Comparative Studies on the Crosslinking Reaction of Bis-Maleimides and Bis-Citraconimides in Squalene and Natural Rubber

1998 ◽  
Vol 71 (5) ◽  
pp. 1073-1086 ◽  
Author(s):  
R. N. Datta ◽  
A. G. Talma ◽  
A. H. M. Schotman

Abstract A study has been made to differentiate between the crosslinking reaction of bis-maleimides and bis-citraconimides in squalene as well as in natural rubber. Bis-maleimides are found to participate in the crosslinking reaction in the absence of curatives such as accelerators and sulfur. Bis-citraconimides take part in the crosslinking reaction when polysulfidic crosslinks are under the process of degradation. Sulfur and accelerator are required for the formation of polysulfidic crosslinks and hence the presence of sulfur and accelerator are essential for bis-citraconimide crosslinking. The differences in the reactivity and the chemistry of crosslinking are studied and discussed. Some application data are provided in order to elaborate the differences.

2014 ◽  
Vol 1025-1026 ◽  
pp. 580-584 ◽  
Author(s):  
Phrutsadee Sukpuang ◽  
Mantana Opaprakasit ◽  
Atitsa Petchsuk ◽  
Pakorn Opaprakasit

Glycolyzed polylactic acid (GPLA)-cured epoxidized natural rubber (ENR) is developed for use as a toughening agent for PLA resin. GPLA is obtained from chemical recycling of PLA resin by a glycolysis reaction. GPLA-cured ENR is then prepared by the crosslinking reaction of ENR with GPLA in an internal mixer. Chemical structures of the cured products are characterized by solvent fractionation and thermogravimetric analysis (TGA). The cured ENR products are blended with PLA resin, by varying the cured ENR contents from 5 to 15% wt. Mechanical properties of the blends, and their toughening mechanisms are examined. The cured ENR materials has higher efficiency in improving toughness of PLA resin, compared to uncured ENR, likely due to their rubbery network nature and higher compatibility with the PLA matrix. The incorporation of 5% wt. GPLA-cured ENR also improves elongation at break with no adverse effect on tensile strength and modulus of PLA.


Polymers ◽  
2021 ◽  
Vol 13 (23) ◽  
pp. 4145
Author(s):  
Kriengsak Damampai ◽  
Skulrat Pichaiyut ◽  
Subhradeep Mandal ◽  
Sven Wießner ◽  
Amit Das ◽  
...  

In this work, studies are carried out to understand the crosslinking reaction of epoxidized natural rubber (50 mol% epoxy, ENR-50) by metal ion namely ferric ion (Fe3+, FeCl3, ferric chloride). It is found that a small amount of FeCl3 can cure ENR to a considerable extent. A direct interaction of the ferric ion with the epoxy group as well as internal polymerization enable the ENR to be cured in an efficient manner. It was also found that with the increased concentration of FeCl3, the crosslinking density of the matrix increased and therefore, the ENR offers higher mechanical properties (i.e., modulus and tensile strength). In addition, the glass transition temperature (tg) of ENR vulcanizate is increased with increasing concentration of FeCl3. Moreover, the thermal degradation temperature (Td) of the ENR-FeCl3 compound was shifted toward higher temperature as increasing concentration FeCl3.


1963 ◽  
Vol 36 (1) ◽  
pp. 219-235
Author(s):  
J. J. Shipman ◽  
M. A. Golub

Abstract Infrared absorption spectroscopy has been employed in the past in efforts to determine the nature of the chemical changes which occur in natural rubber during vulcanization. The most prominent change in the infrared spectrum of rubber accompanying vulcanization with sulfur is the appearance of a new absorption band at 10.4 µ. In the spectra of useful soft rubber vulcanizates this band is barely detectable, while in those of hard rubber (ebonite) it is moderately intense. It has not been known whether this band is related in some way to the actual crosslinking reaction or is the result of some noncrosslinking reaction between the rubber and sulfur molecules.


1992 ◽  
Vol 26 (3) ◽  
pp. 339-356 ◽  
Author(s):  
Yoshiaki Ikarashi ◽  
Kazuhiro Toyoda ◽  
Naomi Ohsawa ◽  
Takao Uchima ◽  
Toshie Tsuchiya ◽  
...  

2014 ◽  
Vol 695 ◽  
pp. 174-178 ◽  
Author(s):  
Norfhairna Baharulrazi ◽  
Hussin Mohd Nor ◽  
Wan Khairuddin Wan Ali

Controlled degradation of deproteinized natural rubber by oxidative method using cobaltous leading to the formation of telechelic liquid natural rubber (TLNR) has been the subject of study in the current work. Hydroxyl-terminated natural rubber (HTNR) having number-average molecular weights (Mn) below 10,000 gmol-1 were synthesized at temperatures 60°C using Cobalt bis (acetyl acetonate) (Co) in the presence of ethanol and Sodium borohydride. The HTNR were characterized structurally by using Fourier Transform Infrared (FTIR), Gel Permeation Chromatography (GPC) and Nuclear Magnetic Resonance (NMR) spectroscopy. From the FTIR and GPC analysis, one hour reaction showed the lowest Mn around 6691 gmol-1 comparable to Hydroxyl terminated polybutadiene (HTPB) which were 7708 gmol-1. As the reaction time increases the Mn also increase up to 8 hours of reaction. FTIR indicated that the synthesized HTNR contained hydroxyl end groups. The study showed the optimum percentage of Cobalt bis (acetyl acetonate) is 5% which produce the lowest molecular weight. Next, the HTNR underwent crosslinking reaction with isophore diisocyanate (IPDI). Then, metal fuel (aluminium powder, Al), oxidizer (ammonium perchlorate, AP) and HTNR were mixed together and went through compression process to produce solid rocket propellant. Burning rate obtained from HTNR samples were 2.78 mms-1 which were equivalent to HTPB samples which is 2.94 mms-1.


1963 ◽  
Vol 36 (3) ◽  
pp. 835-843 ◽  
Author(s):  
R. Russell ◽  
D. A. Smith ◽  
G. N. Welding

Abstract Thiazole-accelerated sulfur vulcanization of natural rubber gum stocks has been investigated at temperatures from 110° to 140° C for times over the range 1 to 104 min. Swelling measurements on the vulcanizates were used to estimate the contribution of chemical reactions to the network chain density, (ρMc−1)chem which was plotted against cure time. The kinetic graphs show clearly an induction period (t0), then fast crosslinking, a slow limited degradation, and finally a “long-time” crosslinking reaction. Apart from the induction period, each of the kinetic graphs is satisfactorily represented by a rate equation assuming three independent additive reactions: first order crosslinking, first order partial degradation, and a long-time reaction assumed to be zero order. The method has been used to compare MBT and MBTS at equal concentration. One consistent difference is the increase of induction period with MBTS in place of MBT, accompanied by a somewhat greater increase of time of cure to maximum network density. Also the completed contributions (Xand Δ) from the initial crosslinking reaction and the degradation reaction, respectively, are greater with MBTS than with MBT. Over the temperature range studied there is no difference between MBT and MBTS in the activation energies for crosslinking, or for t0−1 or for the long-time reaction, but MBTS has a higher activation energy for degradation than MBT. With rising temperature of cure, Δ increases towards X with each accelerator.


1958 ◽  
Vol 31 (1) ◽  
pp. 117-131 ◽  
Author(s):  
O. Lorenz ◽  
Elisabeth Echte

Abstract The vulcanization of natural rubber with sulfur, using mercaptobenzothiazole as accelerator in the presence of zinc stearate, was investigated. The results were as follows: 1) When natural rubber is cured with sulfur in the presence of zinc oxide and mercaptobenzothiazole, as well as zinc stearate, one observes, with the first order diminution of sulfur concentration, an induction period that grows longer as the cure temperature falls. 2) For the disappearance of sulfur there is calculated, from the temperature dependence of the rate constants, an activation energy of 19.5 kcal/mole. This value is considerably smaller than that found for the mercaptobenzothiazole-accelerated sulfur cure when no zinc stearate is present. 3) The percentage loss in mercaptobenzothiazole during vulcanization is, in the presence of zinc stearate, independent of the temperature; there is a reaction, independent of the amount of stearate added of 5 molecules of S8 for each molecule of mercaptobenzothiazole. 4) The diminution in accelerator can also be explained as a first order reaction, and it becomes evident that disappearance of accelerator and sulfur are equivalent-rate processes. 5) The crosslinking, measured by the reciprocal equilibrium swelling, goes according to the first order, yet sulfur disappearance and the crosslinking reaction are not equal rate processes, which was always true in the absence of zinc stearate. In each case the crosslinking rate constants are greater than those for the decrease in sulfur. 6) For the activation energy of network formation, we calculated, from the temperature dependence of the rate constants of the reciprocal equilibrium swelling, 20.5 kcal/mole. 7) When the stearate content is constant, there is a linear relationship between kSI, the rate constant for sulfur diminution, and the given amount of mercaptobenzothiazole, up to an initial molar ratio of mercaptobenzothiazole: stearate = 4:1. Increases in accelerator proportion beyond this initial ratio cause no further rise in the rate constants. 8) The rate constants of the crosslinking reaction also increase with increasing proportion of mercaptobenzothiazole (stearate portion remaining constant), until the initial molar ratio of mercaptobenzothiazole:zinc stearate = 4:1 is reached. Nevertheless, the relationship is not linear. 9) There is a linear relationship between the amount of zinc stearate in the vulcanizate and the maximum sulfur-loss rate constants kSI (max). 10) The maximum rate constants of the network forming reaction kQI (max) are proportional to the square root of the stearate content. 11) The number of crosslinkage points formed is independent of the quantity of mercaptobenzothiazole, when the stearate content is constant. 12) The number of crosslinking points formed increases, however, along with increasing stearate content, and is proportional to the square root of the stearate content. The investigations will be continued.


1993 ◽  
Vol 66 (2) ◽  
pp. 230-241 ◽  
Author(s):  
A. Roychoudhury ◽  
P. P. De ◽  
N. K. Dutta ◽  
N. Roychoudhury ◽  
B. Haidar ◽  
...  

Abstract A blend of chlorosulfonated polyethylene and epoxidized natural rubber undergoes a self-crosslinking reaction at elevated temperatures. Sulfonyl chloride groups of chlorosulfonated polyethylene react with epoxy groups of epoxidized natural rubber to give ether crosslinks between the two polymers. In addition to the self-crosslinking reaction, these two polymers undergo a number of side reactions as evidenced by spectroscopic studies.


Sign in / Sign up

Export Citation Format

Share Document