Journal articles on the topic 'Brine'

To see the other types of publications on this topic, follow the link: Brine.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Brine.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Ajayi, Toluwaleke. "Investigation of PH effect in a mixture of basalt and iron on co2 sequestration in synthetic brines." International Journal of Advanced Geosciences 7, no. 2 (September 4, 2019): 112. http://dx.doi.org/10.14419/ijag.v7i2.29132.

Full text
Abstract:
CO2 sequestration in deep saline aquifers is a critical component of long-term storage options. It is suggested that the precipitation of mineral carbonates is mostly dependent on brine pH and is favoured above a basic pH of 9.0. However, brine pH will drop to acidic values once CO2 is injected into the brine. Therefore, there is a need to raise brine pH and maintain it stable. Synthetic brines were used here instead of natural brines because of the difficulty in obtaining and storing natural brines. Therefore, experiments were conducted to prepare a series of synthetic brines and to compare their suitability to natural brines for carbon sequestration. A typical formation rock (basalt) and a buffer solution (0.3M Tris buffer solution) were selected to buffer brine pH. The results show that synthetic brines prepared can be used as analogues to natural brines for carbon sequestration studies in terms of chemical composition and pH response. This study investigates the effect of iron ( ) in the pH of six synthetic brines prepared as analogue to oil-field brine by conducting a pH stability studies for CO2-brine experiment and CO2-basalt-brine experiment. In a subsequent step, studies were conducted to correlate how brine samples respond in the presence of basalt and the buffer solution. X-Ray powder Diffraction (XRD) analyses were also carried out to characterise the mineralogy of the synthetic brines. The result of the XRD confirmed that calcite was the major component that was dominated in the -brine–experiment while slight occurrence of calcite, iron oxyhydroxides and dolomite precipitated in the -rock-brine experiment. It was observed that ferric iron and its reaction with host rock (basalt) did not contribute to pH instability therefore making it suitable for precipitation of carbonate mineral while ferrous iron in the absence of host rock did not contribute to pH instability therefore making it also suitable for precipitation of carbonate mineral.
APA, Harvard, Vancouver, ISO, and other styles
2

Burnett, David. "Brine Management: Produced Water and Frac Flowback Brine." Journal of Petroleum Technology 63, no. 10 (October 1, 2011): 46–48. http://dx.doi.org/10.2118/1011-0046-jpt.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Joye, S. B., I. R. MacDonald, J. P. Montoya, and M. Peccini. "Geophysical and geochemical signatures of Gulf of Mexico seafloor brines." Biogeosciences Discussions 2, no. 3 (May 31, 2005): 637–71. http://dx.doi.org/10.5194/bgd-2-637-2005.

Full text
Abstract:
Abstract. Geophysical, temperature, and discrete depth-stratified geochemical data illustrate differences between an actively venting mud volcano and a relatively quiescent brine pool in the Gulf of Mexico along the continental slope. Geophysical data, including laser-line scan mosaics and sub-bottom profiles, document the dynamic nature of both environments. Temperature profiles, obtained by lowering a CTD into the brine fluid, show that the venting brine was at least 10°C warmer than the bottom water. At the brine pool, two thermoclines were observed, one directly below the brine-seawater interface and a second about one meter below the first. At the mud volcano, substantial temperature variability was observed, with the core brine temperature being either slightly (~2°C in 1997) or substantially (19°C in 1998) elevated above bottom water temperature. Geochemical samples were obtained using a device called the "brine trapper" and concentrations of dissolved gases, major ions and nutrients were determined using standard techniques. Both brines contained about four times as much salt as seawater and steep concentration gradients of dissolved ions and nutrients versus brine depth were apparent. Differences in the concentrations of calcium, magnesium and potassium between the two brine fluids suggests that the fluids are derived from different sources or that brine-sediment reactions are more important at the mud volcano than the brine pool. Substantial concentrations of methane and ammonium were observed in both brines, suggesting that fluids expelled from deep ocean brines are important sources of methane and dissolved inorganic nitrogen to the surrounding environment.
APA, Harvard, Vancouver, ISO, and other styles
4

Alwazeer, Duried, Menekşe Bulut, and Yasemin Çelebi. "Hydrogen-Rich Water Can Restrict the Formation of Biogenic Amines in Red Beet Pickles." Fermentation 8, no. 12 (December 14, 2022): 741. http://dx.doi.org/10.3390/fermentation8120741.

Full text
Abstract:
Fermented foods are considered the main sources of biogenic amines (BAs) in the human diet while lactic acid bacteria (LAB) are the main producers of BAs. Normal water (NW) and hydrogen-rich water (HRW) were used for preparing red beet pickles, i.e., NWP and HRWP, respectively. The formation of BAs, i.e., aromatic amines (tyramine, 2-phenylethylamine), heterocyclic amines (histamine, tryptamine), and aliphatic di-amines (putrescine), was analyzed in both beet slices and brine of NWPs and HRWPs throughout the fermentation stages. Significant differences in redox value (Eh7) between NWP and HRWP brine samples were noticed during the first and last fermentation stages with lower values found for HRWPs. Total mesophilic aerobic bacteria (TMAB), yeast–mold, and LAB counts were higher for HRWPs than NWPs for all fermentation stages. Throughout fermentation stages, the levels of all BAs were lower in HRWPs than those of NWPs, and their levels in brines were higher than those of beets. At the end of fermentation, the levels (mg/kg) of BAs in NWPs and HRWPs were, respectively: tyramine, 72.76 and 61.74 (beet) and 113.49 and 92.67 (brine), 2-phenylethylamine, 48.00 and 40.00 (beet) and 58.01 and 50.19 (brine), histamine, 67.89 and 49.12 (beet) and 91.74 and 70.92 (brine), tryptamine, 93.14 and 77.23 (beet) and 119.00 and 93.11 (brine), putrescine, 81.11 and 63.56 (beet) and 106.75 and 85.93 (brine). Levels of BAs decreased by (%): 15.15 and 18.35 (tyramine), 16.67 and 13.44 (2-phenylethylamine), 27.65 and 22.7 (histamine), 17.09 and 21.76 (tryptamine), and 21.64 and 19.5 (putrescine) for beet and brine, respectively, when HRW was used in pickle preparation instead of NW. The results of this study suggest that the best method for limiting the formation of BAs in pickles is to use HRW in the fermentation phase then replace the fermentation medium with a new acidified and brined HRW followed by a pasteurization process.
APA, Harvard, Vancouver, ISO, and other styles
5

Joye, S. B., I. R. MacDonald, J. P. Montoya, and M. Peccini. "Geophysical and geochemical signatures of Gulf of Mexico seafloor brines." Biogeosciences 2, no. 3 (October 28, 2005): 295–309. http://dx.doi.org/10.5194/bg-2-295-2005.

Full text
Abstract:
Abstract. Geophysical, temperature, and discrete depth-stratified geochemical data illustrate differences between an actively venting mud volcano and a relatively quiescent brine pool in the Gulf of Mexico along the continental slope. Geophysical data, including laser-line scan mosaics and sub-bottom profiles, document the dynamic nature of both environments. Temperature profiles, obtained by lowering a CTD into the brine fluid, show that the venting brine was at least 10°C warmer than the bottom water. At the brine pool, thermal stratification was observed and only small differences in stratification were documented between three sampling times (1991, 1997 and 1998). In contrast, at the mud volcano, substantial temperature variability was observed, with the core brine temperature being slightly higher than bottom water (by 2°C) in 1997 but substantially higher than bottom water (by 19°C) in 1998. Detailed geochemical samples were obtained in 2002 using a device called the "brine trapper" and concentrations of dissolved gases, major ions and nutrients were determined. Both brines contained about four times as much salt as seawater and steep concentration gradients of dissolved ions and nutrients versus brine depth were apparent. Differences in the concentrations of calcium, magnesium and potassium between the two brine fluids suggest that the fluids are derived from different sources, have different dilution/mixing histories, or that brine-sediment reactions are more important at the mud volcano. Substantial concentrations of methane, ammonium, and silicate were observed in both brines, suggesting that fluids expelled from deep ocean brines are important sources of these constituents to the surrounding environment.
APA, Harvard, Vancouver, ISO, and other styles
6

Merridew, Nancy. "Being of brine." Neurology 95, no. 23 (October 1, 2020): 1059–60. http://dx.doi.org/10.1212/wnl.0000000000010956.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

TREMBLAY, JEAN-FRANÇOIS. "BRINE INTO GOLD." Chemical & Engineering News 88, no. 30 (July 26, 2010): 22. http://dx.doi.org/10.1021/cen-v088n030.p022.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Del Bene, J. V., Gerhard Jirka, and John Largier. "Ocean brine disposal." Desalination 97, no. 1-3 (August 1994): 365–72. http://dx.doi.org/10.1016/0011-9164(94)00100-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Duran, Lena Ballone. "Investigating Brine Shrimp." Science Activities: Classroom Projects and Curriculum Ideas 40, no. 2 (January 2003): 30–34. http://dx.doi.org/10.1080/00368120309601119.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Schmid, R. M. "Lake Torrens brine." Hydrobiologia 158, no. 1 (January 1988): 267–69. http://dx.doi.org/10.1007/bf00026284.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Novak, Stephanie A., and Yoram Eckstein. "Hydrochemical Characterization of Brines and Identification of Brine Contamination in Aquifersa." Ground Water 26, no. 3 (May 1988): 317–24. http://dx.doi.org/10.1111/j.1745-6584.1988.tb00395.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Price, Robert J., Edward F. Melvin, and Jon W. Bell. "Postmortem Changes in Blast, Brine and Brine-Coil Frozen Albacore." Journal of Aquatic Food Product Technology 1, no. 1 (January 20, 1992): 67–84. http://dx.doi.org/10.1300/j030v01n01_08.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Hwang, Chiu-Chu, Yi-Chen Lee, Chung-Yung Huang, Hsien-Feng Kung, Hung-Hui Cheng, and Yung-Hsiang Tsai. "Effect of Brine Concentrations on the Bacteriological and Chemical Quality and Histamine Content of Brined and Dried Milkfish." Foods 9, no. 11 (November 3, 2020): 1597. http://dx.doi.org/10.3390/foods9111597.

Full text
Abstract:
In this research, the occurrence of hygienic quality and histamine in commercial brined and dried milkfish products, and the effects of brine concentrations on the quality of brined and dried milkfish, were studied. Brined and dried milkfish products (n = 20) collected from four retail stores in Taiwan were tested to investigate their histamine-related quality. Among them, five tested samples (25%, 5/20) had histamine contents of more than 5 mg/100 g, the United States Food and Drug Administration guidelines for scombroid fish, while two (10%, 2/20) contained 69 and 301 mg/100 g of histamine, exceeding the 50 mg/100 g potential hazard level. In addition, the effects of brine concentrations (0%, 3%, 6%, 9%, and 15%) on the chemical and bacteriological quality of brined and dried milkfish during sun-drying were evaluated. The results showed that the aerobic plate count (APC), coliform, water activity, total volatile basic nitrogen (TVBN), and histamine content values of the brined and dried milkfish samples decreased with increased brine concentrations, whereas those of salt content and thiobarbituric acid (TBA) increased with increasing brine concentrations. The milkfish samples prepared with 6% NaCl brine had better quality with respect to lower APC, TVBN, TBA, and histamine levels.
APA, Harvard, Vancouver, ISO, and other styles
14

Nasralla, Ramez A., and Hisham A. Nasr-El-Din. "Double-Layer Expansion: Is It a Primary Mechanism of Improved Oil Recovery by Low-Salinity Waterflooding?" SPE Reservoir Evaluation & Engineering 17, no. 01 (January 30, 2014): 49–59. http://dx.doi.org/10.2118/154334-pa.

Full text
Abstract:
Summary Literature review shows that improved oil recovery (IOR) by low-salinity waterflooding could be attributed to several mechanisms, such as sweep-efficiency improvement, interfacial-tension (IFT) reduction, multicomponent ionic exchange, and electrical-double-layer (EDL) expansion. Although these mechanisms might contribute to IOR by low-salinity water, they may not be the primary mechanism. Therefore, the main objective of this study is to investigate if the mechanism of EDL expansion could be the principal reason for IOR during low-salinity waterflooding. Low-salinity water results in a thicker EDL when compared to high-salinity water, so we tried to eliminate the effect of low-salinity brines on double-layer expansion to show to what extent IOR is related to EDL expansion caused by low-salinity water. The double-layer expansion is dependent on the electric surface charge, which is a function of the pH of brine; therefore, the pH levels of low-salinity brines were decreased in this study to provide low-salinity brines that can produce a thinner EDL, similar to high-salinity brines. ζ-potential measurements were performed on both rock/brine and oil/brine interfaces to demonstrate the effect of brine pH and salinity on EDL. Contact angle and coreflood experiments were conducted to test different brine salinities at different pH values, which could assess the effect of water salinity and pH on rock wettability and oil recovery, and hence involvement of EDL expansion in the IOR process. ζ-potential results in this study showed that decreasing the pH of low-salinity brines makes the electrical charges at both oil/brine and brine/rock interfaces slightly negative, which reduces the double-layer expansion caused by low-salinity brine. As a result, the rock becomes more oil-wet, which was confirmed by contact-angle measurements. Moreover, coreflood experiments indicated that injecting low-salinity brine at lower pH values recovered smaller amounts of oil when compared to the original pH because of the elimination of the low-salinity-water effect on the thickness of the double layer. In conclusion, this study demonstrates that expansion of the double layer is a dominant mechanism of oil-recovery improvement by low-salinity waterflooding.
APA, Harvard, Vancouver, ISO, and other styles
15

Wang, Z., Z. He, and H. Li. "Mass transfer dynamics during brining of rabbit meat." World Rabbit Science 25, no. 4 (December 28, 2017): 377. http://dx.doi.org/10.4995/wrs.2017.6687.

Full text
Abstract:
As a traditional processing method, brining is a preliminary, critical and even essential process for many traditional rabbit meat products in China. The aim of this work was to investigate mass transfer of rabbit meat brined in different salt concentration. Rabbit meat (Longissimus dorsi) was brined for 24 h in 5 brine solutions (5, 10, 15, 20 and 25% NaCl [w/w]). Results indicated that mass transfer and kinetics parameters were significantly affected by the brine concentration during brining. When brine concentration increased, the total and water weight changes decreased, whereas the sodium chloride weight changes increased. Higher brine concentrations resulted in a higher degree of protein denaturation and consequently gave lower process yields. Samples treated with higher brine concentrations obtained lower brining kinetic parameter values for total weight changes and water weight changes, whereas they acquired higher values for sodium chloride weight changes.
APA, Harvard, Vancouver, ISO, and other styles
16

Udoh, Tinuola, and Jan Vinogradov. "A Synergy between Controlled Salinity Brine and Biosurfactant Flooding for Improved Oil Recovery: An Experimental Investigation Based on Zeta Potential and Interfacial Tension Measurements." International Journal of Geophysics 2019 (July 4, 2019): 1–15. http://dx.doi.org/10.1155/2019/2495614.

Full text
Abstract:
In this study, we have investigated the effects of brine and biosurfactant compositions on crude-oil-rock-brine interactions, interfacial tension, zeta potential, and oil recovery. The results of this study show that reduced brine salinity does not cause significant change in IFT. However, addition of biosurfactants to both high and low salinity brines resulted in IFT reduction. Also, experimental results suggest that the zeta potential of high salinity formation brine-rock interface is positive, but oil-brine interface was found to be negatively charged for all solutions used in the study. When controlled salinity brine (CSB) with low salinity and CSB with biosurfactants were injected, both the oil-brine and rock-brine interfaces become negatively charged resulting in increased water-wetness and, hence, improved oil recovery. Addition of biosurfactants to CSB further increased electric double layer expansion which invariably resulted in increased electrostatic repulsion between rock-brine and oil-brine interfaces, but the corresponding incremental oil recovery was small compared with injection of low salinity brine alone. Moreover, we found that the effective zeta potential of crude oil-brine-rock systems is correlated with IFT. The results of this study are relevant to enhanced oil recovery in which controlled salinity waterflooding can be combined with injection of biosurfactants to improve oil recovery.
APA, Harvard, Vancouver, ISO, and other styles
17

Cho, H., P. B. Shepson, L. A. Barrie, J. P. Cowin, and R. Zaveri. "NMR Investigation of the Quasi-Brine Layer in Ice/Brine Mixtures." Journal of Physical Chemistry B 106, no. 43 (October 2002): 11226–32. http://dx.doi.org/10.1021/jp020449+.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Pales, AshleyR, ErinN Kinsey, Chunyan Li, Linlin Mu, Lingyun Bai, HeatherM Clifford, and ChristopheJ G. Darnault. "Rheological Properties of Silica Nanoparticles in Brine and Brine-Surfactant Systems." Journal of Nanofluids 6, no. 5 (October 1, 2017): 795–803. http://dx.doi.org/10.1166/jon.2017.1379.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Totten, ClayW. "4765912 Geothermal brine clarifier process for removing solids from geothermal brine." Geothermics 18, no. 3 (January 1989): 478. http://dx.doi.org/10.1016/0375-6505(89)90075-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

GARCÍA-GARCÍA, P., M. BRENES-BALBUENA, D. HORNERO-MÉNDEZ, A. GARCÍA-BORREGO, and A. GARRIDO-FERNÁNDEZ. "Content of Biogenic Amines in Table Olives." Journal of Food Protection 63, no. 1 (January 1, 2000): 111–16. http://dx.doi.org/10.4315/0362-028x-63.1.111.

Full text
Abstract:
Content of biogenic amines in flesh and brines of table olives was determined by high-pressure liquid chromatography analysis of their benzoyl derivatives. No biogenic amines were found in the flesh of fresh fruits at any stage of ripeness. Contents of biogenic amines in Spanish-style green or stored olives increased throughout the brining period but were always higher in the former. Putrescine was the amine found in the highest concentration. Small quantities of cadaverine were found in the samples taken after 3 months of brining. This compound and histamine, tyramine, and tryptamine were also found in samples taken after 12 months. Gordal cultivar showed the highest contents, followed by Manzanilla and Hojiblanca. No relationship was found between contents of biogenic amines and lactic acid production or table olive spoilages, although zapatera olives had considerably higher amounts than those brines that had undergone a normal process. Concentrations in directly brined olives were markedly lower than contents in Spanish-style olives. With respect to partition between flesh and brine, there was equilibrium between both media in the case of Spanish-style olives, whereas the contents in directly brined olives were higher in flesh than brine.
APA, Harvard, Vancouver, ISO, and other styles
21

Jia, Hu, Yao–Xi Hu, Shan–Jie Zhao, and Jin–Zhou Zhao. "The Feasibility for Potassium-Based Phosphate Brines To Serve as High-Density Solid-Free Well-Completion Fluids in High-Temperature/High-Pressure Formations." SPE Journal 24, no. 05 (November 8, 2018): 2033–46. http://dx.doi.org/10.2118/194008-pa.

Full text
Abstract:
Summary Many oil and gas resources in deep–sea environments worldwide are often located in high–temperature/high–pressure (HT/HP) and low–permeability reservoirs. The reservoir–pressure coefficient usually exceeds 1.6, with formation temperature greater than 180°C. Challenges are faced for well drilling and completion in these HT/HP reservoirs. A solid–free well–completion fluid with safety density greater than 1.8 g/cm3 and excellent thermal endurance is strongly needed in the industry. Because of high cost and/or corrosion and toxicity problems, the application of available solid–free well–completion fluids such as cesium formate brines, bromine brines, and zinc brines is limited in some cases. In this paper, novel potassium–based phosphate well–completion fluids were developed. Results show that the fluid can reach the maximum density of 1.815 g/cm3 at room temperature, which makes a breakthrough on the density limit of normal potassium–based phosphate brine. The corrosion rate of N80 steel after the interaction with the target phosphate brine at a high temperature of 180°C is approximately 0.1853 mm/a, and the regained–permeability recovery of the treated sand core can reach up to 86.51%. Scanning–electron–microscope (SEM) pictures also support the corrosion–evaluation results. The phosphate brine shows favorable compatibility with the formation water. The biological toxicity–determination result reveals that it is only slightly toxic and is environmentally acceptable. In addition, phosphate brine is highly effective in inhibiting the performance of clay minerals. The cost of phosphate brine is approximately 44 to 66% less than that of conventional cesium formate, bromine brine, and zinc brine. This study suggests that the phosphate brine can serve as an alternative high–density solid–free well–completion fluid during well drilling and completion in HT/HP reservoirs.
APA, Harvard, Vancouver, ISO, and other styles
22

Tang, Hui, Jin-Kui Ma, Lin Chen, Li-Wen Jiang, Jing Xie, Pao Li, and Jing He. "GC-MS Characterization of Volatile Flavor Compounds in Stinky Tofu Brine by Optimization of Headspace Solid-Phase Microextraction Conditions." Molecules 23, no. 12 (November 30, 2018): 3155. http://dx.doi.org/10.3390/molecules23123155.

Full text
Abstract:
This study optimized the headspace solid phase microextraction (HS-SPME) conditions for the analysis of the volatile flavor compounds of Chinese south stinky tofu brine by gas chromatography-mass spectrometry (GC-MS). The optimum HS-SPME conditions established were as follows: polar column CD-WAX, white 85 μm polyella extractor, extraction temperature 60 °C, equilibrium time 20 min, extraction time 40 min. Under these conditions, a total of 63 volatile flavor compounds in five stinky tofu brines were identified. The offensive odor of the stinky tofu may be derived from some of the volatile flavor compounds such as phenol, p-cresol, 3-methylindole, indole, acetic acid, propionic acid, isobutyric acid, n-butyric acid and 3-methylbutanoic acid. The volatile flavor substances data was examined by principal component analysis (PCA) to visualize the response patterns in the feature space of principal components (PC). PCA analysis results revealed that the Chengshifu brine (STB1) and Baise jingdian brine (STB4) are similar in PC 1, 2, and 3, and the two brines have a similar flavor. Results also indicate that the Huogongdian brine (STB2) and Wangcheng brine (STB3) can be grouped in the same class as they are similar in PC 3. However, PC 1, 2, and 3 of the Luojia brine (STB5) and other brands of brine are different as is the flavor.
APA, Harvard, Vancouver, ISO, and other styles
23

Fauziah, Cut A., Ahmed Z. Al-Yaseri, R. Beloborodov, Mohammed A. Q. Siddiqui, M. Lebedev, D. Parsons, H. Roshan, A. Barifcani, and S. Iglauer. "Carbon Dioxide/Brine, Nitrogen/Brine, and Oil/Brine Wettability of Montmorillonite, Illite, and Kaolinite at Elevated Pressure and Temperature." Energy & Fuels 33, no. 1 (December 2018): 441–48. http://dx.doi.org/10.1021/acs.energyfuels.8b02845.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Koleini, Mohammad Mehdi, Mohammad Hasan Badizad, Zahra Kargozarfard, and Shahab Ayatollahi. "Interactions between Rock/Brine and Oil/Brine Interfaces within Thin Brine Film Wetting Carbonates: A Molecular Dynamics Simulation Study." Energy & Fuels 33, no. 9 (July 29, 2019): 7983–92. http://dx.doi.org/10.1021/acs.energyfuels.9b00496.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Öztürk, Mustafa, and Büşra Gülşah Güncü. "Effect of Brine Calcium Concentration on the Surface Solubilization and Texture of Fresh Perline Mozzarella Cheese." Turkish Journal of Agriculture - Food Science and Technology 9, no. 4 (April 23, 2021): 650–54. http://dx.doi.org/10.24925/turjaf.v9i4.650-654.3764.

Full text
Abstract:
Softening of cheese surface is a common problem especially in brined cheeses. In this study, the effects of the brine calcium concentrations on the texture of fresh perline Mozzarella cheese were investigated. The compositions of cheeses were analyzed 2 weeks after production. Brine protein content were monitored at 2 and 4 week of storage. The effect of the brine calcium concentration on the texture and meltability of cheeses were monitored Texture Profile Analysis (TPA) and Schreiber meltability test at 2 and 4 weeks of storage. The decrease in brine calcium concentration increased the protein transfer from cheese to brine, leading to an increase in the moisture content of cheese. As the calcium concentration increased in brine, an increase in the hardness, and decrease in adhesiveness and meltability of the cheeses were observed during storage. In conclusion, softening/solubilization of the surface of fresh perline Mozzarella cheese can be prevented with increasing the brine calcium concentration.
APA, Harvard, Vancouver, ISO, and other styles
26

Sander, W., and H. J. Herbert. "NaC1 crystallization at the MgCl2/NaC1 solution boundary–a possible natural barrier to the transport of radionuclides." Mineralogical Magazine 49, no. 351 (April 1985): 265–70. http://dx.doi.org/10.1180/minmag.1985.049.351.13.

Full text
Abstract:
AbstractConcentration, conductivity, temperature, and flow logs from sixteen brine-filled shafts in northern Germany have shown that the brines in all former potash salt mines exhibit a very sharp stratification into lower Mg-rich brine, an upper layer of Na-rich brine, and groundwater at the top. Laboratory experiments have shown that, at the MgCl2-brine/NaCl-brine boundary, both solutions become oversaturated with regard to NaCl, due to diffusion processes. NaCl therefore crystallizes from the solutions and forms a salt plug in the boundary region, which considerably reduces further diffusion. It is concluded that the observed effects would also take place in shafts. The backfilling material helps to nucleate the halite crystals and provides a structure on which they might be supported. The results of these experiments show that the density boundaries in the brine bodies act as barriers against transport of matter while the formation of a halite plug growing independently at the MgCl2/NaCl-brine interface acts as an additional barrier.
APA, Harvard, Vancouver, ISO, and other styles
27

Zhang, Shiyao, Yue Xiao, Yongli Jiang, Tao Wang, Shengbao Cai, Xiaosong Hu, and Junjie Yi. "Effects of Brines and Containers on Flavor Production of Chinese Pickled Chili Pepper (Capsicum frutescens L.) during Natural Fermentation." Foods 12, no. 1 (December 25, 2022): 101. http://dx.doi.org/10.3390/foods12010101.

Full text
Abstract:
The effects of (fresh/aged) brine and (pool/jar) containers on the flavor characteristics of pickled chili peppers were investigated based on a multivariate analysis integrated with kinetics modeling. The results showed that the effect of brine on organic acid, sugar, and aroma was more dominant than that of containers, while free amino acids production was more affected by containers than brines. Chili pepper fermented using aged brine exhibited higher acidity (3.71–3.92) and sugar (7.92–8.51 mg/g) than that using fresh brine (respective 3.79–3.96; 6.50–9.25 mg/g). Besides, chili peppers fermented using pool containers showed higher free amino acids content (424.74–478.82 mg/100 g) than using a jar (128.77–242.90 mg/100 g), particularly with aged brine. As for aroma, the number of volatiles in aged brine was higher (88–96) than that in fresh brine (76–80). The contents of the esters, alcohols, and ketones were significantly higher in the aged brine samples than those in fresh brine (p < 0.05), while terpenes in chili pepper fermented using the pool were higher than those using the jar. In general, jar fermentation with aged brine contributed more flavor to pickled chili peppers than other procedures.
APA, Harvard, Vancouver, ISO, and other styles
28

GAILUNAS, K. M., K. E. MATAK, R. R. BOYER, C. Z. ALVARADO, R. C. WILLIAMS, and S. S. SUMNER. "Use of UV Light for the Inactivation of Listeria monocytogenes and Lactic Acid Bacteria Species in Recirculated Chill Brines." Journal of Food Protection 71, no. 3 (March 1, 2008): 629–33. http://dx.doi.org/10.4315/0362-028x-71.3.629.

Full text
Abstract:
Ready-to-eat meat products have been implicated in several foodborne listeriosis outbreaks. Microbial contamination of these products can occur after thermal processing when products are chilled in salt brines. The objective of this study was to evaluate UV radiation on the inactivation of Listeria monocytogenes and lactic acid bacteria in a model brine chiller system. Two concentrations of brine (7.9% [wt/wt] or 13.2% [wt/wt]) were inoculated with a ~6.0 log CFU/ml cocktail of L. monocytogenes or lactic acid bacteria and passed through a UV treatment system for 60 min. Three replications of each bacteria-and-brine combination were performed and resulted in at least a 4.5-log reduction in microbial numbers in all treated brines after exposure to UV light. Bacterial populations were significantly reduced after 5 min of exposure to UV light in the model brine chiller compared with the control, which received no UV light exposure (P &lt; 0.05). The maximum rate of inactivation for both microorganisms in treated brines occurred between minutes 1 and 15 of UV exposure. Results indicate that in-line treatment of chill brines with UV light reduces the number of L. monocytogenes and lactic acid bacteria.
APA, Harvard, Vancouver, ISO, and other styles
29

Hershey, David R., and Stanley A. Rice. "Brine Shrimp "Bioassay" Problems." American Biology Teacher 66, no. 7 (September 2004): 474–75. http://dx.doi.org/10.2307/4451723.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Hershey, David R. "Brine Shrimp “Bioassay” Problems." American Biology Teacher 66, no. 7 (September 2004): 474–75. http://dx.doi.org/10.1662/0002-7685(2004)066[0474:bsbp]2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Sarah, Quazi Sahely, Fatema Chowdhury Anny, and Mir Misbahuddin. "Brine shrimp lethality assay." Bangladesh Journal of Pharmacology 12, no. 2 (June 5, 2017): 5. http://dx.doi.org/10.3329/bjp.v12i2.32796.

Full text
Abstract:
<p>Brine shrimp lethality assay is an important tool for the preliminary cytotoxicity assay of plant extract and others based on the ability to kill a laboratory cultured larvae (nauplii). The nauplii were exposed to different concentrations of plant extract for 24 hours. The number of motile nauplii was calculated for the effectiveness of the extract. It is a simple, cost effective and requires small amount of test material.</p><p><strong>Video Clip of Methodology</strong>:</p><p>Brine Shrimp Lethality Assay: 4 min 9 sec <a href="https://www.youtube.com/v/QJY7SQogD0g">Full Screen</a> <a href="https://www.youtube.com/watch?v=QJY7SQogD0g">Alternate</a> </p>
APA, Harvard, Vancouver, ISO, and other styles
32

Ejeian, Mojtaba, Alexander Grant, Ho Kyong Shon, and Amir Razmjou. "Is lithium brine water?" Desalination 518 (December 2021): 115169. http://dx.doi.org/10.1016/j.desal.2021.115169.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Dole, E. "Brine Bulb Technology ABSTRACT." Proceedings of the Water Environment Federation 2014, no. 7 (October 1, 2014): 4259–60. http://dx.doi.org/10.2175/193864714815941847.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Hanson, B. "Dissolving CO2 in Brine." Science 337, no. 6101 (September 20, 2012): 1435. http://dx.doi.org/10.1126/science.337.6101.1435-c.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Kirkwood, R. "Badger braves the brine." Veterinary Record 165, no. 4 (July 25, 2009): 122. http://dx.doi.org/10.1136/vetrec.165.4.122-b.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Koefod, Scott. "Effect of Prewetting Brines and Mixing on Ice-Melting Rate of Salt at Cold Temperatures: New Tracer Dilution Method." Transportation Research Record: Journal of the Transportation Research Board 2613, no. 1 (January 2017): 71–78. http://dx.doi.org/10.3141/2613-09.

Full text
Abstract:
A novel test method has been developed to measure the ice-melting rate of deicers. The ice-melting rates of prewetted salt were determined by measuring the change in the concentration of chloride (Cl−) or magnesium or calcium cations (Mg2+ or Ca2+, respectively) in the ice melt as tracers. The method is substantially more precise than the SHRP H205.1 standard and has the further advantage of measuring ice-melting and salt dissolution rates simultaneously. Brines were preequilibrated with ice at −19.3°C (−2.7°F) and blended with solid salt to determine the effect of different prewetting brines on the ice-melting rate of the solid salt component only. The measured equilibrium ice-melting capacity of sodium chloride (NaCl) agreed well with the theoretical value calculated from the NaCl freezing point curve. Under a condition of no mixing, solid salt yielded 0.87% of its total available ice-melting capacity after 60 min when wetted with NaCl brine and 9.7% when wetted with calcium chloride (CaCl2) brine. Mixing raised the yield of ice melt to 27.1% and 50.5% after 60 min when wet with NaCl and CaCl2 brines, respectively. The CaCl2 brine was slightly more effective than the magnesium chloride (MgCl2) brine at enhancing the ice-melting rate of salt. The test method promises to be a useful tool for permitting a more precise optimization of prewetting brine composition, concentration, and brine-to-salt ratio at different temperatures. The method may also permit better determination of the cost-effectiveness of different prewetting strategies and provide deeper insights into the mechanism of chemical ice melting.
APA, Harvard, Vancouver, ISO, and other styles
37

Fatoba, Ojo O., Leslie F. Petrik, Wilson M. Gitari, and Emmanuel I. Iwuoha. "Fly ash-brine interactions: Removal of major and trace elements from brine." Journal of Environmental Science and Health, Part A 46, no. 14 (December 2011): 1648–66. http://dx.doi.org/10.1080/10934529.2011.623647.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Xevgenos, D., A. Vidalis, K. Moustakas, D. Malamis, and M. Loizidou. "Sustainable management of brine effluent from desalination plants: the SOL-BRINE system." Desalination and Water Treatment 53, no. 12 (July 4, 2014): 3151–60. http://dx.doi.org/10.1080/19443994.2014.933621.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Weerakone, W. M. S. B., R. C. K. Wong, and A. K. Mehrotra. "Single-Phase (Brine) and Two-Phase (DNAPL–Brine) Flows in Induced Fractures." Transport in Porous Media 89, no. 1 (March 17, 2011): 75–95. http://dx.doi.org/10.1007/s11242-011-9753-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Song, Hyeyeon, Eun-woo Moon, and Ji-Hyoung Ha. "Application of Response Surface Methodology Based on a Box-Behnken Design to Determine Optimal Parameters to Produce Brined Cabbage Used in Kimchi." Foods 10, no. 8 (August 20, 2021): 1935. http://dx.doi.org/10.3390/foods10081935.

Full text
Abstract:
The factors of brine time, concentration, and temperature, affect the high-quality production of brined cabbage used in Kimchi. Although changes in Kimchi cabbage quality depending on brine time and concentration have been reported, changes in brine temperature have not been explored. Here, we optimized the brine process considering specific conditions of temperature (15–25 °C), concentration (10–14%), and osmosis duration (14–18 h) affecting the characteristics such as pH, titratable acidity, soluble solid content, glucose, fructose, and lactic acid bacteria and mass transport (salt gain, water loss, and weight reduction). The optimal parameters were determined using multivariate statistical analysis using the Box–Behnken design combined with response surface methodology. For each response as qualitative characteristics, second order polynomial models were developed using multiple regression analysis. Analysis of variance was performed to check the adequacy and accuracy of the fitted models. The brine temperature and concentration affected salt gain and water loss; the optimal brining temperature, concentration, and time were 19.17 °C, 10.53%, and 15.38 h, respectively. Statistical regression analysis indicated that standardized brined cabbage can be produced efficiently using a brining tank at controllable temperature.
APA, Harvard, Vancouver, ISO, and other styles
41

Kuo, M. H., S. G. Moussa, and V. F. McNeill. "Modeling interfacial liquid layers on environmental ices." Atmospheric Chemistry and Physics 11, no. 18 (September 28, 2011): 9971–82. http://dx.doi.org/10.5194/acp-11-9971-2011.

Full text
Abstract:
Abstract. Interfacial layers on ice significantly influence air-ice chemical interactions. In solute-containing aqueous systems, a liquid brine may form upon freezing due to the exclusion of impurities from the ice crystal lattice coupled with freezing point depression in the concentrated brine. The brine may be segregated to the air-ice interface where it creates a surface layer, in micropockets, or at grain boundaries or triple junctions. We present a model for brines and their associated liquid layers in environmental ice systems that is valid over a wide range of temperatures and solute concentrations. The model is derived from fundamental equlibrium thermodynamics and takes into account nonideal solution behavior in the brine, partitioning of the solute into the ice matrix, and equilibration between the brine and the gas phase for volatile solutes. We find that these phenomena are important to consider when modeling brines in environmental ices, especially at low temperatures. We demonstrate its application for environmentally important volatile and nonvolatile solutes including NaCl, HCl, and HNO3. The model is compared to existing models and experimental data from literature where available. We also identify environmentally relevant regimes where brine is not predicted to exist, but the QLL may significantly impact air-ice chemical interactions. This model can be used to improve the representation of air-ice chemical interactions in polar atmospheric chemistry models.
APA, Harvard, Vancouver, ISO, and other styles
42

Bottomley, Dennis J., Ian D. Clark, Nicholas Battye, and Tom Kotzer. "Geochemical and isotopic evidence for a genetic link between Canadian Shield brines, dolomitization in the Western Canada Sedimentary Basin, and Devonian calcium-chloridic seawater." Canadian Journal of Earth Sciences 42, no. 11 (November 1, 2005): 2059–71. http://dx.doi.org/10.1139/e05-075.

Full text
Abstract:
Hypersaline brines of marine origin, deep within fractured crystalline rocks of the Canadian Shield, are characterized by calcium-chloride compositions with relatively low concentrations of magnesium and sulphate. Such chemistries among crustal fluids are very unusual, regardless of their origins, which makes it difficult to identify possible genetic associations between shield brines and other marine brines. Key conservative chemical and isotopic attributes of shield brines from the Yellowknife and Sudbury areas are similar, however, to those of fluid inclusions in sparry, late-stage hydrothermal dolomites in lower Paleozoic carbonates of the Western Canada Sedimentary Basin, suggesting that both fluid types originated from a residual seawater brine of probable Devonian age. This interpretation is supported by determination of a helium-4 model age for the Yellowknife brine that is consistent with a Devonian origin. Therefore, dolomitization was likely both a major source of Ca and a sink for about half of the initial Mg in the brine prior to its infiltration into the underlying basement rocks. Mass-balance considerations indicate that albitization of plagioclase, predominantly in basinal clastic sediments, was an equally important mineral source of Ca to the infiltrating brine, but interactions with carbonate and silicate rocks cannot account for the entire Ca inventory in the brine. A major contribution to the Ca budget must also have been initially provided by the evaporative concentration of seawater, which is only possible if it were enriched in calcium and depleted in sulphate compared with modern seawater, as was probably the case during Devonian time.
APA, Harvard, Vancouver, ISO, and other styles
43

Yutkin, M. P., C. J. Radke, and T. W. Patzek. "Chemical Compositions in Modified Salinity Waterflooding of Calcium Carbonate Reservoirs: Experiment." Transport in Porous Media 141, no. 2 (January 2022): 255–78. http://dx.doi.org/10.1007/s11242-021-01715-x.

Full text
Abstract:
AbstractModified or low-salinity waterflooding of carbonate oil reservoirs is of considerable economic interest because of potentially inexpensive incremental oil production. The injected modified brine changes the surface chemistry of the carbonate rock and crude oil interfaces and detaches some of adhered crude oil. Composition design of brine modified to enhance oil recovery is determined by labor-intensive trial-and-error laboratory corefloods. Unfortunately, limestone, which predominantly consists of aqueous-reactive calcium carbonate, alters injected brine composition by mineral dissolution/precipitation. Accordingly, the rock reactivity hinders rational design of brines tailored to improve oil recovery. Previously, we presented a theoretical analysis of 1D, single-phase brine injection into calcium carbonate-rock that accounts for mineral dissolution, ion exchange, and dispersion (Yutkin et al. in SPE J 23(01):084–101, 2018. 10.2118/182829-PA). Here, we present the results of single-phase waterflood-brine experiments that verify the theoretical framework. We show that concentration histories eluted from Indiana limestone cores possess features characteristic of fast calcium carbonate dissolution, 2:1 ion exchange, and high dispersion. The injected brine reaches chemical equilibrium inside the porous rock even at injection rates higher than 3.5 $$\times$$ × 10$$^{-3}$$ - 3 m s$$^{-1}$$ - 1 (1000 ft/day). Ion exchange results in salinity waves observed experimentally, while high dispersion is responsible for long concentration history tails. Using the verified theoretical framework, we briefly explore how these processes modify aqueous-phase composition during the injection of designer brines into a calcium-carbonate reservoir. Because of high salinity of the initial and injected brines, ion exchange affects injected concentrations only in high surface area carbonates/limestones, such as chalks. Calcium-carbonate dissolution only affects aqueous solution pH. The rock surface composition is affected by all processes.
APA, Harvard, Vancouver, ISO, and other styles
44

Fang, Chao, Shuyu Sun, and Rui Qiao. "Structure, Thermodynamics, and Dynamics of Thin Brine Films in Oil–Brine–Rock Systems." Langmuir 35, no. 32 (July 21, 2019): 10341–53. http://dx.doi.org/10.1021/acs.langmuir.9b01477.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Spongberg, Martin E., and Wesley P. James. "CONTROL OF NATURAL BRINE SPRINGS IN BRAZOS RIVER BASIN PART II: BRINE DISPOSAL." Journal of the American Water Resources Association 32, no. 3 (June 1996): 485–90. http://dx.doi.org/10.1111/j.1752-1688.1996.tb04046.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Keeley, D. F., and J. R. Meriwether. "Hydrocarbons From U.S. Gulf Coast Geopressured Brines." Journal of Energy Resources Technology 110, no. 3 (September 1, 1988): 177–82. http://dx.doi.org/10.1115/1.3231379.

Full text
Abstract:
U.S. Gulf Coast geopressured brines studied to date contain small amounts of C6+ hydrocarbons which are primarily aromatic in nature. They range from benzene to substituted anthracenes. In addition, the brines contain a variety of ions and light, C1 to C6, aliphatic hydrocarbons. The primarily aromatic hydrocarbon mixture was collected at −78.5°C and is referred to as a “cryocondensate.” It contains at least 95 different compounds and, from the carbon isotropic ratios, appears to be of terrestrial plant origin. For the only U.S. DOE geopressured energy design well studied for an extended period of time, i.e., the Glady’s McCall well, the concentration of the cryocondensate in the brine was observed to increase prior to the onset of oil production. It is postulated that the change in the brine cryocondensate concentration as a function of the cumulative brine volume produced from the wells results from an extraction of additional aromatic components from oil migrating into the production zone from adjacent shale. When sufficient oil has migrated, it is produced along with the brine.
APA, Harvard, Vancouver, ISO, and other styles
47

McKINNEY, JULIE, ROBERT C. WILLIAMS, GREGORY D. BOARDMAN, JOSEPH D. EIFERT, and SUSAN S. SUMNER. "Dose of UV Light Required To Inactivate Listeria monocytogenes in Distilled Water, Fresh Brine, and Spent Brine." Journal of Food Protection 72, no. 10 (October 1, 2009): 2144–50. http://dx.doi.org/10.4315/0362-028x-72.10.2144.

Full text
Abstract:
The purpose of this research was to establish the dose of UV light (253.7 nm) needed to inactivate Listeria monocytogenes in distilled water, fresh brine (9% NaCl), spent brine, and diluted (5, 35, and 55%) spent brine, using uridine as a chemical actinometer. Strains N1-227 (isolated from hot dog batter), N3-031 (isolated from turkey franks), and R2-499 (isolated from meat) were mixed in equal proportions and suspended in each solution prepared so as to contain 10−4 M uridine. Samples were irradiated in sterile quartz cells for 0, 5, 10, 15, 20, 25, or 30 min. Inactivation was evaluated by serially diluting samples in 0.1% peptone, by surface plating in duplicate onto modified Oxford agar and Trypticase soy agar with yeast extract, and by enrichment in brain heart infusion broth, followed by incubation at 37°C for 24 to 48 h. For dose measurements, the absorbance (262 nm) was measured before and after irradiation. Differences were observed in population estimates depending on the solution (P ≤ 0.05). Reductions were as follows from greatest to least: water &gt; fresh brine &gt; 5% spent brine &gt; 35% spent brine &gt; 55% spent brine &gt; undiluted spent brine. UV light did not significantly reduce populations suspended in spent brine solutions. L. monocytogenes decreased to below the detection limit (1 log CFU/ml) at doses greater than 33.2 mJ/cm2 in water and at doses greater than 10.3 mJ/cm2 in fresh brine. Knowledge of UV dosing required to control L. monocytogenes in brines similar to those used for ready-to-eat meat processing will aid manufacturers in establishing appropriate food safety interventions for these products.
APA, Harvard, Vancouver, ISO, and other styles
48

Uttaro, B., M. Badoni, S. Zawadski, and C. O. Gill. "Effects of the pressure, flow rate and delivered volume of brine on the distributions of brine and bacteria in brine-injected meat." Food Control 22, no. 2 (February 2011): 180–85. http://dx.doi.org/10.1016/j.foodcont.2010.06.016.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Abu-Al-Saud, Moataz, Amani Al-Ghamdi, Subhash Ayirala, and Mohammed Al-Otaibi. "A Surface Complexation Model of Alkaline-SmartWater Electrokinetic Interactions in Carbonates." E3S Web of Conferences 146 (2020): 02003. http://dx.doi.org/10.1051/e3sconf/202014602003.

Full text
Abstract:
Understanding the effect of injection water chemistry is becoming crucial, as it has been recently shown to have a major impact on oil recovery processes in carbonate formations. Various studies have concluded that surface charge alteration is the primary mechanism behind the observed change of wettability towards water-wet due to SmartWater injection in carbonates. Therefore, understanding the surface charges at brine/calcite and brine/crude oil interfaces becomes essential to optimize the injection water compositions for enhanced oil recovery (EOR) in carbonate formations. In this work, the physicochemical interactions of different brine recipes with and without alkali in carbonates are evaluated using Surface Complexation Model (SCM). First, the zeta-potential of brine/calcite and brine/crude oil interfaces are determined for Smart Water, NaCl, and Na2SO4 brines at fixed salinity. The high salinity seawater is also included to provide the baseline for comparison. Then, two types of Alkali (NaOH and Na2CO3) are added at 0.1 wt% concentration to the different brine recipes to verify their effects on the computed zeta-potential values in the SCM framework. The SCM results are compared with experimental data of zeta-potentials obtained with calcite in brine and crude oil in brine suspensions using the same brines and the two alkali concentrations. The SCM results follow the same trends observed in experimental data to reasonably match the zeta-potential values at the calcite/brine interface. Generally, the addition of alkaline drives the zeta-potentials towards more negative values. This trend towards negative zeta-potential is confirmed for the Smart Water recipe with the impact being more pronounced for Na2CO3 due to the presence of divalent anion carbonate (CO3)-2. Some discrepancy in the zeta-potential magnitude between the SCM results and experiments is observed at the brine/crude oil interface with the addition of alkali. This discrepancy can be attributed to neglecting the reaction of carboxylic acid groups in the crude oil with strong alkali as NaOH and Na2CO3. The novelty of this work is that it clearly validates the SCM results with experimental zeta-potential data to determine the physicochemical interaction of alkaline chemicals with SmartWater in carbonates. These modeling results provide new insights on defining optimal SmartWater compositions to synergize with alkaline chemicals to further improve oil recovery in carbonate reservoirs.
APA, Harvard, Vancouver, ISO, and other styles
50

Pominski, J., M. A. Mishkin, H. M. Pearce, and P. Wan. "Salting and Hydration of Partially Defatted Peanuts." Peanut Science 23, no. 2 (July 1, 1996): 94–97. http://dx.doi.org/10.3146/i0095-3679-23-2-6.

Full text
Abstract:
Abstract Salt was added to partially defatted peanuts prior to roasting, by hydration in salt and dextrose solutions (brines), and by adding salt after hydration in water. Brine uptake was less than that of water for the same time intervals; i.e., at 2-min hydration time, pickup for water was 27.6% as compared with 18.49% for 15% brine. These procedures resulted in desired salt contents of 2.66 and 2.79%, respectively, in roasted partially defatted peanuts. The brine method was the more efficient procedure for the addition of salt and increased the amount of whole roasted partially defatted peanuts as much as 35.46%.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography