionation of Rosin Acids

Rosin Grades

Gum rosin is differentiated into diverse grades according to the color and the softening point of the rosin. Softening point is the temperature at which rosin may distort under pressure. Generally, as the soften-ing point of the rosin reduces as the rosin color gets darker. Table (1) shows some of the rosin grades and relative properties.

Item X WW WG N M K

Color

slightly yellow

pale yellow yellow deep

yellowyellow brown

Yellow red

Correspond to the standard glass color piece of rosin

Appearance transparent solid

Softening Point(ball &

ring)76oC min 75oC min 74oC min

Acid No. ( mg KOH/g ) 166min 165min 164min

Unsaponifiable Material 5% max 6%max

Alcohol insoluble Material

0.03%max 0.04%max

Ash ( % ), max 0.02 0.03 0.04

Table 2: Rosin Grades and Relative properties

Page 4: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 4 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

Figure 5: Rosins of different grades

Reactions of Rosin Acids

The chemical reactivity of resin acids is based on the presence of both the double- bond system and the COOH group [5]. The overall reac-tions of both function groups are given in Figure 6.

Reactions of the Carboxylic Group

Salt Formation

Numerous metal resin acid salts are known such as the salts of Na, Mg, Ca, Zn, Al and ammonium resin acid salts. These salts can be produced industrially [5]. Na, K and ammonium salts are partially water soluble and were used as soaps in the past. Recently, resin acids sodium salts are mainly used as intermediates in paper sizing.

Esterification of Resin Acids

The esterification of resin acids is performed industrially at high tem-peratures (260–300°C) in presence of metal oxides as catalysts. The most common commercially available esters are those with methanol and polyols like ethylene glycol [11], pentaerythritol [12], diethylene glycol, and glycerol [13]. Glycerol esters were the first to be used in protective coatings, although pentaerythritol esters are harder and more durable for varnishes. The methyl esters are generally used as plasticizing agents, as it will be mentioned later.

Alkoxylation Reactions

The alkoxylation reaction of free resin acids takes place in presence of certain reagents such as ethylene oxide. It yields hydroxyl-terminat-ed esters which can easily undergo polymerization with an excess of ethylene oxide to produce polyethers of different chain lengths [14]. These rosin derivatives can be utilized as chain extenders in the poly-urethane foam industry [14,15].

Anhydride and Nitride Formation

Anhydrides can be prepared by refluxing the analogous acid in acetic anhydride [5]. These derivatives have found modest applications as such e.g. in paper sizing, but they are also useful as precursors to ni-trogen derivatives [5]. Furthermore, the reaction of molten rosin with ammonia at high temperature yields the corresponding nitrils which can be easily converted into Rosin amines [5]:

RCOOH+NH3→ [RCONH2] → RCN → RCH2NH2

Dehydroabietylamine (Figure7) and the analogous ammonium salts find a wide range of applications from cationic flotation agents to anti-oxidants, fungicides and anticorrosion materials [5]. The reduction of methyl dehydroabietate at high temperature yields dehydroabietanol (Figure7), whose light colour and high stability make it an appropriate intermediate for the synthesis of a variety of esters used in protective coatings, adhesives, and plasticizers [5].

Page 5: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 5 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

Figure 6: The Overall reactions of Rosin acids

Figure 7: Structures of dehydroabietyl amine and dehydroabietanol

Reactions with the Olefin System

Although the conjugated double bond system of abietadienoic acids is an attractive reactive centre for additional modifications although it is a source of instability.

Oxidation Reaction

The conjugated double bond is responsible for the change of rosin color due to oxidation, isomerization, and other reactions. In many applications this colouring reflects the loss of product quality. Oxi-dation reaction is normally photochemically induced, leading to the formation of epoxides, hydroxylated derivatives, and endoperoxides

[16] (Figure8). These subsidiary reactions can be concealed by dehy-drogenation or hydrogenation processes.

Figure 8: SOxidation of levopimaric acid with formation of an endoperoxide

Hydrogenation and Dehydrogenation Reactions

Hydrogenation is an effective methodology for rosin stabilization. The reduction of the first (conjugated) double bond of abietadienoic acids with hydrogen in the presence of metal catalysts is relatively uncom-plicated and frequently the resulting dihydro-derivatives are stable enough for most applications. The reduction of the remaining double bond needs harder conditions

Dehydrogenation reaction results in elimination of hydrogen. This leads to converting abietadienoic acids into dehydroabietic acid after

Page 6: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 6 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

double bond rearrangement [5,8,16]. This reaction is normally taken place at high temperatures (200-300°C), in the presence of metal cat-alysts (e.g. Pd, Ni), sulphur or iodine. The eliminated hydrogen then added to the pimaradienoic acids and to the abietadienoic acids. Both reactions result in the rosin stabilization and the stabilized products

are known as disproportionated rosins. Decarboxylation reaction may be carried out if dehydrogenation is performed at extreme tempera-tures for long reaction times to produce the neutral aromatic com-pound (Figure9) [5].

Figure 9: Catalytic conversion of abietadienoic acid(s) into dehydroabietic acid and retene

Functionalization of Dehydroabietic Acid Aromatic Ring

Dehydroabietic acid can undergo classic aromatic substitution reac-tions (e.g. acylation, chlorosulphonation, sulphonation and nitration)

with preferential functionalization of the more reactive 12 position, followed in some cases by the 14 position [13,18]. For brevity, the ni-tration reaction of dehydroabietic acid is given in (Figure10).

Figure 10: Nitration of dehydroabietic acid

Isomerization

Abietic-type acids can undergo thermal isomerization through the conjugated double bond system [18]. Thus, for instance, isomerization of abietadienoic acid mixtures gives an equilibrium mixture of prod-ucts with the following constitution: abietic acid ( 80 %), palustric ( 14 %) and neoabietic acids ( 5 %), whereas levopimaric acid is only formed in trace amounts.

Diels-Alder Reaction

Diels-Alder reaction with dienophiles is used to convert resin acids

into novel materials. This reaction is optimized by the simultaneous isomerization process giving levopimaric acid, which is the major isomer able to give a Diels-Alder adduct. It was shown that palus-tric acid can produce Diels-Alder adduct in some instances [15]. The Diels-Alder adduct between levopimaric acid and maleic anhydride, maleopimaric anhydride (Fig. 11) and the corresponding diacid are definitely the most significant derivatives of this family due to their wide applications. Levopimaric acid can also be utilized in the prepa-ration of other adducts with a wide variety of dienophiles such as fu-maric acid, acrylonitrile, acrylic acid and vinyl acetate [5].

Page 7: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 7 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

Figure 11: Diels-Alder reaction of levopimaric acid with maleic anhydride

Reactions with Formaldehyde and Phenol

The addition reaction of resin acids to formaldehyde and/or phenols can be achieved either under alkaline or acidic media. Since rosin can react with formaldehyde in a similar way as phenol, rosin components

have also been used for preparation of the rosin-phenol-formaldehyde resins for printing ink formulations. The inclusion of rosin compo-nents into phenol-formaldehyde prepolymers can take place through esterification or methylol condensation at one of the unsaturated car-bons (Figure12).

Figure 12: Formation of rosin-modified phenol formaldehyde resins

Rosin Derivatives

The utilization of Rosin and its derivatives as monomers or co-mono-mers in a variety of polymerization systems is widely known [16-18]. For instance, Maleopimaric anhydride can interact with amines to form maleimides. If diamines are used, the corresponding maleim-ide-amines can be polymerized through a step process (Figure13), in-volving the condensation of the carboxylic moiety with the primary amine to form a poly (amide-imide), which was used in the formula-tion of gravure printing inks [17]. These poly (amide-imide) materials were used in synthesis and preparation of blends with other polymeric materials [18] reflecting the increased importance of these materials.

Moreover, water soluble polyamides and poly (amide-imide)s have been recently reported [19]. In addition, photoactive polymers could be prepared by the condensation of a maleopimaric adduct with azo-dye type diamines [20]. The levopimaric adduct with acrylic acid has been used as a diacid in the synthesis of polyesters [16] The two car-boxylic groups of this diacid were transformed into isocyanate func-tions and used as a co-monomer in the preparation of polyurethanes [21].

Furthermore, some metal coatings have been prepared from ma-leopimaric acid acrylopimaric adducts, after a three-step reaction with ethylene glycol, epichlorohydrine and acrylic or methacrylic acid con-

Page 8: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 8 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

sequently [22]. Moreover, maleopimaric anhydride can be converted into diallyl derivatives [23], which can be further converted into di-

meric-type ketones [24]. Both intermediates were used in polymeriza-tion tests that yielded crosslinked materials [23, 24].

Figure 13: Synthesis and polycondensation of rosin-based poly (amide-imide)

Some Applications of Rosin Derivatives

Rosin utilization, a part of the so-called Naval Stores Industry, is as old as the construction of wooden naval vessels. However, the chemistry of this natural source has started only during the first half of the twen-tieth century. Since that, new transformations and applications were developed on a more scientific basis. Rosin is a source of monomers or additives for polymeric materials. A detailed investigation of the extensive applications of rosin and its derivatives will be introduced.

Rosin Derivatives in Medical Field

Rosin Derivatives for Controlled Drug Delivery

Natural products have attracted widespread attention for drug deliv-ery applications due to their accessibility, capability of different chem-ical modifications, environmental compatibility, and biodegradtion. The structures of natural materials are always multifarious than syn-thetic ones. This makes their chemical accessibility more complicat-ed.. Rosin as a natural material satisfies the major requirements for drug delivery applications. Its derivatives are normally eco-friendly, biodegradable and biocompatible [25,26]. The abietic acid which is the main component of rosin isomerizes thermally to levopimaric acid which performs Diels-Alder addition reaction with maleic anhydride to form an adduct [14].

Furthermore, the carboxylic acid group of abietic acid can form esters

readily by reacting with alcohols such as glycerol at higher tempera-ture [25]. Rosin and its derivatives have been pharmaceutically eval-uated as microencapsulating agents [27] and as anhydrous binders in tablets [28] or as controlled release hydrophobic matrix material in tablets [29]. The release kinetics of drugs from rosin-glycerol ester microcapsules were also investigated and correlated to the chemical constitution of the prepared esters [30]. Fulzele et al prepared a novel film forming polymer based on polymerized rosin (PR). The prepared polymer film was characterized and investigated for utilization in drug delivery [31]. Ramani and his coworkers studied both rosin/ma-leoabietic acid and diabietic acid as matrix forming materials [32,33].

Rosin-(2-acryloyloxy) ethyl ester was grafted onto chitosan via micro-wave irradiation using potassium persulfate as an initiator. The resul-tant copolymers were applied as carriers of fenoprofen calcium, and their controlled release behavior in artificial intestinal juice was stud-ied [34]. Recent studies also indicated that rosin nanoparticles can be effective for the encapsulation and delivery of drugs [35].

Rosin Derivatives for Pharmaceutical Coating

The successful employment of films as specific coatings on medica-tions, as vehicles for medicaments or as packaging agents has motivat-ed several studies which investigated various film forming materials for these applications [36,43]. The application of a polymer film coat is a widespread practice in the preparation of sustained and controlled release dosage forms [37]. The discharge of a drug from a coated par-

Page 9: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 9 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

ticle has been studied by various investigators for different conditions [32,33,38,39]. The value of film forming materials has often been char-acterized in terms of their mechanical properties [40], permeability [41,42] and water vapour transmission [43].

Rosin based polymer has been used as film coating materials; coated pellets were prepared using diclofenac sodium as a model drug and sustained release of the drug was achieved [44]. Rosin polymer has used as the trans- dermal drug delivery system. Blending of Rosin with polyvinyl pyrrolidone and dibutyl phthalate (30 % w/w) pro-duces smooth film with superior elongation and tensile strength [29]. Rosin and rosin esters such as pentaerythritol rosin ester gum were evaluated as drug coating materials [45,46]. Furthermore, Sahu et al [47] performed valuable in-vivo biodegradation studies of Rosin-glyc-erol ester derivative, whereas, Dheorey and Dorle [48,49] investigated the release kinetics of drugs from rosin-glycerol ester microcapsules prepared by solvent evaporation technique. Satturwar and his cowork-ers [50] synthesized and evaluated some rosin-based polymers as film coating materials. In the same area of interest, Lakshmana et al [51] prepared and evaluated some rosin-coated microspheres for sustained release of aceclofenac.

Rosin Derivatives in Industrial Field

Rosin and its derivatives also find wide applications in different in-dustries.

Rosin derivatives in paper industry

Specific additives are introduced into most papers to give the intrinsic hydrophylic character of cellulose fibers and also to shrink this ten-dency to different extents to fit specific use. For example, to lessen the penetration of aqueous liquids or the excessive wetting associated with

certain printing processes [52]. This partial hydrophobization can be achieved by bulk or surface treatment with adequate agents, which bind to the fibers’ surface, either chemically or by other less strong in-teractions. This treatment is known as paper sizing (Figure 14). Rosin and alum (aluminium sulphate), alkenyl succinic anhydrides (ASA) and alkyl ketene dimers (AKD) are the most significant reagents ap-plied for this purpose [52,55]. Sizing with rosin/alum was invented in 1807 [55,56] and had been the most important process up to the middle of the last century. However, the acidic conditions related to its completion were the main reason for its progressive abandonment due to two main drawbacks:

1. Acidic condition cause acid-catalyzed hydrolysis for cellulose. This may also result in progressive loss in the mechanical strength of the paper sheet.

2. Acidic papers are incompatible with calcium carbonate which is very common filler.

These problems forced a steady decrease in the utilization of ros-in alum in favor of the alternative use of AKD and ASA. For all the above-mentioned problems, research dealing with the implementa-tion of rosin sizing under neutral conditions has been a continuous in-terest of investigation [53]. Rosin can be used in paper sizing in three forms, namely as sodium salts (known as neutral rosin), free rosin acids and modified rosin products. One of the most important rosin derivatives in this context is the maleopimaric acid adduct (Figure 11), whose incorporation (about 8-15 %) into the sizing composition (a process known as fortification) provides a marked improvement in ef-ficiency compared with unmodified rosin [52]. Similar modifications can be reached with fumaric or itaconic acid adducts, among others, although the maleopimaric acid adduct dominates the industrial ap-plications of this family of agents.

Figure 14: Interaction of aluminium resinate with cellulose surface

Page 10: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 10 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

The addition of maleopimaric acid adduct induces two main changes: • The increased hydrophilicity associated with the tricarboxylic

functionality of the maleopimaric acid adduct

• The equivalent increased anionicity reduces the tendency of thesizing agent to agglomerate and promotes a better dispersion over the sheet surface, while also increasing the reactivity with alum to form the corresponding aluminium rosinate, which is the actual sizing agent.

The mechanism of paper sizing with rosin was discussed by several theories [52,56]. The most sensible theory is the one based on the po-larity of cellulose surface. As cellulose surface is negatively charged, it can interact with the sizing agent leading to the in situ formation of an aluminium resinate, where the aluminium ions establish an elec-trostatic interaction between the negatively charged carboxylic groups of the resinate and the sheet surface (Figure 14). Many additives have been introduced in order to enhance the cationicity of the aluminium species in neutral to basic conditions, so that rosin sizing can be ap-plied to papermaking without the problems related to acidity. Among them, cellulose polyamines [57,58] and cellulose polyimines [59,60], sometimes in conjunction with rosin esters [61], have attracted much attention because of their considerable cationic charge in neutral con-ditions.

Rosin Derivatives in Printing Ink Industry

Generally, the properties of rosin itself make it impossible as a compo-nent of any printing inks formulations. However, several of its deriva-tives have been, and are being, broadly used in the manufacture of all types of printing inks.

The most common rosin derivatives utilized in this purpose are rosin oligomers and their esters, metal resinates, modified phenol formalde-hyde (Figure12) and alkyd resins, ester gums, maleic and fumaric acid adducts and their esters [62,64]. Basically, rosin-based components can be used in the manufacture of all types of printing inks because they afford good miscibility and compatibility with most film formers and other ink additives. Of the numerous rosin-based esters used in printing inks, the simplest are those formed by the direct condensa-tion reaction of rosin or its oligomers with glycerol and pentaerythri-tol at high temperature [65].

Rosin Derivatives as Adhesive

Adhesive tack refers to the ability of instantaneous binding when an adhesive and a surface are brought into contact [66,67]. When com-pared to terpenes and hydrocarbon resins, it was found that Rosin esters have superior adhesive properties. Thus, when they are incor-porated into a variety of polymer substrates, they enhance their tack. These polymers include ethylene-vinyl acetate copolymers [68,69], natural rubber [70] and thermoplastic elastomers [71, 72], polyure-thanes, where the rosin derivatives participate chemically in the poly-mer growth [73], and as water emulsions [74].

Rosin Derivatives for Coating Applications

Unsaturated polyester resins have been identified as basic matrix materials in the field of reinforced plastics and coatings. However, they show low acid and alkali resistances and low adhesion with steel when cured. Incorporating long chain aliphatic compounds into the chemical structure of UP resins greatly improves resin flexibility. In this respect, Atta et al. [75] introduced specially designed UP resins and hardeners based on rosin derivatives to produce cured UP resins which have good durability with excellent mechanical properties.

Based on the sources and application requirements, varying weights of maleic anhydride or fumaric acid are reacted with rosin. A typi-cal commercial rosin adduct for coating applications is prepared by incorporating rosin with about 4-6 wt.% maleic anhydride. The hy-drophobic skeleton of Rosin has in combination with the hydrophilic carboxyl groups afforded rosin an excellent solubility and compatibil-ity with a variety of other synthetic resins. It is useful to modify these resin acids for many coatings applications. For example, the incorporation of rosin or its adduct into alkyds and varnishes modified their drying properties and resistance towards dif-ferent chemicals [76,77]. Furthermore, Resin acid dimer adducts with maleic anhydride and acrylic acids were also used to prepare epoxy resins [78]. The epoxy precursors were prepared by the reaction of the Diels-Alder adducts with di(ethanol)amine, followed by treatment with epichlorohydrine under alkaline conditions, as shown in Fig. 15 for the acrylic acid adduct [79]. After curing, the resulting materials create high stability coats.

Page 11: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 11 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

Figure 15: Synthesis of epoxy rosin acid dimer adduct with acrylic acid

Rosin Derivatives in Some Petroleum Applications

Resinate salts can act as surfactants due to their strong amphiphilic character. Thus, various types of surfactants can be prepared by insert-ing variant hydrophilic groups into rosin acids.

Chemical modification of dehydroabietic acid into surfactants is achieved by incorporation of different functional groups, e.g., carbox-ylic acid, amine, ester, alcohol, and an aromatic group [80]. The func-tional groups are useful as the linking unit to the hydrophilic part in synthesizing surfactants. Furthermore, a number of surfactants were synthesized from sugars and natural hydrophobic compounds such as rosin esters. Monosaccharide’s, including D-glucose, 2-deoxy -2-ami-no-D-glucose, D-fructose and D-(+)-glucono-1,5-lactone, were used as the hydrophilic moiety of the surfactants. The hydrophobic moiety consisted of steroids, monoterpenes, rosin acids, fatty acids and long chain alkyl groups, as well as aromatic compounds. In general, the synthetic procedures gave relatively high yields in simple steps. Some new surfactants were synthesized in high yields in one simple synthet-ic step [81].

Oil Spill Dispersants from Rosin Derivatives

Rosin derivatives were used as oil spill dispersants due to their non-toxicity. Atta and El-Saeed [82] prepared some nonionic rosin based polymeric surfactants via esterification of rosin with different molecular weights of polyethylene glycol (PEG 400, 600, 1000, 2000)

to create rosin ester surfactants. The esterified rosin was then reacted with maleic anhydride followed by reaction with diaminobutane or triethylene tetramine to produce rosin-imide. The surfactants were applied for dispersing sludge versus time.

In the same trend, Atta et al. prepared some water-based oil spill dis-persants based on rosin formaldehyde resins [83]. Furthermore, the surface and thermodynamic properties of the prepared nonionic sur-factants were studied [84]. Atta et al could also prepared nonionic Sur-factants from Rosin-imides Maleic Anhydride Adduct to be used as oil spill dispersants [85].

Moreover, Tall oil fatty acids were epoxized and the resultant sur-factants were investigated as oil spill dispersants [86]. Svensson [87] modified rosin acids into aclicyclic surfactants compounds and ap-plied those surfactants in some petroleum applications.

Corrosion Inhibitors from Rosin Derivatives

Corrosion of pipelines and production equipments is one of the most important problems encountered during petroleum production and exploration. Synthesis of corrosion inhibitors derived from or based on natural materials has attracted great attention [88,89]. Atta et al [90] introduced novel unsaturated polyesters derived from rosin ma-leic anhydride adduct as corrosion inhibitors for steel.

Moreover, Dehydroabietylamine (Fig.7) and the corresponding am-

Page 12: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 12 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

monium salts find a wide diversity of uses from cationic flotation agents to antioxidants, fungicides, and anticorrosion materials [5]. More recently, water soluble polyamides and poly (amide-imides)s as well as polyhydroxyimides [91] structurally equivalent to those de-scribed in Figure13 have been proposed as corrosion inhibitors for carbon steel.

The anti-corrosion efficiency of six water-soluble rosin-based inhibi-tor formulations on the corrosion rates of aluminum and steel in sea-water were investigated. Corrosion rates were measured by the weight loss of specimens following long-term immersion in artificial seawater containing inhibitors at the developer’s suggested dosages [92].

Emulsifiers from Rosin Derivatives

Rosin is acknowledged as green petroleum because it is renewable, not expensive, and environmental friendly. Rosin and its derivatives are frequently used as surfactants having different chemical structures. Rosin was used as raw material to prepare rosinyl amine salt (RAS) surfactant. The prepared surfactant had strong emulsification activity and high foam stability [93]. Some polymers from rosin acrylic-acid adduct were also used as emulsifiers for stabilizing re-constituted pe-troleum emulsions [94]. Warwel et al. [95] could prepare various types of surfactants based on natural products such as cellulose and rosin. The prepared compounds were evaluated as emulsifiers and also as demulsifies for petroleum crude oil. Recently, some Rosin-Based non-ionic surfactants were synthesizes and applied as emulsifiers for high internal phase emulsions [96].

Future Trends

Extensive consideration is being given to the manufacture of biode-gradable synthetic polymers [97,98]. Materials derived from biode-gradable natural polymers obtained from constantly renewed vegeta-ble raw material are of ever-increasing interest, too. A special place among them is occupied by rosin-based plastics. Rosin is a typical bio-degradable, cheap and available. The design and synthesis of nano- to micro-sized polymer composite with core–shell nanostructures have attracted more and more attention because of their wide applicability in modern material science and their technological importance in the areas of colloid and interface science. Many synthetic strategies have been developed for the preparation of ferro core-shell particles that consist of nano ferro cores and hydrophilic or hydrophobic polymeric shells.

Rosin Nanocomposites in Industrial water Treatment

Saving water is very necessary to save the planet and to make the fu-ture of mankind safe. With the growth of mankind, society, science,

technology our world is reaching to new high horizons but the cost which we are paying or will pay in near future is surely going to be too high. Among the consequences of this rapid growth is environmental disorder with a big pollution problem.

Life on earth is threatened by two major dangers. Namely, heavy metal pollution -resulted from rapid industrialization- and the increase in the world population. Metal ions are persistent to biodegradation and have accumulative effect inside the living tissues. The metals that have high environmental impacts are Cr, Hg, Cu, Ni, Zn and Cd, because they are commonly associated with pollution and toxicity problems [99]. Many processes have been proposed to restrict heavy metal pol-lution, including chemical precipitation, electrode deposition, solvent extraction, ion-exchange, activated carbon adsorption and biological methods [100,101]. Among these methods, adsorption has increas-ingly received more attention in recent years because it is simple, relatively low-cost, and effective in removing heavy metal ions from wastewaters [102,103]. Nanotechnology has been considered as one of the most important advancements in science and technology. Its es-sence is the ability to fabricate and engineer the materials and systems with the desired structures and functionalities using the nano-sized building blocks [104]. Nanoparticles are one of the important build-ing blocks in fabrication of nanomaterials. Their basic properties, extremely small size and high surface-area-to volume ratio, provide better kinetics for the adsorption of metal ions from aqueous solu-tions. However, for such an application, it is necessary to use a method of purification that does not generate secondary waste and involves materials that can be recycled and easily used on an industrial scale. Magnetic separation has been shown to be a very promising method for solid–liquid phase separation technique. To facilitate the recovery and manipulation of nanoparticles, magnetism is incorporated with the nanoparticles. An extensive literature on iron-oxide and ferrite nanostructures and their composites [104,105] attests to the vast technological importance of these materials for broad application in the nanotechnologies of information storage, bioprocess and ferro-fluids [106,107]. The particles synthesized and dispersed in aqueous medium can be trapped in different polymeric solid matrices keep-ing a high dispersion state. The oxide/polymer mass ratio determines the average distance between the nuclei existing in the sol, isolated particles or aggregates. This makes magnetic nanoparticles excellent candidates for combining metal binding and selective adsorption properties with ease of phase separation [107]. Numerous types of magnetic nanoparticles for various applications could be tailored by using functionalized natural or synthetic polymers to impart surface reactivity. For instance, a novel magnetic nano-adsorbent was devel-oped for the adsorption of metal ions by the surface modification of Fe3O4 nanoparticles with gum Arabic which is a natural, harmless and environment friendly polymer containing active functional groups like carboxylate and amine groups [107]. Gum arabic was attached to

Page 13: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 13 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

Fe3O4 via the interaction between the carboxylic groups of gum arabic and the surface hydroxyl groups of Fe3O4.

Conclusions

Rosin is a versatile natural product which has been used long time ago. The rapid industrialization pushed scientist’s efforts towards replacing petroleum based products with natural materials in order to reduce the environmental impacts and to save natural resources. In this re-spect, rosin and its derivatives have been utilized in a versatile num-ber of applications, from medicine to industry. Rosin is an interesting candidate for chemical modification through a series of chemical re-actions that can be performed on its functional groups.

References

1. Sjostrom E (1981) Wood Chemistry: Fundamentals and Applications, Academic Press New York.

2. Enos HI, Harris GC, Hedrick GW, in Mark HF, McKetta JJ Jr et al. (1968) Encyclopedia of Chemical Technology, 2nd edn., Wiley, New York 17: 475.

3. Weaver JC, in Mark HF, Othmer DF, Overberger CG, Seaborg GT et al. (1982) Encyclopedia of Chemical Technology, 3rd edn. Wiley, New York 20: 19.

4. Zachary LG, Bajak HW, Eveline FJ (1965) Tall Oil and Its Uses, Mc-Graw-Hill New York.

5. DE Zinkel, J Russel (1989) Naval Stores. Production, Chemistry and Utilization, Pulp Chemical Association, New York.

6. Maiti S, Ray SS, Kundu AK (1989) Rosin-A renewable resource for polymers and polymer chemicals. Prog Polym Sci 14(3): 297-338.

7. Stauffer DF (1989) Production, Markets and Economics, In Naval Stores. Production, Chemistry and Utilization (Eds. Zinkel D.F.and J. Russel Pulp Chemical Association, New York 39-82.

8. Hone GA, Coppen JJW (1995) Gum Naval Stores: Turpentine and Rosin from Pine Resin, FAO Rome

9. Brower M (1990) Cool Energy: The Renewable Solution to Global Warming, Union of Concerned Scientists, Cambridge MA.

10. Tsujimoto T, Ueda H, Kageyama H, Uyama H (2016) Bio-based Poly-meric Materials from Epoxidized Triglyceride and Rosin Derivatives, Trends in Green Chemistry 2(1): 1.

11. Kanlaya P, Sumrit W, Amorn P (2016) Synthesis and Characterization of Water Soluble Rosin-Polyethylene Glycol 1500 Derivative, Int. J. of Chem. Eng and Appl Vol 7( 4).

12. Pathak YV, Dorle AK (1987) Study of rosin and rosin derivatives as coating materials for controlled release of drug, J. of controlled release 5(1): 63-68.

13. Bhupendra Kumar Yadav, Bina Gidwani, Amber Vyas (2015) Rosin: Recent advances and potential applications in novel drug delivery sys-tem, Journal of Bioactive and Compatible Polymers 31(2): 111-126.

14. Portugal I, Vital J, Lobo L (1992) Resin acids isomerization - A kinetic study Chem. Eng. Sci 47(9-11): 2671-2676.

15. Stirna U, Sevastyanova I, Misanea M, Cabulisa U, Beverte I et al. (2006) Structure and properties of polyurethane foams obtained from rapeseed oil polyols, Proc. Estonian Acad. Sci Chem 55(2): 101-110.

16. Kotsuki H, Kataoka M, Matsuo K, Suetomo S, Shiro M et al. (1993) High-pressure organic-chemistry. 17. Diels–Alder reaction of methyl palustrate with maleic-anhydride and N -phenylmaleimide, J. Chem. Soc. Perkin Trans. I(22): 2773-2776.

17. Maiti S, Ray S, Kundu AK (1989) Rosin – A renewable resource for polymers and polymer chemicals, Prog. Polym. Sci 14(3): 297-338.

18. Kundu AK, Ray SS, Maiti S (1986) Polymer blends. 4. Blends of resol with polyamideimide synthesized from rosin, Eur. Polym J 22(10): 821-825.

19. Bicu I, Mustafa F (2000) Water soluble polymers from Diels–Al-der adducts of abietic acid as paper additives, Macromol. Mater Eng 280(7-8): 47-53.

20. Kim SJ, Kim BJ, Jang DW, Kim SH, Park SY et al. (2001) Photoac-tive polyamideimides synthesized by the polycondensation of azo-dye diamines and rosin derivative , J. Appl. Polym Sci 79(4): 687- 695.

21. Bingham JF, Marvel CS (1972) Preparation and polymerization of a diisocyanate from Diels–Alder adduct of levopimaric acid and acrylic acid J Polym Sci A1 Pol Chem 10(3): 921–929.

22. Atta AM, Sh El-Saeed, Farag R (2006) New vinyl ester resins based on rosin for coating applications, React. Funct. Polym 66(12): 1596 – 1608.

23. Bicu I, Mustafa F (1997) Allylic polymers from resin acids, Angew Makromol Chem 246(1): 11-22.

24. Barabde UV, Fulzele SV, Satturwar PM, Dorle AK, Joshi SB et al. (2005) Film coating and biodegradation studies of new rosin deriva-tive, React. Funct Polym 62(3): 241-248.

25. Prashant M, Fulzele V, Avinash K (2003) Biodegradation and in vivo biocompatibility of rosin: a natural film-forming polymer, AAPS Pharm Sci Tech 4(4): 434-439.

Page 14: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 14 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

26. Mandaogade PM, Satturwar PM, SV Fulzele, BB Gogte, Dorle AK et al. (2002) Rosin derivatives: novel film forming materials for con-trolled drug delivery, React. Funct. Polym 50(3): 233–242.

27. Satturwar PM, Fulzele SV, Panyam J, Mandaogade PM, Mundhada DR et al. (2004) Evaluation of new rosin derivatives for pharmaceuti-cal coating, International Journal of Pharmaceutics 270(1-2): 27–36.

28. Pathak YV, Dorle AK (1990) Rosin and rosin derivatives as hydro-phobic matrix materials for controlled release of drugs. Drug Design Deliv 6(3): 223–227.

29. Sheorey DS, Dorle AK (1991) Release kinetics of drugs from ros-in–glycerol ester microcapsules prepared by solvent evaporation tech-nique. J. Microencapsulation 8(2): 243–246.

30. Fulzele SV, Satturwar PM, Dorle AK (2002) Polymerized rosin: novel film forming polymer for drug delivery, International Journal of Pharmaceutics 249(1-2): 175-184.

31. Ramani CC, Puranik PK, Dorle AK, (1996a) Study of rosin/maleoa-bietic acid as a matrix forming material. International Journal of Phar-maceutics. Adv. 1, 344/352.

32. Ramani CC, Puranik PK, Dorle AK, (1996b) Study of diabietic acid as matrix forming material. International Journal of Pharmaceutics 137(1): 11-19.

33. Wengui D, Chunhong C, Linbin J, Guang L, (2008) Prepara-tion and characterization of the graft copolymer of chitosan with poly[rosin-(2-acryloyloxy)ethyl ester], Carbohydrate Polymers 73(4): 582–586.

34. Lee CM, Lim S, Kim GY, Kim D, Lee KY et al. (2004) Rosin nanopar-ticles as drug carriers: Influence of various solvents on the formation of particles and sustained-release of indomethacin, Biotechnol. Bio-process Eng 9(6): 476-481.

35. Nagarajan V, Mohanty AK, Misra M (2013) Sustainable green com-posites: Value addition to agricultural residues and perennial grasses. ACS Sustainable Chem Eng 1(3): 325-333.

36. Wang R, Schuman TP (2013) Vegetable oil-derived epoxy monomers and polymer blends: A comparative study with review. EXPRESS Polym Lett 7(3): 272-292.

37. Phuapradit W, Railkar NH, Williams W, Infeld L (1995) In vitro char-acterization of polymeric membrane used for controlled release appli-cation. Drug Dev Ind Pharm 21(8): 955-963.

38. Quirino RL, Garrison TF, Kessler MR (2014) Matrices from vege-table oils, cashew nut shell liquid, and other relevant systems for bio-composite applications Green Chem 16(4): 1700-1715.

39. Kokubo H, Obara S, Nishiyama Y (1998) Application of extremely low viscosity methyl cellulose for pellet film coating. Chem. Pharm Bull 46(11): 1803-1806.

40. Chang-Moon Lee, Seung Lim, Gwang-Yun Kim, Doman Kim, Dong-Woon Kim et al. (2004) Rosin nanoparticles as drug carriers: Influence of various solvents on the formation of particles and sustained-release of indomethacin, Biotechnol. Bioprocess Eng 9(6): 476-481.

41. Park JH, Ye M, Park K (2005) Biodegradable Polymers for Microen-capsulation of Drugs Molecules 10(1): 146-161.

42. Prashant MS, Suniket VF, Avinash KD (2005) Evaluation of po-lymerized rosin for the formulation and development of transdermal drug delivery system: A Technical Note, AAPS PharmSciTech 6(4): E649-E654.

43. Zhang Y, Chu CC (2002) In vitro release behavior of insulin from bio-degradable hybrid hydrogel networks of polysaccharide and synthetic biodegradable polyester, J Biomater Appl 16(4): 305-325.

44. Mandaogade PM, Satturwar PM, Fulzele SV, Gogte BB, Dorle AK et al. (2002) Rosin derivatives: novel film forming materials for con-trolled drug delivery, React Funct Polym 50(3): 233-242.

45. Pathak YV, Nikor RL, Dorle AK (1985) Study of rosin and rosin esters as coating material Int J Pharm 24(2-3): 351-354.

46. Pathak YV, Dorle AK (1986) Evaluation of pentaerythritol rosin ester gum as coating material. Drug Dev Ind Pharm 12(11-13): 2217-2229.

47. Sahu NH, Mandaogade PM, Deshmukh AM, Meghre VS, Dorle AK et al. (1999) Bio-degradation studies of Rosin–glycerol ester deriva-tive. J. Bioact Comp Polym 14: 344-360.

48. Dheorey DS, Dorle AK (1990) Preparation and in vitro evaluation of rosin microcapsules: solvent evaporation technique. J Microencapsu-lation 7(2): 261-264.

49. Sheorey DS, Dorle AK (1991) Release kinetics of drugs from ros-in–glycerol ester microcapsules prepared by solvent evaporation tech-nique. J Microencapsulation 8(2): 243–246.

50. Mandaogade PM, Satturwar PM, Fulzele SV, Gogte BB, Dorle AK et al. (2002) Synthesis and evaluation of rosin-based polymers as film coating materials, Drug Dev Ind Pharm 28(4): 381-387.

51. Lakshmana PS, Shirwaikar AA, Shirwaikar A, Kumar A (2009) For-mulation and evaluation of sustained release microspheres of rosin containing aceclofenac, J. Microencapsulation 50(2): 51-62.

52. Niemo L (1999) Internal sizing of paper, in Papermaking Chemistry (Ed. Niemo L.) Tappi Press Helsinki.

Page 15: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 15 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

53. Roberts JC (1996 ) Neutral and alkaline sizing , in Paper Chemistry (Ed. Roberts J.C. ) Blackie Academic & Professional London pp 140-160.

54. Gess JM (1996) The sizing of paper with rosin and alum at acid pHs, in Paper Chemistry (Ed. Roberts J.C. ) , Blackie Academic and Profes-sional Glasgow pp 120-139.

55. Strazdins E (1989) Paper sizes and sizing, in Naval Stores. Pro-duction, Chemistry, Utilization (Eds. Zinkel D.F. and Russel J Pulp Chemical Association New York pp 575-624.

56. Strazdins E (1989) Paper sizes and sizing, in Naval Stores. Produc-tion, Chemistry, Utilization (Eds. Zinkel D.F. and Russel J.), Pulp Chemical Association New York pp 666-714.

57. Wu ZH, Chen SP (2001) Functions of polyamines in rosin sizing under neutral papermaking conditions, Chin. J. Polym. Sci 19(2): 217-221.

58. Hartong B, Deng Y (2004) Evidence of ester bond contribution to neutral to alkaline rosin sizing using polyethyleneimine-epichlorohy-drin as a mordant, J Pulp Paper Sci 30(7): 203-209.

59. Xu Y, Hartong B, Deng Y (2002) Neutral to alkaline rosin sizing using polyethyleneimine-epichlorohydrin (PEI-epi) as a mordant, J. Pulp Pap Sci 28(2): 39-44.

60. Li H, Ni Y, Sain M (2003) Further understanding on polyethylenei-mine-induced rosin ester sizing, Nord. Pulp Pap Res J 18(1): 5-9.

61. Wang F, Kitaoka T, Tanaka H (2003) Supramolecular structure and sizing performance of rosin-based emulsion size microparticles, Coll. Surf. A Physicochem. Eng Aspects 221(1-3): 19-28.

62. LeBlanc PJ, Peter S (1993) Rosin And Fatty Acid Based Pigment Grinding Aids For Water-based Ink formulations, US patent 5: 182-326.

63. Leach RH, Pierce RS (1993) The printing ink manual, Dordrecht, Kluwer Academic Publishers.

64. Pekarovicova A, Bhide H, Fleming PD, Pekarovic J (2003) Phase-change inks , J. Coating Technol 75( 936 ): 65-72.

65. Comyn J (1995) Surface characterization of pentaerythritol rosin ester, Int. J. Adhesion and Adhesives 15(1): 9- 14.

66. Kennedy NL, Krajca KE, Russel J (1989) Rosin in adhesive tacki-fiers, in Naval Stores. Production, Chemistry, Utilization (Zinkel D.F. and Russel J.), Pulp Chemical Association New York 645-664.

67. Fattah H (1996) Adhesives and sealants-Rosin resin makers fight back, Chem. Week 158(45): 45.

68. Takemoto M, Karasawa T, Mizumachi H, Kajiyama M (2000) Misci-bility between ethylene vinyl acetate copolymers and tackifier resins J Adhesion 72(1): 85-96.

69. Barrueso-Martinez ML, Ferrandiz-Gomez TD, Romero-Sanchez MD, Martin-Martinez JM (2003) Characterization of EVA-based adhesives containing different amounts of rosin ester or polyterpene tackifier, J. Adhesion 79(8-9): 805-824.

70. Fujita M, Kajiyama M, Takemura A, Mizumachi H, Hayashi S et al. (1998) Effects of miscibility on probe tack of natural rubber- based pressure-sensitive adhesives, J. Appl. Polym Sci 70(4): 771-776.

71. Kim DJ, Kim HJ, Yoon GH (2004) Effects of blending and coating methods on the performance of SIS (styrene-isoprene-styrene)-based pressure-sensitive adhesives, J. Adhes. Sci Technol 18(15-16): 1783-1797.

72. Yang H, Sa U, Kang M, Ryu CY, Cho K et al. (2006 ) Near-surface morphology effect on tack behavior of poly(styrene-butadiene-sty-rene) triblock copolymer/rosin films Polymer 47(11): 3889-3895.

73. Aran-Ais F, Torro-Pala AM, Orgiles-Barcelo AC, Martin-Martinez JM (2005) Addition of rosin acid during thermoplastic polyurethane synthesis to improve its immediate adhesion to PVC, Int. J Adhes 25(1): 31-38.

74. Kim BJ, Kim SE, Do HS, Kim S, Kim HJ (2007) Probe tack of tack-ified acrylic emulsion PSAs, Int. J Adhes 27(2): 102-107.

75. van de Pas DJ, Torr KM (2017) Biobased Epoxy Resins from Decon-structed Native Softwood Lignin, Biomacromolecules 18(8): 2640-2648.

76. Hedrick GW, Lawrence RV, Sum HB (1985) Levopimuric Acid, USDA Report ARS pp 35-72.

77. Aigbodion AI, Okiemen FE, Iukhuoria EU, Bakare IO, Obazee EO (2003) Rubber seed oil modified with maleic anhydride and fumaric acid and their alkyd resins as binders in water-reducible coatings. J Appl Polym Sci 89(12): 3256-3259.

78. Atta AM, Mansour R, Abdou MI, Sayed AM (2004) Epoxy resins from rosin acids: Synthesis and characterization, Polym. Advan Tech-nol 15(9): 514-522.

79. Atta AM, Mansour R, Abdou MI, Sayed AM (2005) Synthesis and characterization of tetra-functional epoxy resins from rosin, J Polym Res 12(2): 127-138.

80. Peter S, Björn H, Torbjörn N (2002) Synthesis and characterization of dehydroabietic acid derivatives suitable for surfactant synthesis, Journal of Surfactants and Detergents 5(2): 165-168.

Page 16: Rosin: Chemistry, Derivatives, and Applications: a review · of terpene and terpenoid derivatives [1,2]. Moreover, the extractives Moreover, the extractives content changes after

BAOJ Chem, an open access journal Volume 4; Issue 1; 039

Page 16 of 16Citation: Manar El-Sayed Abdel-Raouf, Abdul-Raheim Mahmoud Abdul-Raheim (2018) Rosin: Chemistry, Derivatives, and Applications: a

review. BAOJ Chem 4:039

81. Piispanen S (2002) Synthesis and Characterization of Surfactants Based on Natural Products. M.Sc. thesis, Kungl Tekniska Högskolan, Stockholm.

82. Atta AM, El-Saeed A (2011) Use of rosin-based nonionic surfactants as petroleum crude oil sludge dispersants, Journal of Applied Polymer Science 122(1): 183-192.

83. Atta AM, Abdul-Raouf ME, Nermine E Maysour, Amani k Gafer (2010) Water Based Oil Spill Dispersants Based on Rosin Formalde-hyde Resins, Journal of Dispersion Science and Technology 31(5): 583-595.

84. Atta AM, El-Kafrawy AF, Abdul-Raouf ME, Maysour NE, Gafer AK (2010) Surface and Thermodynamic Properties of Nonionic Surfac-tants Based on Rosin- Maleic Anhydride and Acrylic Acid Adducts, Journal of Dispersion Science and Technology 31(4): 567-576.

85. Atta AM, Ramadan AM, El-Shafay KA, Mohamed AM, Ragab NS et al. (2009) Synthesis and Properties of Nonionic Surfactants from Rosin-imides Maleic Anhydride Adduct, Journal of dispersion science and technology 30(7): 1100-1110.

86. Bjorn H, Piispanena P, El-Ouafi A, Norin T (2003) Synthesis and Characterization of Surfactants via Epoxidation of Tall Oil Fatty Acid, Journal of Surfactants and Detergents 6(1): 47-53.

87. Svensson M (1998) Sterol Surfactants and Other Alicyclic Com-pounds, in Novel Surfactants: Preparation, Applications and Biode-gradability, edited by K. Holmberg, Marcel Dekker, New York 179.

88. Chirkunov AA, Yu I Kuznetsov , Gusakova MA (2007) Protection of Low-Carbon Steel in Aqueous Solutions by Lignosulfonate Inhibitors, Protection of Metals 43(4): 367–372.

89. Quraishi MA, Jamal D, Saeed MT (2000) Fatty Acid Derivatives as Corrosion Inhibitors for Mild Steel and Oil-Well Tubular Steel in 15% Boiling Hydrochloric Acid JAOCS 77(3): 265-268.

90. Atta AM, Nassar IF, Bedawy HM (2007) Unsaturated polyester res-ins based on rosin maleic anhydride adduct as corrosion protections of steel, Reactive & Functional Polymers 67(7): 617-626.

91. Mustata F, Bicu I (2000) Polyhydroxyimides from resinic acids Po-limery 45(4): 258-263.

92. Lambrakos SG, Trzaskoma-Paulette PP, Cooper KP, Tran NE (2004) Properties and Effects of Water-Soluble Inhibitors on the Corrosion Rates of Structural Metals JMEPEG 13: 766-774.

93. Yan J, Shuang Y, Shu J, Zhang LJ (2011) Synthesis and Characteri-zation of Rosinyl amine salt surfactant, Journal of advanved material research 183-185:1888-1891.

94. Roy SS, Kundu AK, Maiti S (1990) Polymers from renewable re-sources. 13. Polymers from rosin acrylic-acid adduct, Eur Polym J 26(4): 471-474.

95. Warwel S, Brüse F, Demes C, Kunz M, Klaas MR (2001) Polymers and Surfactants on the Basis of Renewable Resources, Chemosphere 43(1): 39-48.

96. Lim H, Kassim A, Huang N, Yarmo MA (2009) Rosin-Based Non-ionic Surfactants as Emulsifiers for High Internal Phase Emulsions, J Surfact Deterg 12(4): 355-362.

97. Fomin VA, Guzeev VV (2001) “Biodegradable polymers, state of the art and prospectives of application” Plast. Massy (2): 42-48.

98. Loshadkin DV (2002) “Biodegradable plastics: types of materials, their basic properties, and prospective industrial applications” Plast Massy (7): 41-44.

99. Mido Y, Satake M (1995) Chemicals in the Environment, Discovery Publishing House New Delhi.

100. Banerjee SS, Jayaram RV, Joshi MV (2005) Removal of Cr(VI) and Hg(II) from aqueous solutions using fly ash and impregnated fly ash, Sep Sci Technol 39(7): 1611-1629.

101. Murlidhara HS (1986) Advances in Solid–Liquid Separation, Batelle Press, Colombus Richland Ohio.

102. Siegel RW, Hu E, Roco MC (1999) Nanostructure Science and Tech-nology, A Worldwide Study, WTEC, Loyola College Kluwer Academ-ic Baltimore MD.

103. Ngomsik A, Bee A, Draye M, Cote G, Cabuil V (2005) Magnetic nano and microparticles for metal removal and environmental applica-tions: a review CR Chimie 8(6-7): 963-970.

104. Amara D, Felner I, Nowik I, Margel S (2009) Synthesis and charac-terization of Fe and Fe3O4 nanoparticles by thermal decomposition of triiron dodecacarbonyl. Colloids Surf. A: Physicochem. Eng Aspects 339(1): 106-110.

105. Aslam K (2008) Preparation and characterization of magnetic nanoparticles embedded in microgels. Mater Lett 62(6-7): 898-902.

106. Hong RY, SZ, Li GQ, Zheng HZ (2008) Y and J. Ding; Preparation, characterization and application of Fe3O4/ZnO core/shell magnetic nanoparticles. Mater. Res Bull 43: 2457-2468.

107. Shashwat S, Banerjee S, Dong-Hwang C (2007) Fast removal of copper ions by gum Arabic modified magnetic nano-adsorbent. Jour-nal of Hazardous Materials 147(3): 792-799.