@article{franks_jeltema_luck_beckley_foegeding_vinyard_2020, title={Morphological and masticatory performance variation of mouth behavior groups}, volume={51}, ISSN={["1745-4603"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-85074800855&partnerID=MN8TOARS}, DOI={10.1111/jtxs.12483}, abstractNote={Abstract}, number={2}, journal={JOURNAL OF TEXTURE STUDIES}, author={Franks, Erin M. and Jeltema, Melissa and Luck, Paige J. and Beckley, Jacqueline and Foegeding, E. Allen and Vinyard, Christopher J.}, year={2020}, month={Apr}, pages={343–351} } @article{wagoner_luck_foegeding_2016, title={Caramel as a Model System for Evaluating the Roles of Mechanical Properties and Oral Processing on Sensory Perception of Texture}, volume={81}, ISSN={["1750-3841"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84959510916&partnerID=MN8TOARS}, DOI={10.1111/1750-3841.13237}, abstractNote={Abstract}, number={3}, journal={JOURNAL OF FOOD SCIENCE}, author={Wagoner, Ty B. and Luck, Paige J. and Foegeding, E. Allen}, year={2016}, month={Mar}, pages={S736–S744} } @article{wilson_luck_woods_foegeding_morgenstern_2016, title={Comparison of jaw tracking by single video camera with 3D electromagnetic system}, volume={190}, ISSN={["1873-5770"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84975832612&partnerID=MN8TOARS}, DOI={10.1016/j.jfoodeng.2016.06.008}, abstractNote={The breakdown of food during chewing is both a physical process and a sensory experience. In trying to understand the differing sensory responses of consumers to food products it is useful to be able to measure their physical chewing action. In this paper we report the results of a comparison between a simple 2D video jaw tracking method with a 3D method using the JT-3D™ Jaw Tracker (BioRESEARCH Assoc., Inc., Milwaukee), on four model gel systems. The video and JT3D systems gave similar values for number of chews, chewing time, chewing cycle time, chewing frequency, opening velocity, and proportion of crossed/crescent/circle shaped cycles. Although timing of the three phases of a chew and vertical and lateral movement were different between the two methods, the effect of the different model gels on these parameters are similar in direction by the two methods for vertical and lateral movement, and opening and closing velocity and closing time. Our study demonstrates that for sensory evaluation of foods and consumer preference, where large numbers of participants are required to cover the variation in human populations, the simple 2D video method allows jaw movement to be tracked with sufficient accuracy to detect the effects produced by different foods.}, journal={JOURNAL OF FOOD ENGINEERING}, author={Wilson, Arran and Luck, Paige and Woods, Carmel and Foegeding, E. Allen and Morgenstern, Marco}, year={2016}, month={Dec}, pages={22–33} } @article{luck_varum_foegeding_2015, title={Charge related astringency of chitosans}, volume={48}, ISSN={["1873-7137"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84924240333&partnerID=MN8TOARS}, DOI={10.1016/j.foodhyd.2015.02.024}, abstractNote={Chitosan is a biopolymer that can be used in food applications such as forming edible coatings and nanoparticles. It is a linear polysaccharide composed of neutral (acetylated) and positively (deacetylated) sugar units and can be prepared with varying content of the two units. One factor limiting the use of chitosan is causing an unacceptable level of astringency. We hypothesized that astringency was related to the positive charge density and that decreasing the charge density would decrease astringency. Chitosans with fraction of acetylated units (FA) ranging from 0.01 to 0.49 (percentage degree of acetylation from 1 to 49) were analyzed for astringency using a time intensity sensory method. In addition, the ability of chitosans to precipitate saliva proteins was determined. Increasing FA from 0.01 to 0.49 caused a decrease in maximum astringency sensation from 10.9 to 7.2. Mucins and other saliva proteins were precipitated by chitosans but a simple association with astringency intensity was not observed. It can be concluded that altering the charge density of chitosans is a means to alter chitosan astringency.}, journal={FOOD HYDROCOLLOIDS}, author={Luck, Paige and Varum, Kjell M. and Foegeding, E. Allen}, year={2015}, month={Jun}, pages={174–178} } @article{luck_vardhanabhuti_yong_laundon_barbano_foegeding_2013, title={Comparison of functional properties of 34% and 80% whey protein and milk serum protein concentrates}, volume={96}, ISSN={["0022-0302"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84882905470&partnerID=MN8TOARS}, DOI={10.3168/jds.2013-6617}, abstractNote={This study compared the functional properties of serum protein concentrate (SPC) with whey protein concentrate (WPC) made from the same milk and with commercial WPC. The experimental SPC and WPC were produced at 34% or 80% protein from the same lot of milk. Protein contents of WPC and SPC were comparable; however, fat content was much lower in SPC compared with WPC and commercial WPC. The effect of drying methods (freeze vs. spray drying) was studied for 34% WPC and SPC. Few differences due to drying method were found in turbidity and gelation; however, drying method made a large difference in foam formation for WPC but not SPC. Between pH 3 and 7, SPC was found to have lower turbidity than WPC; however, protein solubility was similar between SPC and WPC. Foaming and gelation properties of SPC were better than those of WPC. Differences in functional properties may be explained by differences in composition and extent of denaturation or aggregation.}, number={9}, journal={JOURNAL OF DAIRY SCIENCE}, author={Luck, P. J. and Vardhanabhuti, B. and Yong, Y. H. and Laundon, T. and Barbano, D. M. and Foegeding, E. A.}, year={2013}, month={Sep}, pages={5522–5531} } @inbook{foegeding_luck_vardhanabhuti_2011, title={Milk Protein Products: Whey Protein Products}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-85016959262&partnerID=MN8TOARS}, DOI={10.1016/B978-0-12-374407-4.00350-2}, abstractNote={Whey proteins, a coproduct of cheesemaking and casein manufacture, represent a rich and heterogeneous mixture of proteins with a broad range of nutritional and functional properties. Significant progress in the utilization of whey has been made in the past 30 years, especially with the current research findings on the health benefits of whey proteins and their derivatives. The future growth of whey utilization is expected to be led by the industry’s increasing focus on nutritional products, particularly in the dietary, sports, and clinical segments of the market. Functional properties of whey proteins in foods include solubility, dispersibility, heat stability, network formation (gels and edible films), and surface activity (emulsions and foams). This article discusses the production, processing, and applications of whey proteins and their derivatives.}, booktitle={Encyclopedia of Dairy Sciences: Second Edition}, author={Foegeding, E.A. and Luck, P. and Vardhanabhuti, B.}, year={2011}, pages={873–878} } @article{kelly_vardhanabhuti_luck_drake_osborne_foegeding_2010, title={Role of protein concentration and protein-saliva interactions in the astringency of whey proteins at low pH}, volume={93}, ISSN={["1525-3198"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-77952051521&partnerID=MN8TOARS}, DOI={10.3168/jds.2009-2853}, abstractNote={Whey protein beverages are adjusted to pH <4.5 to enhance clarity and stability, but this pH level is also associated with increased astringency. The goal of this investigation was to determine the effects of protein concentration on astringency and interactions between whey and salivary proteins. Whey protein beverages containing 0.25 to 13% (wt/wt) beta-lactoglobulin and 0.017% (wt/wt) sucralose at pH 2.6 to 4.2 were examined using descriptive sensory analysis. Controls were similar pH phosphate buffers at phosphate concentrations equivalent to the amount of phosphoric acid required to adjust the pH of the protein solution. Changes in astringency with protein concentration depended on pH. At pH 3.5, astringency significantly increased with protein concentration from 0.25 to 4% (wt/wt) and then remained constant from 4 to 13% (wt/wt). Conversely, at pH 2.6, astringency decreased with an increase in protein concentration [0.5-10% (wt/wt)]. This suggests a complex relationship that includes pH and buffering capacity of the beverages. Furthermore, saliva flow rates increased with increasing protein concentrations, showing that the physiological conditions in the mouth change with protein concentration. Maximum turbidity of whey protein-saliva mixtures was observed between pH 4.6 and 5.2. Both sensory evaluation and in vitro study of interactions between beta-LG and saliva indicate that astringency of whey proteins is a complex process determined by the extent of aggregation occurring in the mouth, which depends on the whey protein beverage pH and buffering capacity in addition to saliva flow rate.}, number={5}, journal={JOURNAL OF DAIRY SCIENCE}, author={Kelly, M. and Vardhanabhuti, B. and Luck, P. and Drake, M. A. and Osborne, J. and Foegeding, E. A.}, year={2010}, month={May}, pages={1900–1909} } @article{vardhanabhuti_kelly_luck_drake_foegeding_2010, title={Roles of charge interactions on astringency of whey proteins at low pH}, volume={93}, ISSN={["1525-3198"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-77952066147&partnerID=MN8TOARS}, DOI={10.3168/jds.2009-2780}, abstractNote={Whey proteins are a major ingredient in sports drink and functional beverages. At low pH, whey proteins are astringent, which may be undesirable in some applications. Understanding the astringency mechanism of whey proteins at low pH could lead to developing ways to minimize the astringency. This study compared the astringency of beta-lactoglobulin (beta-LG) at low pH with phosphate buffer controls having the same amount of phosphate and at similar pH. Results showed that beta-LG samples were more astringent than phosphate buffers, indicating that astringency was not caused by acid alone and that proteins contribute to astringency. When comparing among various whey protein isolates (WPI) and lactoferrin at pH 3.5, 4.5, and 7.0, lactoferrin was astringent at pH 7.0 where no acid was added. In contrast, astringency of all WPI decreased at pH 7.0. This can be explained by lactoferrin remaining positively charged at pH 7.0 and able to interact with negatively charged saliva proteins, whereas the negatively charged WPI would not interact. Charge interactions were further supported by beta-LG or lactoferrin and salivary proteins precipitating when mixed at conditions where beta-LG, lactoferrin, or saliva themselves did not precipitate. It can be concluded that interactions between positively charged whey proteins and salivary proteins play a role in astringency of proteins at low pH.}, number={5}, journal={JOURNAL OF DAIRY SCIENCE}, author={Vardhanabhuti, B. and Kelly, M. A. and Luck, P. J. and Drake, M. A. and Foegeding, E. A.}, year={2010}, month={May}, pages={1890–1899} } @article{beecher_drake_luck_foegeding_2008, title={Factors regulating astringency of whey protein beverages}, volume={91}, ISSN={["1525-3198"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-47149106271&partnerID=MN8TOARS}, DOI={10.3168/jds.2008-1083}, abstractNote={A rapidly growing area of whey protein use is in beverages. There are 2 types of whey protein-containing beverages: those at neutral pH and those at low pH. Astringency is very pronounced at low pH. Astringency is thought to be caused by compounds in foods that bind with and precipitate salivary proteins; however, the mechanism of astringency of whey proteins is not understood. The effect of viscosity and pH on the astringency of a model beverage containing whey protein isolate was investigated. Trained sensory panelists (n = 8) evaluated the viscosity and pH effects on astringency and basic tastes of whey protein beverages containing 6% wt/vol protein. Unlike what has been shown for alum and polyphenols, increasing viscosity (1.6 to 7.7 mPa.s) did not decrease the perception of astringency. In contrast, the pH of the whey protein solution had a major effect on astringency. A pH 6.8 whey protein beverage had a maximum astringency intensity of 1.2 (15-point scale), whereas that of a pH 3.4 beverage was 8.8 (15-point scale). Astringency decreased between pH 3.4 and 2.6, coinciding with an increase in sourness. Decreases in astringency corresponded to decreases in protein aggregation as observed by turbidity. We propose that astringency is related to interactions between positively charged whey proteins and negatively charged saliva proteins. As the pH decreased between 3.4 and 2.6, the negative charge on the saliva proteins decreased, causing the interactions with whey proteins to decrease.}, number={7}, journal={JOURNAL OF DAIRY SCIENCE}, author={Beecher, J. W. and Drake, M. A. and Luck, P. J. and Foegeding, E. A.}, year={2008}, month={Jul}, pages={2553–2560} } @article{luck_foegeding_2008, title={The role of copper in protein foams}, volume={3}, ISSN={["1557-1858"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-43749110248&partnerID=MN8TOARS}, DOI={10.1007/s11483-008-9060-0}, number={2}, journal={FOOD BIOPHYSICS}, author={Luck, Paige J. and Foegeding, Edward A.}, year={2008}, month={Jun}, pages={255–260} } @article{whetstine_luck_drake_foegeding_gerard_barbano_2007, title={Characterization of flavor and texture development within large (291 kg) blocks of Cheddar cheese}, volume={90}, ISSN={["1525-3198"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-35748980863&partnerID=MN8TOARS}, DOI={10.3168/jds.2006-755}, abstractNote={Cheddar cheese is a natural product that has a variable flavor and texture profile. Many companies produce 291-kg blocks of Cheddar cheese, which are subsequently cut and shipped, or stored and subsequently cut. Previous research has shown that compositional differences exist within 291-kg blocks and that these differences may influence flavor and texture development. The objectives of this study were to systematically characterize flavor and texture differences within 291-kg blocks. On 2 different occasions, a 291-kg block was manufactured at each of 4 manufacturing facilities. After 7 d, the 291-kg blocks were sliced into sixteen 18-kg sample portions using a predetermined diagram, and each portion was labeled appropriately (outer corner, inner corner, etc.) and stored at 7 degrees C. Cheese from different locations within the 291-kg blocks was evaluated at 1, 4, 8, and 12 mo. At each time point, two 18-kg portions representing an inside and outside location with the 291-kg block cross-section (from inside to outside) were sampled. The moisture content was lower in the inner than outer locations within the 291-kg blocks. Protein hydrolysis was higher in the inner location and inner locations developed aged Cheddar flavors sulfur, nutty, and brothy more rapidly than the outer locations. However, plant-to-plant differences in aging were often larger than differences caused by block location. These differences were due to differences in cheese manufacturing practices among plants. Dynamic headspace results for flavor volatiles were consistent with descriptive sensory flavor results, documenting differences between inner and outer locations within 291-kg blocks. The inner locations were more fracturable and the outer locations were more cohesive and had more residual in the mouth. Inner locations had greater fracture strain than outer locations. Documenting the differences in aging of 291-kg blocks of Cheddar cheese is important in understanding how to make a consistent high-quality Cheddar cheese.}, number={7}, journal={JOURNAL OF DAIRY SCIENCE}, author={Whetstine, M. E. Carunchia and Luck, P. J. and Drake, M. A. and Foegeding, E. A. and Gerard, P. D. and Barbano, D. M.}, year={2007}, month={Jul}, pages={3091–3109} } @article{foegeding_luck_davis_2006, title={Factors determining the physical properties of protein foams}, volume={20}, ISSN={["0268-005X"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-25844500998&partnerID=MN8TOARS}, DOI={10.1016/j.foodhyd.2005.03.014}, abstractNote={Protein foams are an integral component of many foods such as meringue, nougat and angel food cake. With all these applications, the protein foam must first obtain the desired level of air phase volume (foamability), and then maintain stability when subjected to a variety of processes including mixing, cutting and heating. Therefore, factors determining foamability and stability to mechanical and thermal processing are important to proper food applications of protein foams. We have investigated the effects of protein type, protein modification and co-solutes on overrun, stability and yield stress. The level of overrun generated by different proteins was: whey protein isolate hydrolysates >whey protein isolate=β-lactoglobulin=egg white>α-lactalbumin. The level of yield stress generated by different proteins was: egg white>whey protein isolate hydrolysates≥β-lactoglobulin>whey protein isolate>α-lactalbumin. Factors that decreased surface charge (pH∼pI or high ionic strength) caused a more rapid adsorption of protein at the air–water interface, generally increased dilatational viscoelasticity and increased foam yield stress. The elastic component of the dilatational modulus of the air–water interface was correlated with foam yield stress. The properties of foams did not predict performance in making angel food cakes. A model for foam performance in angel food cakes is proposed.}, number={2-3}, journal={FOOD HYDROCOLLOIDS}, author={Foegeding, EA and Luck, PJ and Davis, JP}, year={2006}, pages={284–292} } @article{pernell_foegeding_luck_davis_2004, title={Properties of whey and egg white protein foams (vol 204, pg 9, 2002)}, volume={238}, ISSN={["1873-4359"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-2442575970&partnerID=MN8TOARS}, DOI={10.1016/j.colsurfa.2004.03.003}, number={1-3}, journal={COLLOIDS AND SURFACES A-PHYSICOCHEMICAL AND ENGINEERING ASPECTS}, author={Pernell, CW and Foegeding, EA and Luck, PJ and Davis, JP}, year={2004}, month={May}, pages={161–161} } @article{luck_bray_foegeding_2002, title={Factors determining yield stress and overrun of whey protein foams}, volume={67}, ISSN={["0022-1147"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0036324489&partnerID=MN8TOARS}, DOI={10.1111/j.1365-2621.2002.tb08704.x}, abstractNote={ABSTRACT: Foams were formed by whipping whey protein solutions (15% w/v protein) containing NaCl, CaCl2, lactose, or glycine. Foam overrun and yield stress were determined. Foams made from whey protein ingredients have greater overrun and yield stress if the concentration of β‐lactoglobulin is high relative to a‐lactalbumin. The presence of 0.4 M CaCl2 in the foaming solution increases overrun and yield stress for β‐lactoglobulin and a‐lactalbumin. The high yield stress of β‐lactoglobulin and a‐lactalbumin foams made from solutions containing CaCl2 suggests that CaCl2 is altering rheological properties of the interfacial protein film and/or contributing to protein aggregation or network formation in the lamellae.}, number={5}, journal={JOURNAL OF FOOD SCIENCE}, author={Luck, PJ and Bray, N and Foegeding, EA}, year={2002}, pages={1677–1681} } @article{pernell_luck_foegeding_daubert_2002, title={Heat-induced changes in angel food cakes containing egg-white protein or whey protein isolate}, volume={67}, ISSN={["0022-1147"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0036812822&partnerID=MN8TOARS}, DOI={10.1111/j.1365-2621.2002.tb08843.x}, abstractNote={ABSTRACT: Angel food cakes made from egg white or whey protein foams were compared. Cakes were evaluated based on final volume, dynamic volume change, and rheological transitions during baking. Cake expansion during baking was a function of protein concentration regardless of protein type. Cakes containing whey proteins had a lower ability to prevent collapse once starch gelatinization started during baking. Heat‐treating whey proteins or adding xanthan gum increases cake volume, but not to the extent of egg‐white proteins. Cakes containing egg‐white proteins became more elastic at 60 to 85 °C than those containing whey proteins, indicating physical differences in the heat‐set protein foam network associated with protein type.}, number={8}, journal={JOURNAL OF FOOD SCIENCE}, author={Pernell, CW and Luck, PJ and Foegeding, EA and Daubert, CR}, year={2002}, month={Oct}, pages={2945–2951} } @article{pernell_foegeding_luck_davis_2002, title={Properties of whey and egg white protein foams}, volume={204}, ISSN={["0927-7757"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0037161616&partnerID=MN8TOARS}, DOI={10.1016/S0927-7757(01)01061-5}, abstractNote={Foams made of varying concentrations (2–20% w/v protein) of egg white and whey protein isolate were compared by measuring rheological and microstructural properties. Egg white proteins formed foams with higher yield stress (τ) at lower protein concentrations and less whipping time than whey protein isolate foams. The model of Princen and Kiss [J. Coll. Interf. Sci. 128 (1989) 176] predicts a relationship among τ, surface tension (σ), phase volume (φ), and bubble size (R32). This was supported by τ increasing with φ, and the relationship between τ and φ1/3 becoming more linear as protein concentration increased. However, egg white foams had yield stress values as much as 100 Pa greater than whey protein foams, despite having similar phase volumes, bubble size, and lower surface tensions. The experimentally determined factors, Y(φ), for egg white and whey protein isolate foams were within the range determined by Princen and Kiss [J. Coll. Interf. Sci. 128 (1989) 176] for concentrated emulsions. Egg white foams were different in that the values for Y(φ) increased at lower phase volumes than for whey protein isolate foams or concentrated emulsions. These results suggest that specific proteins contribute to foam τ by some means in addition to altering surface properties.}, number={1-3}, journal={COLLOIDS AND SURFACES A-PHYSICOCHEMICAL AND ENGINEERING ASPECTS}, author={Pernell, CW and Foegeding, EA and Luck, PJ and Davis, JP}, year={2002}, month={May}, pages={9–21} } @article{luck_lanier_daubert_kwanyuen_2002, title={Viscoelastic behavior of commercially processed soy isolate pastes during heating and cooling}, volume={67}, ISSN={["0022-1147"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0036293043&partnerID=MN8TOARS}, DOI={10.1111/j.1365-2621.2002.tb10293.x}, abstractNote={ABSTRACT}, number={4}, journal={JOURNAL OF FOOD SCIENCE}, author={Luck, PJ and Lanier, TC and Daubert, CR and Kwanyuen, P}, year={2002}, month={May}, pages={1379–1382} }