首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A copolymer prepared by the copolymerization of ethyl vinyl ether and maleic anhydride underwent amidoacidation with N,N‐dimethylaminopropylamine. The obtained ethyl vinyl ether/dimethylaminopropyl maleamidic acid copolymer was then reacted with methyl iodide to yield poly(methyl iodide quaternized ethyl vinyl ether/N,N′‐dimethylaminopropyl maleamidic acid) (MIQEDMAPMA). The greatest difference from other polyelectrolytes was the carboxylic group on the polymer chain unit of MIQEDMAPMA. Its aqueous solution properties in various salts and at various pH values were studied by measurements of the reduced viscosity and intrinsic viscosity. The reduced viscosity and intrinsic viscosity of this cationic polyelectrolyte were related to the types and concentrations of the added salts. The tendency of the salt effect was similar to that of other polyelectrolytes; that is, soft salt anions were more easily bound to the quaternary ammonium (R4N+) of MIQEDMAPMA than hard salt anions. Some salt ions strongly attracted the quaternary ammonium of the cationic polymeric side chain for the agglomeration of the polymers. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 89: 2261–2269, 2003  相似文献   

2.
The grafting of ADPEA onto natural rubber was executed with UV radiation. Benzoyl peroxide was used to initiate the free‐radical grafting copolymerization. Natural rubber‐graftN‐(4‐aminodiphenylether) acrylamide (NR‐g‐ADPEA) was characterized with an IR technique. The ultrasonic velocities of both longitudinal and shear waves were measured in thermoplastic discs of NBR vulcanizates as a function of aging time. Ultrasonic velocity measurements were taken at 2 MHz ultrasonic frequency using the pulse echo method. We studied the effect of aging on the mechanical properties and the swelling and extraction phenomena for acrylonitrile–butadiene copolymer (NBR) vulcanizates, which contained the prepared NR‐g‐ADPEA and a commercial antioxidant, N‐isopropyl‐N′‐phenyl‐p‐phenylenediamine. The prepared antioxidant enhanced both the mechanical properties of the NBR vulcanizates and the permanence of the ingredients in these vulcanizates. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

3.
A hydrated, 1% by weight Na‐kaolinite suspension in deionized water was prepared, completely characterized, and reproducible measures of flocculation efficiency were validated. Flocculation tests of copolymers of acrylamide (AM) with dimethylacrylamide (DMA), methacrylamide (MeAM), or N‐t‐butylacrylamide (NTBAM) with 1% Na‐kaolinite suspensions gave average settling rate rates which decreased as the proportion of DMA, MeAM, or NTBAM in the copolymer increased. However, for a similar weight‐average molecular weight and slightly lower 〈rg〉, the copolymer from DMA‐co‐AM‐3 gave settling rates and supernatant turbidities comparable to similar types of commercial polymers. This new copolymer was also more resistant to changes in pH or the presence of an electrolyte than were the tested commercial polymers. Cationic derivatives of the new copolymers gave lower average settling rates and higher supernatant turbidities than those of Percol 721 (cationic PAM), probably because of their lower charge densities. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 84: 2090–2108, 2002; DOI 10.1002/app.10562  相似文献   

4.
Poly(acrylamide/dimethyl diallyl ammonium chloride) (PAD) samples were synthesized via solution polymerization. Hydrolyzed poly(acrylamide/dimethyl diallyl ammonium chloride) (HPAD) was prepared with NaOH as a hydrolyzation agent. Both PAD and HPAD were characterized by IR and NMR spectroscopy. Na2SO3 was added to slow down the decomposition of HPAD under elevated temperatures. The rheological and filtrate properties of the drilling fluids were investigated in saltwater mud. We found that 30 wt % NaCl hampered those properties. Increasing HPAD's percentage in the clay suspensions reduced the filtrate volumes and whereas increased the viscosities. A greater amount of HPAD added to the saltwater mud resulted in a smaller particle medium size and caused the ζ potential to move toward a less negative value. The hydrolysis degree of HPAD needed to be controlled in a proper range to optimize the reduction in filtration. Scanning electron microscopy photos showed that HPAD was distributed in the surface of the high‐temperature and high‐pressure filter cake as bridges. The drilling fluids with 3.5 wt % HPAD had an excellent tolerance to 30 wt % NaCl and a temperature of 200°C. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 41581.  相似文献   

5.
The aqueous solution properties of an ampholytic naphthalene-labelled acrylamide/N,N-dimethyl maleimido propyl ammonium propane sulphonate copolymer, poly(ADMMAPS)/NA, are examined in this study by measuring reduced and intrinsic viscosities. The intrinsic viscosity of the poly(ADMMAPS)/NA is related to the type and concentration of the salt added. This behaviour, resulting from the associations of the polymer chains, is in contrast with cationic and anionic polyelectrolytes. Smaller anions, such as F-, with a common cation (K+) are found to be the most difficult to be bound to the sulphonate group, indicating that a lower intrinsic viscosity of the poly(ADMMAPS)/NA would be found in aqueous KF solution. Smaller cations, such as Li+, with a common anion (Cl-) are found to be the most difficult to be bound to the quaternary ammonium group, indicating that a lower intrinsic viscosity of the poly(ADMMAPS)/NA would be found in LiCl salt aqueous solution. Because a naphthalene label was introduced into the polymer, the behaviour of the solution properties of the poly(ADMMAPS)/NA could be clearly defined in term of macroscopic and microscopic analysis. The models proposed in this study can account for the poly(ADMMAPS)/NA solution viscosities. © 1998 SCI.  相似文献   

6.
The properties of a novel cationic, naphthalene-labeled cationic poly(dimethyl sulfate quaternized acrylamide/N,N-dimethylaminopropylmaleimide copolymer), poly-(DSQADMAPM)/NA in aqueous solution are examined in this study, measuring intrinsic viscosity, reduced viscosity, and ionic strength. This cationic poly(DSQADMAPM)/NA’s intrinsic viscosity is dependent on the type and concentration of salt added to the aqueous solution. The intrinsic viscosity behavior of the cationic poly(DSQADMAPM)/NA resulting from the electrostatic repulsive force of the polymer chain is contrasted with polyampholyte. Smaller anions such as F with a common cation (K+) are found to be the most difficult to be bound to the end group, indicating that a higher intrinsic viscosity of the poly(DSQADMAPM)/NA would be in KF salt aqueous solution. Smaller cations such as Li+ with a common anion (Cl) are found to be the most difficult to be bound to the quaternary ammonium group, indicating that a higher intrinsic viscosity of the poly(DSQADMAPM)/NA would also be in LiCI salt aqueous solution. Models are proposed to account for the poly(DSQADMAPM)/NA solution viscometrics.  相似文献   

7.
A series of copolymers PDAMs were synthesized with varying monomer ratio of acrylamide (AM) and N,N‐dimethylaminoethyl methacrylate (DMAEMA). The resulting copolymer solution shows an interesting property of viscosity‐response which is CO2‐triggered and N2‐enabled. Tertiary amine groups of PDAMs experience a reversible transition between hydrophobic and hydrophilic state upon CO2 addition and its removal, which induced different rheological behavior. A combination of zeta‐potential, laser particle‐size analysis, and electrical conductivity analysis indicated that, when the monomer mole ratio of DMAEMA and AM is less than or equal to 3 : 7, the hydrophobic association structure between the copolymer molecules was destroyed by the leading of CO2 and caused a viscosity decrease in its solution. On the contrary, when the monomer mole ratio of DMAEMA and AM is more than 3 : 7, a more extended conformation due to the protonated tertiary amine groups is formed and the enhanced repulsive interactions among the copolymer molecule results in a rise of its solution viscosity. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2014 , 131, 40872.  相似文献   

8.
The effect of the deaeration process on the behavior of acrylonitrile–acrylamide copolymer solutions was demonstrated experimentally in detail. The corresponding mechanical properties and morphology of the resultant precursors were also been examined. It was found that the viscosity of the copolymer solutions at rest increased continuously with the deaeration time prolonged at every fixed temperature stage, but it considerably increased when the deaeration temperature decreased. The changes in the viscosity of the solutions at 80°C were less prominent than those of the solutions at 20°C in the beginning stage; beyond 60 min, the changes became remarkable. In concentrated copolymer solutions, the dimethyl sulfoxide composition weight percentage decreased with an increase in the deaeration temperature; the increasing temperature reduced the solvent power continuously at a high deaeration temperature, so more solid–elastic gels were formed in the solutions with an increase in time. At the same degree of vacuum, when the temperature varied from 20 to 80°C and, in particular, the deaeration temperature was beyond 60°C, the products from the copolymer seemed to have slightly narrower molecular weight distributions. With the deaeration temperature increasing, the tensile strength, elongation at break, and bulk density of the resultant precursors increased, but their fineness and coefficient of variation decreased. The morphology of the precursors was more compact and more round beyond 60°C than below 60°C. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

9.
Viscosity behavior of dimethyl formamide solutions of acrylonitrile–acrylamide copolymer is discussed. The rheological kinetics of the solutions was studied for contrast. It is shown that the solutions behave the same as Newtonian flow as the rotor speed goes beyond 6 revolutions min−1. With an increase of temperature, the apparent viscosity of acrylonitrile–acrylamide copolymer solutions shows a decreasing trend. The changes of the apparent flow–activation energy of solutions calculated by Arrhenius equation become less prominent along with the changes of the molecular weight of acrylonitrile–acrylamide copolymers. The apparent flow–activation energy of the copolymer solutions increases continuously with an increase of copolymer concentration. The viscosity of copolymer solutions decreases continuously at concentrations of KCl and NaCl up to 0.02 mol L−1 and then increases. The apparent flow–activation energy of acrylonitrile–acrylamide copolymer solutions shows an obvious trend of decrease with addition of alkali salts and the changes of the apparent flow–activation energy of solutions containing NaCl are more prominent than those of solutions containing KCl. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 97: 527–531, 2005  相似文献   

10.
Copolymer of N,N‐dimethylacrylamide (NNDAM) and sodium 2‐acrylamido‐2‐methylpropanesulfonate (NaAMPS) have been prepared by free‐radical copolymerization and characterized with the help of molecular weight, molecular weight distribution, intrinsic viscosity, and monomer ratio in the copolymer. The solution behavior of a copolymer containing 26.62 wt % NaAMPS is studied in different solvents, namely, water (W), dimethyl sulfoxide (DMSO), ethylene glycol (EG), and ethanol (EtOH). The reduced viscosity of the copolymer is highly dependent on the ionic strength of the copolymer solution. The reduced viscosity decreases as a function of solvent selection in the order W > DMSO > EtOH > EG. The shapes of the ηsp / C vs. C plots indicate the polyelectrolyte behavior of the copolymer, except for the case of EG solutions, where nonpolyelectrolyte behavior is observed. However, at a certain degree of ionization attained by adding W as cosolvent, the copolymer begins to demonstrate polyelectrolyte behavior. For this copolymer, there exists a minimum concentration of brine (NaCl, CaCl2, etc.) above which solution viscosity is not further reduced. The copolymer solution behaves as a power law fluid, and exhibits time‐dependent thixotropic behavior. The copolymer cannot regain its solution viscosity when allowed to shear at a constant rate for long period of time. The reduced viscosities of copolymer solutions increase with increasing temperature in W and DMSO, yet decreases with increasing temperature in EG. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 85: 1521–1529, 2002  相似文献   

11.
Vinyl chloride–diallyl phthalate (VC–DAP) suspension copolymerization was carried out in a 5‐L autoclave and 200‐mL stainless steel vessel at 45°C. The apparent reactivity ratios of VC–DAP suspension copolymerization system were calculated as rVC = 0.77 and rDAP = 0.37. It shows that VC–DAP copolymer contains no gel when the feed concentration of DAP (f) is lower than a critical concentration (fcr, inside the range of 0.466–0.493 mmol/mol VC at 80–85% conversion), the polymerization degree (DP) of copolymer increases with the increase of f and conversion. VC–DAP copolymer is composed of gel and sol fractions when f is larger than fcr. The DP of sol fraction decreases as f increases, but the gel content and the crosslinking density of gel increase. The gel content also increases as conversion increases. The results also show that the index of polydispersity of molecular weight of sol changes with f, a maximum value appears when f is close to fcr. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 77: 156–162, 2000  相似文献   

12.
Poly(diallyldimethyl ammonium chloride–vinyl trimethoxysilane) [P(DADMAC–VTMS)] and poly(diallyldimethyl ammonium chloride–acrylamide–vinyl trimethoxysilane) [P(DADMAC–AM–VTMS)], the latter a new cationically charged and hydrophobically modified flocculant, were obtained by radical polymerization initiated by potassium persulfate. The effects of the vinyl trimethoxysilane (VTMS) feed ratio on the intrinsic viscosity and solubility of the polymers were examined. The effects of the flocculants on turbidity removal, decolorization, and oil removal in water treatment were also studied. The introduction of VTMS increased the intrinsic viscosities of P(DADMAC–VTMS) and P(DADMAC–AM–VTMS) in comparison with the viscosities of poly(diallyldimethyl ammonium chloride) and poly(diallyldimethyl ammonium chloride–acrylamide), respectively, but reduced their solubilities. The introduction of VTMS also enhanced the flocculation properties of P(DADMAC–VTMS) and P(DADMAC–AM–VTMS), including turbidity removal, decolorization, and oil removal. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 84: 335–342, 2002; DOI 10.1002/app.10339  相似文献   

13.
The viscosity behavior of dimethyl sulfoxide (DMSO) solutions of acrylonitrile–ammonium itaconate copolymer is discussed. The intrinsic viscosity was determined by an Ubbelohde viscometer. It is shown that an increase in the viscosity of acrylonitrile–ammonium itaconate copolymer solutions with time is considerably reduced by mechanical mixing. The viscosity of copolymer solutions with dimethyl formamide as an additive decreases monotonically. The viscosity of copolymer solutions decreases with the addition of H2O, and as the content of H2O in DMSO goes beyond 3 wt %, the viscosity shows a trend toward increase. The intrinsic viscosity decreases quickly with the addition of NaCl. As the concentration of NaCl goes beyond 0.015 mol/L, the viscosity increases solwly. The viscosity of copolymer solutions containing sodium ethoxide and sodium hydroxide appears to increase after an initial drop. The viscosity of copolymer solutions upon the addition of diethylamine increases continuously with time and changes in viscosity become less prominent after a period of a few hours. The effects of bases on the tacticity of highly isotactic copolymers follow the order of base strength. Diethylamine is more effective for moderating the stabilization exotherm of highly isotactic copolymers compared to sodium ethoxide and sodium hydroxide. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 93: 2622–2626, 2004  相似文献   

14.
Poly (acrylamide‐dimethyldiallylammonium chloride) (PAM‐DMDAAC) particles have been prepared via dispersion polymerizations using poly (vinyl pyrrolidone) (PVP) as a steric stabilizer in ethanol‐water media. The monomer reactivity ratios of acrylamide (AM, r1) and dimethyldiallylammonium chloride (DMDAAC, r2) were determined as 6.664 and 0.120, respectively, which means that PAM‐DMDAAC is a nonideal copolymer. The effects of various polymerization parameters (e.g., concentration of monomer and initiator, medium polarity, the ratio of AM to DMDAAC, initial temperature and ethylene diamine tetraacetic acid disodium (EDTA)) on the intrinsic viscosity and conversion of copolymer have been investigated. The copolymer was characterized by FTIR and NMR. The optimum operating conditions for preparing PAM‐DMDAAC were determined as ethanol content 50%, Cmonomer (wt %) 40%, nAM: nDMDAAC 8 : 2, Cinitiator (wt %) 0.04% and initiate temperature 40°C. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

15.
Copolymers of acrylamide, 2‐acrylamide‐2‐methylpropanesulfate (AMPS), and hydrophobic monomer N‐arylalkylacrylamide (BAAM) were synthesized by free‐radical micellar copolymerization. The effects of the copolymer, BAAM, AMPS, and NaCl concentrations and the pH value on the apparent viscosity of the copolymers were studied. The solution viscosities increased sharply when the copolymer concentration was higher than the critical associating concentration. The apparent viscosities of aqueous solutions of poly(N‐arylalkylacrylamide‐co‐acrylamide‐co‐2‐acrylamide‐2‐methylpropanesulfate) (PBAMS) increased with increasing BAAM and AMPS concentrations. PBAMS exhibited good salt resistance. With increasing pH, the apparent viscosities first increased and then decreased. Dilute PBAMS solutions exhibited Newtonian behavior, whereas semidilute aqueous and salt solutions exhibited shear‐thickening behavior at a lower shear rate and pseudoplastic behavior at a higher rate. Upon the removal of shear, the aqueous solution viscosities recovered and became even greater than the original viscosity, but the salt solution viscosities could not recover instantaneously. The elastic properties of PBAMS solutions were more dominant than the viscous properties, and this suggested a significant buildup of a network structure. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 97: 316–321, 2005  相似文献   

16.
The transport properties of polyelectrolytes with N‐alkyl‐N,N‐dimethyl‐N‐(2‐hydroxypropyl)ammonium chloride pendent groups (alkyl = octyl, dodecyl, or cetyl) randomly distributed along a polysaccharide backbone (dextran) in solutions were studied with viscometry and conductometry. This investigation mainly considered the influence of the side‐chain length and the solvent polarity on the polyelectrolyte behavior in salt‐free aqueous solutions. The viscometric data indicated that all the copolymers exhibited polyelectrolyte behavior, and they were plotted in terms of the Rao equation. The viscosity of the polycation with dodecyl as alkyl substituent was also measured in solvent mixtures of water and methanol. The conductometric behavior of these cationic polysaccharides was found to be influenced by the alkyl side‐chain length and the solvent polarity. A nonlinear dependence of the equivalent conductivity on the dielectric constant of the solvent was observed. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

17.
In this study, the viscosity behavior and surface and interfacial activities of associative water‐soluble polymers, which were prepared by an aqueous micellar copolymerization technique from acrylamide and small amounts of N‐phenyl acrylamide (1.5 and 5 mol %), were investigated under various conditions, including the polymer concentration, shear rate, temperature, and salinity. The copolymer solutions exhibited increased viscosity due to intermolecular hydrophobic associations, as the solution viscosity of the copolymers increased sharply with increasing polymer concentration, especially above a critical overlap concentration. An almost shear‐rate‐independent viscosity (Newtonian plateau) was also displayed at high shear rates, and typical non‐Newtonian shear‐thinning behavior was exhibited at low shear rates and high temperatures. Furthermore, the copolymers exhibited high air–water and oil–water interfacial activities, as the surface and interfacial tensions decreased with increasing polymer concentration and salinity. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 89: 2290–2300, 2003  相似文献   

18.
将十二烷基(Dod)与十六烷基(Cet)以酯键方式分别引入经羧甲基化改性的香豆胶大分子骨架上,制备了基于香豆胶的疏水改性阴离子聚电解质衍生物(HmCmFG)。通过荧光探针芘的激发光谱、发射光谱以及紫外光谱研究了HmCmFG溶液中的疏水缔合行为,研究发现,此疏水缔合行为受到HmCmFG大分子上烷基链长与取代度以及溶液中小分子电解质NaCl的显著影响,增加烷基链长、提高烷基取代度或增大NaCl浓度均有利于溶液中疏水缔合微区的形成。采用黏度法研究了溶液中HmCmFG大分子与NaCl、表面活性剂十六烷基三甲基溴化铵(CTAB)及十二烷基硫酸钠(SDS)的相互作用。结果表明,Dod取代度小于5.7时,NaCl的加入以增强HmCmFG分子间缔合为主,溶液黏度增加,加入CTAB对于Dod取代度小于5.7时的溶液黏度值的影响高于SDS;Dod取代度为10.2或Cet取代度为6.3时,加入NaCl、CTAB或SDS后使HmCmFG分子内缔合占优势,溶液黏度降低。  相似文献   

19.
A novel hydrophobically modified and cationic flocculant poly(acrylamide‐methacryloxyethyltrimethyl ammonium chloride‐methacryloxypropyltrimethoxy silane) (P(AM‐DMC‐MAPMS)) was synthesized by inverse emulsion polymerization. The molecular structure of hydrophobically cationic polyacrylamide (HCPAM) was characterized by FTIR and 1H‐NMR. The effects of DMC and MAPMS feed ratio on intrinsic viscosity and solubility were measured. The effects of hydrophobically cationic flocculants on reactive brilliant red X‐3B solution and kaolin suspension were studied. It was found that the introduction of MAPMS could increase the intrinsic viscosities of P(AM‐DMC‐MAPMS) and enhance the flocculation properties to anionic dye solution and kaolin suspension, but reduced their water‐solubility. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

20.
Monomethoxy poly(ethylene glycol) acrylate (MPEGA) was synthesized through the esterification reaction of acrylic acid and methoxy poly(ethylene glycol)s (MPEGs) of different molecular weights. Then, MPEGA was copolymerized with methyl methacrylate, butyl acrylate, and β‐carboxyethyl acrylate neutralized with potassium hydroxide via conventional solution polymerization. In this way, a single‐ion conductive polymer‐electrolyte‐based antistatic agent (PEAA), in which potassium (K) ions were used as charge carriers, was obtained. The molecular structure, coordination effects between ether oxygen (EO) groups and K cations, ionic conductivity, and crystallization ability of the copolymer were characterized with Fourier transform infrared, conductivity measurements, polarizing optical microscopy, and differential scanning calorimetry, respectively. The crystallinity of the synthesized PEAA apparently decreased with the molecular weight of MPEG and the EO/K molar ratio decreasing, and this led to a corresponding enhancement of the conductivity. The dependence of the conductivity of the copolymer on temperature could be divided into different linear parts, and each was in good agreement with the Arrhenius equation. Moreover, the dependence of the conductivity on the relative humidity (RH) revealed that the PEAA could maintain high ionic conductivity (∼10−6 S/cm) even at the low RH of 10%. This implies the potential widespread application of PEAAs for the preparation of antistatic composites and especially poly(vinyl chloride)‐ and poly(methyl methacrylate)‐related composites because of their considerable miscibility. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号