You are on page 1of 14

pubs.acs.

org/JAFC Article

Effect of Processing of Whey Protein Ingredient on Maillard


Reactions and Protein Structural Changes in Powdered Infant
Formula
Pernille Lund, Mie Rostved Bechshøft, Colin A. Ray, and Marianne N. Lund*
Cite This: J. Agric. Food Chem. 2022, 70, 319−332 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
Downloaded via UNIV TECNICA FEDERICO SANTA MARIA on April 4, 2023 at 19:39:46 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The most widely used whey protein ingredient in an infant formula (IF) is the whey protein concentrate (WPC). The
processing steps used in the manufacturing of both a powdered IF and a WPC introduce protein modifications that may decrease the
nutritional quality. A gently processed whey protein ingredient (serum protein concentrate; SPC) was manufactured and used for
the production of a powdered IF. The SPC and the SPC-based IF were compared to the WPC and the powdered WPC-based IF.
Structural protein modifications were evaluated, and Maillard reaction products, covering furosine, α-dicarbonyls, furans, and
advanced glycation end products, were quantified in the IFs and their protein ingredients. IF processing was responsible for higher
levels of protein modifications compared to the levels observed in the SPC and WPC. Furosine levels and aggregation were most
pronounced in the WPC, but the SPC contained a high level of methylglyoxal, revealing that other processing factors should be
considered in addition to thermal processing.
KEYWORDS: AGEs, CML, α-dicarbonyls, HMF, furosine, protein crosslinks, LAL, LC−MS/MS

■ INTRODUCTION
Various processing steps are used in the manufacturing of a
The majority of AGEs formed in dairy foods are protein
bound and may therefore affect the protein structure.9 In
powdered infant formula (IF), and high-temperature treat- addition, thermal processing is known to induce protein
ments are usually included to ensure microbiological safety, denaturation and aggregation of especially whey proteins.
which is crucial for infants. However, this can also induce Depending on the severity, denaturation and aggregation can
protein modifications, such as those formed by the Maillard affect the digestibility.10 Lanthionine (LAN) and lysinoalanine
reaction, also referred to as protein glycation or non-enzymatic (LAL) are non-reducible protein crosslinks that are associated
browning. The Maillard reaction is initiated by the with processed foods and should be considered in the IF since
condensation of a reducing sugar, such as lactose, with an their formation may result in decreased protein digestibility or
amino group on a protein, peptide, or amino acid to form a impaired nutritional quality.11 Caseins are rich in phosphoser-
Schiff base. Further re-arrangements and condensation ine residues, and whey proteins have high levels of cysteine.
reactions lead to a wide range of Maillard reaction products, Both residues can undergo β-elimination to form dehydroala-
which affect the product quality and possibly also human nine during thermal processing. Dehydroalanine can react with
health. 5-Hydroxymethylfurfural (HMF) and α-dicarbonyls are lysine or cysteine to form LAL or LAN, respectively. Overall,
both reactive Maillard reaction intermediates, while α- thermal processing can result in the loss of essential amino
dicarbonyls in particular may pose negative health con- acids, leading to a decrease in nutritional quality.
sequences.1−4 Numerous studies have been conducted in To mimic the composition of human breast milk, the IF has
order to investigate the possible health implications of a whey protein-to-casein ratio of approximately 60:40. The
advanced glycation end products (AGEs), but it is still not majority of the protein fraction in the IF thus consists of whey
understood to which degree dietary AGEs impact human proteins. The most common source of whey protein in the IF
health.5,6 Also, AGEs may not be fully absorbed by the is the whey protein concentrate (WPC), originating from
intestine, where unabsorbed Nε-carboxymethyl-lysine (CML) cheese/sweet whey. The WPC undergoes various processing
and pyrraline have been shown to be a source of nutrition for steps such as thermal treatments which may be unfavorable
the human gut microbiota.7 Due to its abundance, CML is one from a nutritional point of view.12 Indeed, reduced thermal
of the most studied AGEs, is often used as a marker for AGEs,
and can be found at 100-fold higher levels in the IF compared Received: September 8, 2021
to human breastmilk.8 However, several other AGEs have been Revised: November 24, 2021
quantified in dairy products such as IF, where some have been Accepted: December 15, 2021
shown to differ in digestion and absorption.5 Therefore, Published: December 30, 2021
quantification of AGEs has to be broader than the analysis of
CML exclusively.

© 2021 American Chemical Society https://doi.org/10.1021/acs.jafc.1c05612


319 J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

Table 1. Sample Composition [% w/w]


form batch dry matter protein lactose fat
whey protein ingredient SPC liquid 1 20.66 3.95 15.29 low, not determined
2 20.84 3.96 15.98
3 20.57 3.90 15.18
WPC powder 1 94.90 77.50 7.80 5.50
2 95.00 76.90 8.00 5.00
3 95.14 76.80 8.00 5.50
casein ingredient MPC liquid 1 14.10 8.10 4.67 0.10
2 14.28 8.08 4.68 0.11
3 14.13 8.06 4.69 0.12
SM (pasteurized SM) liquid 1 9.17 3.49 4.70a 0.10a
2 8.77 3.51 4.70a 0.10a
3 9.24 3.62 4.70a 0.10a
IF SPC-IF (SPC- and MPC-based IF) powder 1 97.74 11.15 51.00 26.61
2 97.89 10.29 48.95 29.29
3 97.85 11.42 49.60 27.95
WPC-IF (WPC- and SM-based IF) powder 1 97.63 11.84 48.75 28.47
2 97.76 11.69 49.05 28.00
3 97.47 11.63 48.95 27.97
a
The lactose and fat content of SM was from the packaging label.

processing of the WPC has shown to benefit the intestinal standards were purchased from Iris Biotech GmbH (Marktredwirtz,
integrity in pre-term and near-term piglets.13 High levels of Germany), with net weight values given as percentage: N-ε-
protein modifications have been observed in the whey protein Carboxyethyl-lysine (CEL, 89.6%), CML (95.5%), glyoxal-lysine
ingredient when compared to the final formulation,14 but a dimer TFA salt (GOLD, 94.1%), methylglyoxal-hydroimidazolone 3
TFA salt (MG-H3, 50.5%), glyoxal-hydroimidazolone 1 (GO-H1,
large variation in the whey protein ingredient used for the IF
80.7%), MG-H1-d3 acetate salt (90.5%), furosine HCl salt (72.7%),
has also been observed.15 LAL HCl salt (mixture of two diastereoisomers, 62.7%), CEL-d4
The present study aims to evaluate how the processing (94.4%), CML-d4 (94.4%), MOLD-15N2 acetic acid salt (88.9%),
history of ingredients would impact the extent and type of furosine-d4 HCl salt (52.8%), glyoxal-lysine-amide HCL salt (GOLA,
protein modification in the final IF. A gently processed whey 56.9%), glycolic acid-lysine-amide (GALA, 96.2%), argpyrimidine
protein ingredient (serum protein concentrate; SPC) was TFA salt (51.3%), and pyrraline (98%). Net weight values were not
manufactured and compared with the WPC, and both provided by the company for the following standards and internal
ingredients were used in the manufacturing of generic standards (% values given in brackets are chromatographic purities of
powdered IFs. Furthermore, casein ingredients from both the AGE compound): methylglyoxal-lysine dimer acetate salt
manufacturing lines were included to account for possible (MOLD, ≥95%), pentosidine TFA salt (≥97%), GO-H1-13C2
differences in protein modifications. In total, 27 Maillard (≥95%), and GOLD-15N2 acetic acid salt (≥92%); consequently,
reaction products were quantified using liquid chromatography MOLD and pentosidine could only be determined qualitatively, but
GO-H1-13C2 and GOLD-15N2 could still be used as internal standards.
with UV or MS/MS detection to cover early, intermediate, and
Deionized water was obtained from a Milli-Q water system (Millipore
advanced stages of the Maillard reaction. Protein cross-linking Corporation, Bedford, MA, USA) and used for all solutions.
and aggregation were characterized by sodium dodecyl sulfate Production of Samples. All IFs and ingredients, with the
polyacrylamide gel electrophoresis (SDS-PAGE) and size exception of skimmed milk (SM), were produced by Arla Foods
exclusion chromatography (SEC). In addition, LAL and Ingredients Group P/S (Viby, Denmark) in three independent
LAN were quantified. batches. Powdered WPC80 (WPC) was produced from sweet whey

■ MATERIALS AND METHODS


Chemicals. HMF (99%), 5-methyl-2-furaldehyde (MF, ≥98.5%),
derived from cheese processing. The whey was pasteurized (72 °C for
15 s), followed by concentration by ultrafiltration with diafiltration to
reduce the lactose and mineral contents in order to obtain ∼80%
protein on dry matter basis. The retentate was concentrated by
2-furylmethylketone (FMC, ≥98.5%), glyoxal (40% (w/w) in H2O),
2-methylglyoxal (40% (w/w) in H2O), glucosone (≥97%), L-lysine reverse-osmosis and spray-dried using standard drying conditions.
(≥98%), L-arginine (≥98%), d4-lysine (≥98%), trifluoroacetic acid The liquid milk protein concentrate (MPC) and the SPC were
(TFA), DL-LAN (≥98%), bovine serum albumin (≥96%), lactoferrin produced by milk fractionation. Pasteurized SM (72 °C for 15 s) was
(∼90%), α-lactalbumin (α-LA, type III, calcium depleted, ≥85%), β- concentrated by ultrafiltration to reduce the contents of lactose and
lactoglobulin (β-LG, ≥90%), αs-casein (≥70%), β-casein (≥98%), minerals. The retentate stream (MPC) was further subjected to
and κ-casein (≥70%) were purchased from Sigma-Aldrich Co. (St. microfiltration to separate caseins from serum proteins.16 The
Louis, MO, USA). Caseinoglycomacropeptide (cGMP, ≥79%) was permeate from the microfiltration (SPC), which contained serum
obtained from Arla Foods Ingredients P/S (Viby, Denmark). 3,4- proteins, was subsequently concentrated to remove water and
Dideoxyglucosone-3-ene (3,4-DGE, ≥98%) and 2-(2′,3′,4′- minerals to obtain ∼22% protein on dry matter basis. The resulting
trihydroxybutyl)quinoxaline (3-DG quinoxaline, ≥97%) were pur- SPC stream was then demineralized and subjected to a heat treatment
chased from Carbosynth Ltd (Berkshire, UK). NuPAGE 3-(N- at 65 °C for 15 s. Three packages of organic unhomogenized
morpholino)propanesulfonic acid (MOPS) SDS running buffer 20× pasteurized SM from different manufacturing dates (Arla Foods amba,
and NuPAGE lithium dodecyl sulfate (LDS) sample buffer 4× were Slagelse, Denmark) were purchased at a local supermarket.
purchased from Thermo Fisher Scientific (Carlsbad, CA, USA). Two powdered IFs were produced at Arla Foods Ingredients
Diacetyl (≥97%) and furfural (≥98%) were purchased from Merck Group P/S (Viby, Denmark), where lactose was mixed with protein
KGaA (Darmstadt, Germany). The following standards and internal ingredients to reach the desired lactose content and a whey protein-

320 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

to-casein ratio of 60:40. A vegetable oil blend, minerals, vitamins, with oxalic acid) since the presence of oxalic acid caused the intercept
galactooligosaccharides (GOS), and fructooligosaccharides (FOS) of the calibration curve to increase slightly.
were added to the mix before processing into final IF powders using a Quantification of α-Dicarbonyls. α-Dicarbonyls were converted
standard IF process comprising sequential direct steam injection to their respective quinoxalines by derivatization with o-phenylenedi-
pasteurization, evaporation, and spray-drying. Different whey protein amine (OPD)23 and analyzed by RP-HPLC−MS/MS. The samples
and casein ingredients were used for the two IFs. For WPC-IF, the were dissolved or diluted in Milli-Q water to ∼40 mg protein/mL and
protein ingredients consisted of powdered WPC described above derivatized as described previously24 with the following modifications.
dissolved in pasteurized SM (72 °C for 15 s), and for SPC-IF, the Aliquots of 800 μL sample and a centrifugation speed of 22,000g were
protein ingredients consisted of liquid SPC and liquid MPC described used for protein precipitation. These derivatization conditions (37 °C,
above. Powders were vacuum-sealed, and liquid samples were kept in 1 h, in the dark) have previously been validated for glucosone, 3-DG,
small aliquots at −60 °C to avoid repeated thawing/freezing cycles. methylglyoxal, glyoxal, and diacetyl,25 where similar conditions at 37
The composition of the samples was analyzed using standard methods °C for 2 h have been validated for 3,4-DGE.26 3,4-DGE was
(Table 1).17−20 In order to compare the samples, all results were derivatized at 37 °C for 1 and 2 h with no significant differences
normalized to the protein content of the sample, as ingredients were observed for the quantified level of the quinoxaline of 3,4-DGE. Thus,
added to the liquid IF stream based on their protein content. If not derivatization at 37 °C for 1 h was found to be sufficient for the
further specified, solutions of the resuspended powder were freshly quantification of 3,4-DGE. Five microliters of the filtered (0.22 μm)
prepared prior to analysis by dissolving the powder in Milli-Q water sample was injected on a Dionex Ultimate 3000 LC system coupled
with stirring for 1 h at room temperature before final volume with a Thermo Scientific OrbiTrap Q Exactive mass spectrometer
adjustment. (Thermo Fisher Scientific Inc., Waltham, USA) with an ACQUITY
Color and pH. Samples were diluted or dissolved in Milli-Q water UPLC BEH Phenyl (2.1 × 100 mm, 1.7 μm) column (Waters,
to 35 mg protein/mL, and pH and CIELAB values were measured Eschborn, Germany). Electrospray ionization was operated in the
according to a previous approach.21 The total color (E*) was positive mode (3.5 kV spray voltage, 320 °C capillary temperature,
calculated using eq 1. sheath gas flow rate 30 arbitrary units, auxiliary gas flow rate 10
arbitrary units, sweep gas flow rate 2.13 arbitrary units, and 55% S-
E* = (L*)2 + (a*)2 + (b*)2 (1) lens RF levels). Separation was carried out at 55 °C with a flow rate of
0.350 mL/min by the following gradient: 0−9.7 min 3−25% B, 9.7−
Protein Composition. The protein composition of the samples 11.5 min 25−50% B, 11.5−11.6 min 50−80% B, and 11.6−13.3 min
was determined by reversed-phase (RP) liquid chromatography with 80% B using 0.2% formic acid in Milli-Q water (mobile phase A) and
UV detection. The samples were prepared and analyzed as previously 100% methanol (mobile phase B). The gradient had been optimized
described21 with the following modifications. The injection volume to allow separation and quantification of 11 different α-dicarbonyls,
used was 10 μL. The gradient was modified to 0−4 min 15−31% B, including galactosone, glucosone, 1-deoxyglucosone (1-DG), 3-DG,
4−9 min 31−39% B, 9−10 min 39% B, 10−11 min 39−41% B, 11− 3-deoxygalactosone (3-DGal), 1-deoxypentosulose/1-deoxypentosone
14 min 41% B, 14−18 min 41−44% B, and 18−22 min 44% B. (1-DPs), 3-deoxypentosulose/3-deoxypentosone (3-DPs), glyoxal,
Quantification was achieved by the use of external calibration curves 3,4-DGE, methylglyoxal, and diacetyl (Supporting Information, Figure
of cGMP, κ-casein, αs-casein, lactoferrin, β-casein, bovine serum S3).
albumin, α-LA, and β-LG, ranging from 0.07 to 1.2 mg/mL with Parallel reaction monitoring (PRM) was used for quantification at a
coefficients of determination (r2) ≥ 0.9947. The chromatogram of resolution of 35,000, an automatic gain control target of 1 × 106, a
standards is shown in the Supporting Information, Figure S1. maximum injection time of 128 ms, and with optimized normalized
Individual standard dilutions were used for each calibration curve. collision energies (NCE) to obtain the highest signal of the selected
Each protein content was quantified as a percentage of the total area product ion (see Table 2). Glucosone, 3-DG, glyoxal, 3,4-DGE,
under the curves of the chromatogram, where the unidentified protein
was described as unexplained. Table 2. Mass Spectrometry Parameters for α-Dicarbonyls
Quantification of Free and Total Furans. Four furans (HMF,
furfural, FMC, and MF) were analyzed by RP-HPLC and UV retention time precursor NCE product
analyte [min] ion [V] ion
detection. Samples were prepared in order to quantify free and total
furans according to the previously described methodology22 with galactosone 4.75 251.1026 35 173.0710
some modifications. For quantification of total furan, 1 mL of glucosone 5.05 251.1026 35 173.0710
dissolved powder (0.02−0.07 mg protein/mL) or liquid sample was 1-DG 6.70 235.1077 40 199.0867
mixed by vortexing with 200 μL of freshly prepared 0.6 M oxalic acid 3-DG 7.13 235.1077 40 199.0867
and heated in a water bath (90 °C, 25 min). For free furans, oxalic 3-DGal 7.38 235.1077 40 199.0867
acid was replaced by Milli-Q water, and the heating step was omitted. 1-DPs 8.11 205.0972 35 187.0867
After cooling on ice, 300 μL of cold 75% (v/w) trichloroacetic acid
3-DPs 8.31 205.0972 35 187.0867
(TCA) was added and mixed by vortexing. For the IFs, a higher
glyoxal 8.59 131.0606 120 95.0497
concentration of TCA was needed in order to disrupt the emulsion.
Therefore, 500 μL was mixed with 100 μL of 0.6 M oxalic acid or 3,4-DGE 11.11 217.0971 35 169.0761
Milli-Q water, and 300 μL of 75% (v/w) TCA was added. The methylglyoxal 11.34 145.0761 110 95.0497
samples were centrifuged (22,000g, 10 min, 4 °C), and compact stable diacetyl 12.83 159.0918 110 95.0497
pellets were observed in all samples. The supernatant was filtered
(0.22 μm), and the samples were injected (20 μL) on the same
UHPLC system as described above equipped with a Zorbax SB-C18 methylglyoxal, and diacetyl were commercially available, where
(4.6 × 12.5 mm, 5 μm) guard column and a Zorbax Eclipse XDB-C18 methylglyoxal, glyoxal, and 3-DG were also available as their
column (4.6 × 150 mm, 5 μm) (Agilent Technologies, Santa Clara, quinoxalines. Quinoxaline and 2-methylquinoxaline, which are
USA). Isocratic separation (22 min) was carried out at 25 °C at 1 quinoxalines of glyoxal and methylglyoxal, have been used to quantify
mL/min using 5% ACN as the mobile phase in order to obtain glyoxal and methylglyoxal in various sample types.24,25,27,28 Standards
baseline separation (see the Supporting Information, Figure S2). were derivatized in the same manner as samples, where glyoxal,
Furans were detected at their UV max: 284 nm (HMF), 276 nm methylglyoxal, and 3-DG were compared to quinoxaline, 2-
(furfural), 274 nm (FMC), and 291 nm (MF). External calibration methylquinoxaline, and 3-DG quinoxaline, respectively, to evaluate
curves in the range of 10−1400 ng/mL were used for quantification, the most suited approach for quantification of these α-dicarbonyls.
with coefficients of determination (r2) ≥ 0.9999. Standards were Small signal intensity differences were observed between derivatized
prepared in the same manner as samples (with and without heating glyoxal and quinoxaline and between derivatized methylglyoxal and 2-

321 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

Figure 1. Extracted ion chromatograms of quantified AGEs, LAN, LAL, and furosine from enzymatic (dotted gray) or acid hydrolysate (solid
black).

Table 3. Mass Spectrometry Parameters for Analytes Detected by MS/MS


analyte retention time [min] precursor ion NCE [V] product ion internal standard
standards pyrraline 5.33 255.1343 10 175.1232 MG-H1-d3
argpyrimidine 6.84 255.1452 33 140.0821 MG-H1-d3
GALA 6.98 205.1183 28 142.0865 MG-H1-d3
furosine 7.53 255.1339 33 84.0814 furosine-d4
MG-H1/2/3 7.56 229.1295 33 114.0666 MG-H1-d3
GO-H1/2/3 7.75 215.1139 33 152.0819 GO-H1-13C2
LAN 8.13 209.0591 20 120.0119 CEL-d4
CEL 8.25 219.1339 25 130.0865 CEL-d4
pentosidine 8.48 380.2166 25 379.2088 CML-d4
CML 8.42 205.1182 28 130.0864 CML-d4
GOLA 8.65 333.2133 33 169.0978 CML-d4
LAL 8.68 234.1448 33 130.0866 CML-d4
MOLD 8.77 342.2261 30 341.2183 MOLD-15N2
GOLD 8.85 328.2105 30 327.2026 GOLD-15N2
internal standards furosine-d4 7.53 259.1590 33 88.1064
MG-H1-d3 7.57 232.1484 37 117.0855
GO-H1-13C2 7.75 217.1206 33 154.0888
CEL-d4 8.26 223.1593 33 88.1064
CML-d4 8.41 209.1434 20 134.1117
MOLD-15N2 8.77 344.2202 33 343.2134
GOLD-15N2 8.86 330.2046 33 329.1986

methylquinoxaline. Derivatized glyoxal and methylglyoxal were used GOLA, GOLD, and MOLD), LAN, LAL, and furosine were separated
for quantification since they were derivatized in the same manner as by HILIC and detected by MS/MS. Two different approaches were
glyoxal and methylglyoxal present in the samples. Derivatized 3-DG required for protein hydrolysis. MG-H1/2/3, GO-H1/2/3, CEL,
resulted in a lower signal than 3-DG quinoxaline, which was ascribed pentosidine, GOLD, MOLD, LAN, LAN, and furosine were
to the low purity of 3-DG (≥75%), implying overestimation of 3-DG quantified after acid hydrolysis, while enzymatic hydrolysis was used
in samples if using derivatized 3-DG for the preparation of a for pyrraline, argpyrimidine, GOLA, and GALA since they are not
calibration curve. Consequently, 3-DG quinoxaline was used for stabile in acid. Furthermore, CML was quantified in the enzymatic
quantification. Galactosone, 1-DG, 3-DGal, 1-DPs, and 3-DPs were hydrolysates since acidic hydrolysis may cause artifactual formation of
not commercially available, so glucosone was used to quantify CML and thus an overestimation.
galactosone, and 3-DG quinoxaline was used to quantify 1-DG, 3- Acidic Hydrolysis. Samples were diluted or dissolved in Milli-Q
DGal, 1-DPs, and 3-DPs. Identification of galactosone, 1-DG, 3-DGal, water to 5 mg protein/mL and hydrolyzed with HCl using a Biotage
1-DPs, and 3-DPs was based on their elution order and MS Initiator + microwave synthesizer (Biotage, Uppsala, Sweden) as
spectra.25,29,30 Quantification was achieved based on the external previously described.31
calibration curves, ranging from 1 to 200 and 5−2500 ng/mL for 3- Enzymatic Hydrolysis. The samples were diluted or dissolved in
DG quinoxaline, with coefficients of determination (r2) ≥ 0.9902. The 0.02 M HCl to 3 mg protein/mL, and enzymatic hydrolysis was
samples and standards were all included in the same sequence, and a carried out according to a method described elsewhere.32 Enzymati-
standard was included for every 10th run to monitor the variation in cally hydrolyzed samples were centrifuged (15,000g, 5 min, 4 °C) in
instrument response, where a variation of ≤ 10% was allowed. order to precipitate the undigested protein.
Quantification of AGEs, Furosine, LAN, and LAL by Hydrophilic Interaction Liquid Chromatography-Tandem Mass
Hydrophilic Interaction Liquid Chromatography-Tandem Spectrometry. The method previously described33 was used with the
Mass Spectrometry. Eleven different AGEs (pyrraline, argpyrimi- following modifications. ACN/Milli-Q water (65:35) was used for
dine, GOLA, MG-H1/2/3, GO-H1/2/3, CEL, pentosidine, CML, needle wash to match the solvent composition while still being polar

322 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

Table 4. Mass Spectrometry Parameters for Lysine and Arginine


analyte retention time [min] product ion precursor ion NCE [V] internal standard
standards arginine 3.79 70.0657 175.1190 45 lysine-d4
lysine 4.17 84.0813 147.1128 40 lysine-d4
internal standard lysine-d4 4.16 88.1065 151.1379 33

enough to remove analytes. An injection volume of 20 μL was used. UHPLC system as described above equipped with a Bio SEC-5 guard
This is a high injection volume for HILIC, and peak broadening was column (7.8 × 50 mm, 500 Å, 5 μm) and a Bio SEC-5 column (7.8 ×
observed for early eluting analytes when standards of the highest 300 mm, 500 Å, 5 μm) from Agilent Technologies. Isocratic elution
concentrations were analyzed. However, since no difference in the was carried out at room temperature at a flow rate of 1 mL/min with a
linearity of the calibration curves was observed, 20 μL injection was run time of 20 min with a detection at 214 nm. Based on a protein
chosen since this allowed for better detection of analytes in low standard mix 15−600 kDa (Sigma-Aldrich, Schnelldorf, Germany)
concentrations. The gradient was optimized in order to separate and and individual standards of α-LA and β-LG, chromatograms were
detect all internal standards and analytes within a single run (see divided into three areas describing the protein material larger than
Figure 1): 90% B 0−1 min, 90−75% B 1−4 min, 75−60% B 4−5 min, 670 kDa, around 150 kDa, and below 45 kDa (Figure 2).
60% B 5−6 min, 60−50% B 6−7 min, and 50% B 7−10 min, using 5
mM ammonium formate in Milli-Q water (solvent A) and in 94.6%
ACN (solvent B), both adjusted to pH 2.5 with formic acid. Cleaning
and re-equilibration were carried out at an increased flow in order to
allow for a total run time of 16.5 min.
PRM was used as described previously with optimized NCEs to
obtain the highest signal of the selected product ion (see Table 3).
Quantification was based on internal standard calibration (80 ng/mL
internal standard, except 320 ng/mL of furosine-d4) with calibration
curves ranging from 2 to 150 and 20 to 3000 ng/mL for furosine and
coefficients of determination (r2) ≥ 0.9964, based on the net weight
values of standards. Because isotopically labeled standards for
pyrraline, argpyrimidine, GALA, LAN, pentosidine, GOLA, and
LAL were not available, an internal standard was chosen based on the
retention time to ensure that matrix effects on ionization could be
accounted for. Since the method did not allow for the separation of
GO-H isomers and MG-H isomers, these isomers were quantified as
GO-H1 equivalents as suggested by others.33 MG-H1 is the major
isomer formed in foods but has been found to rearrange into MG-H3 Figure 2. Size exclusion chromatogram of β-LG (green), α-LA (blue),
under acidic conditions.33,34 Consequently, MG-H isomers were and a protein standard mix (gray), including areas described as > 670,
quantified as MG-H3 equivalents. ∼150, and < 45 kDa.
Quantification of Unmodified Lysine and Arginine. Un-
modified lysine and arginine were quantified in acid hydrolysates Statistical Analysis and Determination of Limit of Detection
using the same approach as for AGEs with the following and Quantification. Each sample was produced in three
modifications. Hydrolysates were diluted to 0.05 mg protein/mL in independent batches, and each batch was analyzed in technical
ACN/Milli-Q water (75:25). Five microliters of the sample was duplicates since variation in technical duplicates was minor. For each
injected, and isocratic separation was carried out at 0.400 mL/min sample, quantified values are given as mean values with standard
with 65% B within 7 min using 10 mM ammonium formate in Milli-Q deviation. Significant difference at p < 0.05 was calculated for all
water at pH 4.5 (mobile phase A) and 100% ACN (mobile phase B). pairwise comparisons using the Tukey’s honest significant difference
PRM was used for each analyte, and NCE was optimized to obtain the (HSD) test. Limit of detection (LOD) and limit of quantification
highest signal of the selected product ion (see Table 4). (LOQ) were calculated based on the standard error of y-intercepts
and slope of the full calibration curve.


Quantification was based on internal standard calibration (300 ng/
mL lysine-d4) with calibration curves ranging from 100 to 10,000 ng/
mL for lysine and 100 to 6000 ng/mL for arginine and coefficients of RESULTS AND DISCUSSION
determination (r2) ≥ 0.9962. Protein Composition and Unmodified Lysine and
Sodium Dodecyl Sulfate Polyacrylamide Gel Electropho-
resis. Samples were diluted or dissolved in 5% (w/v) SDS to a Arginine. Protein composition was determined by an HPLC
protein concentration of 12.5 mg/mL, mixed by vortex, and left at method with UV detection at 214 nm (Figure 3). The relative
room temperature on a rocking table for ∼1 h to enhance protein abundance of α-LA and β-LG was higher in SPC compared to
solubilization. Preparation of loading samples, electrophoresis, and that in WPC, which can be explained by the absence of cGMP
staining with Coomassie Brilliant Blue were performed as previously in SPC. With the applied sample preparation, cGMP will not
described apart from minor modifications.21 Electrophoresis was elute as a single peak due to phosphorylation and glycosylation
initially run at 100 V until all the protein material had entered the gel (see Figure S1 in the Supporting Information). Protein
(6 min), followed by a stepwise increase to 150 V for 6 min until 200 material with the same elution time as the cGMP standard
V was reached, and the electrophoresis was then continued for 58 was detected in SPC and SPC-IF, but the elution profile
min. This promoted separation into clear protein bands when differed significantly, and therefore, this protein material
analyzing samples with large protein aggregates.
Size Exclusion Chromatography. Powdered samples were
(∼2%) was categorized as unexplained. β-CN was observed
dissolved in Milli-Q water to 20−60 mg protein/mL. Liquid samples in SPC and has been previously observed to follow the
and dissolved powdered samples were diluted with mobile phase (100 retentate stream during the membrane filtration of SM at 50
mM phosphate buffer, pH 7.0) to 1 mg protein/mL prior to °C.35 Approximately 31% of the protein material in WPC
centrifugation (8000g, 10 min, 4 °C). The supernatant was filtered could not be identified, whereas only ∼14% of the unexplained
(0.22 μm), and the samples were injected (10 μL) on the same protein material was observed in SPC. The unexplained part is
323 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

Figure 3. Protein composition calculated as percentage of total protein, where “unexplained” covers the unidentified protein material.

Figure 4. Quantified unmodified arginine (A) and lysine (B), where different letters indicate significant difference (Tukey’s HSD, p < 0.05).

Table 5. Quantified AGEs [ng/mg Protein], Where n.d. Indicates Not Detecteda
CML CEL pyrraline GO-H1 eqv. MG-H3 eqv.
SPC 29.7 ± 1.5a <LOQ 12.4 ± 3.2a 14.8 ± 0.8a 12.9 ± 0.5a
WPC 22.0 ± 2.9b <LOD 13.7 ± 2.1a n.d. <LOQ
MPC 13.5 ± 1.5c <LOD <LOD <LOQ <LOQ
SM 13.7 ± 1.4c <LOD <LOD <LOQ 4.53 ± 0.22b
SPC-IF 139 ± 12d 9.16 ± 0.54a 28.2 ± 8.4b 17.7 ± 2.1b 19.3 ± 1.4c
WPC-IF 143 ± 15d 8.99 ± 1.93a 36.2 ± 2.9c 17.8 ± 1.2b 25.0 ± 1.7d
LOD/LOQ 1.89/6.21 2.35/7.83 2.44/8.12 1.51/5.02 0.90/3.01
a
Different letters indicate significant differences within a column (Tukey’s HSD, p < 0.05).

expected to cover proteins and peptides, which were not arginine were quantified in all samples (Figure 4). It should be
targeted and quantified with the employed method in addition noted that these results cannot be evaluated as available lysine
to modified and non-native proteins. Higher levels of α-LA and and arginine since nutritionally unavailable derivatives can
β-LG were quantified in SPC-IF compared to those in WPC- revert during acid hydrolysis.36 Furthermore, the absorption of
IF, but no difference was observed in the unexplained protein lysine and arginine in the small intestine should also be
material. No major differences in protein composition were considered if evaluating the availability of unmodified lysine
observed between MPC and SM, supporting that the observed and arginine. Caseins contain more arginine than whey
differences in the protein compositions of IFs are caused by the proteins, which explains the high content of arginine in
differences in the SPC and WPC. MPC and SM (Figure 4A). Furthermore, unmodified arginine
Lysine and arginine are essential and conditionally essential levels were significantly higher in WPC than in SPC, which
amino acids, respectively, and are precursors of several AGEs. were also observed for IFs. The content of unmodified lysine
A decrease in the content of these amino acid residues can was significantly higher in SPC than in WPC, but no significant
result in a decreased nutritional value. Unmodified lysine and difference was observed between IFs (Figure 4B). The
324 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

Table 6. CIELAB Color Values (L*, a*, and b*) Measured at 35 mg Protein/mL with Calculated Total Color (E*)a
L* a* b* E*
SPC 53.8 ± 2.2a −2.97 ± 0.79a 30.8 ± 1.3a 62.1 ± 2.6a
WPC 52.3 ± 1.5b −2.41 ± 0.13b −2.92 ± 0.41b 52.4 ± 1.3b
MPC 79.6 ± 0.3c −4.78 ± 0.26c −3.31 ± 0.58b 79.9 ± 0.3c
SM 79.2 ± 0.1c −5.26 ± 0.10d −2.17 ± 0.03c 79.4 ± 0.3c
SPC-IF 86.9 ± 0.7d 0.338 ± 0.749e 11.8 ± 1.4d 87.7 ± 0.8d
WPC-IF 87.9 ± 0.3e −1.65 ± 0.10f 13.7 ± 0.9e 88.9 ± 0.5e
a
Different letters indicate significant differences within a column (Tukey’s HSD, p < 0.05).

decreased lysine content may be related to the process-induced lysine and glyoxal but were not detected in any of the samples.
modification of WPC but could also be a result of differences MOLD is formed by the cross-linking of two lysine residues by
in composition since α-LA and β-LG are more rich in lysine reaction with methylglyoxal but was also not detected in any of
than cGMP. It was not possible to calculate the theoretical the samples. The higher level of lysine-derived AGEs in SPC
level of lysine and arginine based on the protein composition when compared to that in WPC does not correlate with the
of the different ingredients and IFs due to the large amount of low level of unmodified lysine in WPC, indicating that the
the unexplained protein material in especially WPC and IFs. differences in unmodified lysine are related to differences in
The intermediate stage of the Maillard reaction can be protein composition.
accompanied by a decrease in pH, and this pH decrease may Of the arginine-derived AGEs, only GO-H and MG-H
influence structural changes. Of the ingredients, MPC and SM isomers were quantified, and pentosidine and argpyrimidine
had highest pH values (6.7 ± 0.1 and 6.6 ± 0, respectively), were not detected in any of the samples. GO-H isomers are
while WPC (6.4 ± 0.1) and SPC (6.0 ± 0.1) had significantly formed by the reaction of glyoxal and arginine and were
lower pH values. No significant difference was observed quantified as GO-H1 equivalents as described previously.33 Of
between the pH of SPC-IF (6.6 ± 0.1) and WPC-IF (6.7 ± the protein ingredients, GO-H isomers were only quantified in
0.2). SPC, whereas levels were below LOQ in casein ingredients and
Quantification of Advanced Maillard Reaction Prod- not detected in WPC. As with the other AGEs, levels increased
ucts. Previous studies have found that the majority of AGEs in significantly during IF processing, but the choice of whey
dairy products are protein bound.24,37 CML can be formed by protein ingredient had no impact on GO-H isomer formation
multiple pathways, such as reaction of lysine with glyoxal or and no significant difference was observed between the two
from the oxidation of Amadori products.38 CML was the most IFs. MG-H isomers were formed by the reaction of
abundant AGE of the 11 investigated AGEs and could be methylglyoxal and arginine and were quantified as MG-H3
quantified in all samples (Table 5). The lowest CML levels equivalents.33 As for CML and GO-H isomers, the highest
were detected in casein ingredients, with no difference between level of MG-H isomers in the ingredients was observed in SPC,
MPC and SM. These levels were at comparable levels as while the levels were below LOQ in WPC and MPC but was
detected by others in pasteurized semi-SM.39 Higher levels of quantifiable in SM. IF processing resulted in significant
CML were detected in the whey protein ingredients, which formation of MG-H isomers, and WPC-IF had a significantly
correlate with their higher lysine content. The CML higher level of MG-H isomers than SPC-IF. The conversion of
concentration was significantly higher in SPC when compared MG-H1 to MG-H3 during acid hydrolysis34 may depend on
to that in WPC, which was surprising since increased thermal the sample type and has been shown to be affected by
processing is known to promote CML formation. IF processing hydrolysis conditions,33 which should be considered when
caused CML levels to increase significantly. However, the comparing the MG-H isomer levels with other studies.
choice of whey protein ingredient had no significant impact on Furthermore, the difference in response of MG-H1 and MG-
CML levels in the IFs. Other studies have reported the CML H3 may result in overestimation of MG-H isomers, if MG-H3
levels of 26−140 ng/mg protein in powdered IFs, where acid is the dominant isomer, but quantified as MG-H1. The levels
hydrolysis with reduction has been used as sample of arginine-derived AGEs (GO-H and MG-H isomers) in
preparation.39,40 Comparison of CML results obtained by ingredients did not reflect the levels of unmodified arginine,
different hydrolysis approaches should be done with care, but where SPC had significantly lower arginine concentration than
comparable levels have been found for enzymatic hydrolysis WPC and almost half the levels as for MPC and SM (Figure
and acid hydrolysis with the use of a reduction step.37 4A). Overall, pyrraline, GO-H, and MG-H isomers indicated a
CEL is formed from the reaction between lysine and higher degree of Maillard reaction in WPC when compared to
methylglyoxal and was detected in all samples but could only that in SPC. IF processing was responsible for the formation of
be quantified in IFs. No significant differences were observed five AGEs with levels found in the following order: CML ≫
between the two IFs, and similar values have been detected in pyrraline = MG-H > GO-H > CEL. Formation of AGEs was
liquid infant milk.41 Pyrraline is also a lysine-derived AGE affected differently by IF processing, which emphasizes the
formed from the reaction of lysine and 3-deoxyosone. Pyrraline relevance of quantifying several AGEs, especially when their
was quantified in whey protein ingredients (12−14 ng/mg possible implications on health are not fully understood. The
protein) but was below LOD in casein ingredients. IF CML levels of IFs were ∼4-fold higher than pyrraline, which
processing caused significant formation of pyrraline (30−39 was the second most abundant AGE, making CML the
ng/mg protein), and WPC-IF contained significantly higher quantitatively most relevant AGE.
pyrraline levels than SPC-IF even though no significant Melanoidins are formed in the final stage of the Maillard
difference was observed between WPC and SPC ingredients. reaction and are observed as browning, which was evaluated by
GALA, GOLA, and GOLD originate from cross-linking of measuring CIELAB color values (Table 6). No difference in E*
325 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

was observed for casein ingredients even though differences in raw milk and reported levels of 6−9 ng/mg protein in
were detected in a* and b* values. A significant color pasteurized whey.43 A wide range of HMF levels have been
difference was observed between SPC and WPC, where high reported in powdered IFs and are in general lower than the
b* was measured for SPC, which is expected to relate to the values reported in the present study with values ranging from
high lactose content of SPC, giving it a pronounced yellow 2.6 to 7.7 ng/mg protein for HMF, 13.3−45.2 ng/mg protein
color. IFs had similar lactose levels and are therefore more for total HMF, and 2.2−14.2 ng/mg protein for total
comparable than whey protein ingredients. Significantly higher furfural.43−45 However, a higher level of HMF (∼31 ng/mg
E* was observed for WPC-IF, indicating increased formation protein) has been reported in powdered IFs with added GOS
of melanoidins. and FOS,46 and a higher level of total HMF has been reported
Quantification of Intermediate Maillard Reaction in a commercial liquid IF (253 ng/mg protein), where free
Products. Furans such as HMF, furfural, MF, and FMC are furfural was also found.47
intermediate Maillard reaction products found as free (not α-Dicarbonyls are generated from carbohydrates either by
protein-bound) compounds, where HMF in particular has
the Maillard reaction, the Wolff pathway (oxidation of
been used as an indicator of heat damage in dairy products.
carbohydrates), or sugar condensation. α-Dicarbonyls are
When heating in the presence of oxalic acid, furans are
reactive and can further degrade or react with amino acid
generated from precursors, predominantly Amadori product
but also other Maillard reaction intermediates.42 The value residues to form advanced Maillard reaction products. The α-
obtained after oxalic acid treatment is referred to as “total”, dicarbonyls that are usually detected in dairy foods are formed
covering precursors and free furan. MF and FMC were not by re-arrangement and dehydration of glucose (glucosone, 1-
detected in any samples, which is in agreement with other DG, and 3-DG) and galactose (galactosone and 3-DGal) with
studies and not surprising since these furans are associated 3,4-DGE as an intermediate between 3-DG and 3-DGal. Small
with high heat loads.43,44 Free furfural was not detected in any α-dicarbonyls (glyoxal, methylglyoxal, and diacetyl) can be
of the samples, and precursors of furfural were only detected in formed by fragmentation of 1-DG or isomerization and
SPC and IFs (Table 7). IF processing caused furfural subsequent retro-aldolization of, for example, glucose.48
SPC had ∼6-fold higher methylglyoxal content than WPC,
Table 7. Quantified Furfural and 5-Hydromethylfurfural which was surprising since no harsh heat treatments were
[ng/mg Protein], Both as Total (Determined after Heat applied during the production of SPC. No significant increase
Treatment in Oxalic Acid) and Free Compound (No Oxalic was observed when comparing SPC and SPC-IF, whereas
Acid Treatment)a methylglyoxal increased when comparing WPC and WPC-IF
(Table 8). Methylglyoxal levels somewhat reflected the levels
total furfural total HMF free HMF
of methylglyoxal-derived AGEs (CEL and MG-H isomers).
SPC <LOQ n.d. n.d. However, even though the methylglyoxal content of SPC was
WPC n.d. 1.3 ± 0.5a n.d.
higher than that of WPC, the levels of MG-H isomers were
MPC n.d. n.d. n.d.
significantly higher in WPC-IF than in SPC-IF, indicating
SM n.d. n.d. n.d.
increased potential for MG-H formation in WPC. During
SPC-IF 6.5 ± 0.2a 73 ± 21c 21 ± 12b
processing, SPC and WPC undergo different filtration steps,
WPC-IF 5.0 ± 0.4a 77 ± 15c 26 ± 10b
a where SPC was derived from a permeate stream and WPC
Total furfural was <LOQ (0.75 ng/mg protein) in SPC and n.d.
originated from a retentate stream, where diafiltration had been
indicates not detected. Different letters indicate significant difference
(Tukey’s HSD, p < 0.05). applied. Consequently, molecules smaller than the pore size of
the membrane would be concentrated in SPC but diluted in
precursors to increase, with no significant difference between WPC. The lactose content was reduced during the filtration of
SPC-IF and WPC-IF. HMF was the only furan detected in free WPC, whereas SPC contained significant amounts of lactose.
form and was quantified in IFs with no significant difference The molecular weights of α-dicarbonyls are similar to or
between SPC-IF and WPC-IF. In addition to the formation of smaller than that of lactose. It is therefore suggested that α-
free HMF, IF processing induced a significant increase in HMF dicarbonyls followed the lactose stream during the filtration
precursors. HMF precursors were present in WPC but not in steps of WPC manufacturing. Furthermore, it is proposed that
SPC, which could be explained by the increased processing of the high level of methylglyoxal in SPC was related to its high
WPC. However, difference in the content of HMF precursors lactose content since α-dicarbonyls may be formed via lactose
in whey protein ingredients had no impact on the levels degradation in the absence of proteins. Consequently, it may
detected in IFs. Another study did not detect HMF precursors not be appropriate to normalize quantified α-dicarbonyls to the

Table 8. Quantified α-Dicarbonyls [ng/mg Protein], Where n.d. Indicates Not Detecteda
methylglyoxal glyoxal diacetyl glucosone galactosone 3-DG 3,4-DGE 3-DGal 1-DPs 3-DPs
SPC 6.0 ± 1.2a <LOD <LOD 0.66 ± 0.06a <LOQ 3.1 ± 0.6a <LOD <LOD <LOQ 6.2 ± 2.9a
WPC 1.0 ± 0.2b <LOQ 0.62 ± 0.08a <LOQ 0.75 ± 0.15a <LOQ 10 ± 2ab 2.3 ± 0.7a <LOD <LOD
MPC 1.7 ± 0.3c <LOD <LOD n.d. n.d. <LOD n.d. n.d. <LOD <LOQ
SM 2.6 ± 0.6d <LOD <LOD n.d. n.d. <LOQ <LOD <LOD <LOD 2.0 ± 0.6b
SPC-IF 6.4 ± 0.7a 3.9 ± 0.5a 0.69 ± 0.06b 22 ± 2b 11 ± 1b 165 ± 51b 8.5 ± 1.2a 24 ± 9b 20 ± 2a 4.2 ± 0.3c
WPC-IF 6.1 ± 0.5a 2.5 ± 0.4b < LOQ 16 ± 2c 11 ± 1b 157 ± 18b 13 ± 6b 25 ± 3b 14 ± 2b 3.3 ± 0.2c
LOD/LOQ 0.26/0.85 0.54/1.8 0.16/0.55 0.11/0.36 0.11/0.36 0.43/1.4 0.41/1.4 0.43/1.4 0.43/1.4 0.43/1.4

a
Different letters indicate significant differences within a column (Tukey’s HSD, p < 0.05).

326 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

protein content when comparing protein ingredients with dicarbonyls as compared to the ingredients. Other studies have
different lactose-to-protein ratios. found 3-DG to be the most abundant α-dicarbonyl in UHT
Glyoxal was only quantified in IFs, where SPC-IF had a and infant milk but at lower levels than reported here.24,27,41
significantly higher level than WPC-IF. In terms of glyoxal- To our knowledge, this is the first time α-dicarbonyls have
derived AGEs, no significant difference in CML or GO-H been quantified in the IF.
isomers was observed for IFs, but both AGEs were most Quantification of Early Maillard Reaction Products.
abundant in SPC than in WPC. A significant increase in Amadori products are early Maillard reactions products, which
diacetyl was observed for SPC and SPC-IF, whereas a decrease are converted into furosine during acid hydrolysis. Increased
was observed when comparing WPC and WPC-IF. The furosine levels are observed with increasing thermal processing
diacetyl content of WPC could be attributed to the presence of of dairy products.54 WPC contained 972 ± 246 ng furosine/
lactic acid bacteria during cheese production.49 Diacetyl has mg protein, which was significantly higher than SPC, which
shown to be formed from methylglyoxal at pH 5.50 SPC-IF had contained 107 ± 5 ng/mg protein (Figure 5). The high
pH ∼6.6, but it could be speculated that the diacetyl formation
during processing was related to the high methylglyoxal
content of SPC.
Both WPC and SPC were manufactured from SM with no
addition of other carbohydrates. Therefore, it was unexpected
that SPC and WPC differed in levels of galactosone and
glucosone. However, the presence of lactic acid bacteria during
cheese production has shown to cause accumulation of
galactose in the cheese whey,51 which would explain why
significantly higher galactosone levels were observed in WPC.
Galactosone and glucosone were not detected in casein
ingredients. During IF processing, significant levels of Figure 5. Quantified furosine as an indirect measure of Amadori
glucosone and galactosone were formed, where glucosone products. Different letters indicate significant difference (Tukey’s
was significantly higher in SPC-IF. SPC and WPC differed in HSD, p < 0.05).
3-DG and 3-DGal levels similarly to what was observed for
glucosone and galactosone. IF processing caused significant furosine content of WPC indicated that the low level of
formation of these 3-deoxyosones, with more than a 50-fold unmodified lysine in WPC (Figure 4B) was in fact related to
increase for 3-DG. As mentioned, CML can be formed from increased processing and not solely due to lower levels of α-LA
both glyoxal and 3-deoxyosones. Comparison of low glyoxal and β-LG in WPC as discussed above. Interestingly, different
levels and high levels of 3-DG and 3-DGal suggests that trends were observed in furosine and α-dicarbonyl levels of
formation of CML from 3-deoxyosones was the pronounced WPC and SPC, with furosine levels being highest in WPC and
pathway during IF processing. α-dicarbonyl levels being highest in SPC (apart from 3,4-DGE,
3,4-DGE is an intermediate between 3-DG and 3-DGal and 3-DGal, galactosone, and diacetyl), indicating that different
can lead to the formation of HMF by dehydration.1 In the pathways are favored in the two whey ingredients. No
ingredients, 3,4-DGE was only quantified in WPC, where levels significant difference in furosine levels was observed between
exceeded that of 3-DG and 3-DGal. No increase was observed MPC and SM (Figure 5). Despite a higher level of MG-H
when comparing WPC and WPC-IF, whereas 3,4-DGE isomers and methylglyoxal in SM, quantification of Maillard
formation was significantly increased in SPC-IF as compared reaction products suggests negligible differences in processing
to that in SPC. No differences in total HMF were observed of the casein ingredients and further supports that differences
between IFs (Table 7), which is in agreement with 3-DG and in IFs were related to whey protein ingredients.
3-DGal levels. However, the significantly higher level of 3,4- IF processing resulted in increased furosine levels, and the
DGE in WPC-IF did not result in a significantly higher level of significant difference observed between whey protein ingre-
total HMF. Pyrraline can be formed from 3-DG, and WPC-IF dients was still present when comparing IFs, with furosine
had significantly higher pyrraline levels than SPC-IF (Table 5), levels of 2571 ± 222 and 3238 ± 276 ng/mg protein for SPC-
but no significant differences were observed for 3-DG levels in IF and WPC-IF, respectively. These levels are in agreement
the IFs (Table 8), which indicated that 3,4-DGE levels also with previous studies, where furosine levels of 1700−14,600
affected pyrraline formation. All samples had pH < 7, where ng/mg protein have been found in powdered IFs.39,46,55−57
1,2-enolisation is favored, resulting in higher 3-DG formation Similar levels have also been found in liquid IFs, ranging from
than 1-DG, which is consistent with the observed results. 1910 to 9740 ng/mg protein, where increasing heat load has
3-DPs have been suggested to be formed from α-1,4 been correlated with increased furosine.14,41,47 The increased
glucans,52 such as GOS and FOS, which were present in the levels of pyrraline and MG-H isomers in WPC-IF correlated
IFs. Therefore, it was not surprising that 1-DPs and 3-DPs with the increased furosine levels, when compared to SPC-IF.
were formed during IF processing. The highest level of 3-DPs Overall, the high furosine content of WPC could explain the
was observed in SPC, which was surprising since SPC only significant formation of MG-H and GO-H isomers in WPC-IF
contained lactose and had been subjected to less processing since methylglyoxal and glyoxal were only observed in lower
than WPC. However, 1-DPs and 3-DPs have been quantified levels when compared to SPC and were therefore not clearly
in heated solutions of lysine and glucose.53 In summary, 10 α- linked to MG-H and GO-H isomer formation. This further
dicarbonyls were quantified in the IFs in the following order: supports that other processing parameters apart from the
3-DG ≫ 3-DGal > glucosone > 1-DPs > galactosone = 3,4- applied heat load should be evaluated in order to understand
DGE > methylglyoxal > glyoxal = 3-DPs > diacetyl. IF Maillard-derived protein modifications in the IF, as well as the
processing caused an increase in the majority of these α- relevance of quantifying early-, intermediate-, and advanced
327 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

Figure 6. Representative SDS-PAGE gels of ingredients under non-reduced (A) and reduced conditions (B) and of IFs under non-reduced and
reduced conditions (C). A molecular mass marker (M) and bands belonging to caseins (CN), β-lactoglobulin (β-LG), and α-lactalbumin (α-LA)
are included for reference.

Maillard reaction products. During storage, the Maillard and non-reducing conditions but was more pronounced under
reaction will proceed, with the rate and extent being dependent non-reduced conditions, indicating both disulfide- and non-
on storage conditions.43,47 The higher furosine level in WPC- reducible crosslinks of proteins and/or peptides with β-LG in
IF, compared to SPC-IF, could indicate increased potential for WPC. No obvious differences were observed between MPC
the formation of Maillard reaction products during storage. and SM on the SDS-PAGE gels. In both IFs, significant
Evaluation of Structural Protein Modifications. formation of large disulfide-linked aggregates was observed
Disulfide- and non-reducible crosslinks were evaluated by (comparison between reduced and non-reduced conditions;
SDS-PAGE (Figure 6). A high-molecular-weight (HMW) Figure 6C). Despite clear differences between the whey protein
smear was observed in all batches of WPC under non-reduced ingredients, no clear differences in the amount of aggregated
conditions (Figure 6A), where only a faint HMW smear could material were observed between SPC-IF and WPC-IF. This
be observed under reduced conditions (Figure 6B). This could be linked to differences in protein composition. When
suggests the presence of large disulfide-linked aggregates and compared to WPC-IF, SPC-IF had a higher α-LA and β-LG
some non-reducible crosslinks in WPC, whereas no pro- content due to the absence of cGMP. Consequently, the
nounced HMW smears could be observed in SPC under both protein composition of SPC-IF may make it more prone to
non-reduced and reduced conditions. Furthermore, a smear heat-induced modifications since cGMP does not contain
above β-LG was observed in WPC, which was not observed for cysteine and therefore cannot take part in heat-induced thiol-
SPC. This smear was present on gels run under both reducing disulfide exchange reactions such as α-LA and β-LG. It was
328 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

Figure 7. Quantified LAL (A) and LAN (B). Different letters indicate significant difference (Tukey’s HSD, p < 0.05).

Figure 8. Representative size exclusion chromatograms measured at 214 nm (A), where ★ indicates large soluble aggregates in WPC and arrows
indicates cGMP in WPC and WPC-IF. The chromatograms have been divided into areas described as <45, ∼150, and >670 protein material (B).

evident that large aggregates were formed by disulfide in the production of liquid IF can cause increased formation of
crosslinking during IF processing regardless of whether the LAL when compared to processing steps of powdered IF. Last,
whey protein ingredient was SPC or WPC. alkaline conditions in the liquid IF promote the formation of
LAL and LAN are non-reducible crosslinks derived from LAL.14 Consequently, the neutral pH of SPC-IF and WPC-IF
lysine and cysteine, respectively. LAL was quantified in all (pH ∼ 6.7) may be a contributing cause for limited formation
samples (Figure 7A). LAN was only quantified in whey protein of LAL during processing. The quantified levels of LAN and
ingredients (Figure 7B), which were below LOQ (10 ng/mg LAL by MS/MS showed the same trend as the SDS-PAGE
protein) in IFs and not detected in casein ingredients. No results, where non-reducible crosslinks were most abundant in
significant difference in LAL was observed between the casein WPC than in SPC under reduced conditions (Figure 6B,C).
ingredients. WPC contained significantly higher levels of LAL Hydrophobic interactions also contribute to protein
and LAN when compared to SPC, which can be explained by aggregation but are disrupted in SDS-PAGE analyses. SEC
the increased processing of WPC. This significant difference was therefore used to assess possible differences in aggregation
was also evident when evaluating the LAL content of the two of ingredients and IFs (Figure 8). SEC requires a minimum of
IFs. Apparently, IF processing did not contribute with sample preparation, and it thus allows for analysis of intact
substantial formation of LAL or LAN since IFs had lower protein material, but centrifugation and filtration of samples
levels of LAL and LAN than the whey protein ingredients. could not be omitted. A small pellet was observed in WPC
Previous studies have shown that LAL levels were lower in after centrifugation, and large pellets were observed in IFs, with
powdered IFs than in liquid IFs, but large variations in no difference between SPC-IF and WPC-IF. This means that
commercial products have been observed, indicating great large aggregates present in WPC and IFs were not included in
impact of process parameters within the production of the SEC analysis. This is visually evident as a lower total area in
powdered or liquid IFs.39,58,59 In liquid IFs, up to 20% of WPC and IF samples when compared to SPC (Figure 8B).
LAL was found to originate from the whey protein Removal of large aggregates during SEC sample preparation
ingredient.14 This is contradictory to what was observed in has also been seen when heating WPC at increasing holding
the present study but indicates that processing steps involved times.60 Furthermore, pellet formation correlated with the
329 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

unexplained protein material in Figure 3. No significant same trend as for structural changes, which was in agreement
differences in protein aggregation were observed between with the hypothesis that increased thermal processing of WPC
MPC and SM. A lower total area in MPC and SM was assumed would lead to increased levels of Maillard reaction product in
to relate to large casein micelles being removed during the whey protein ingredient and the final IF. WPC-IF (∼3200
filtration even though no clear pellet could be seen during ng/mg protein) had a significantly higher furosine level than
centrifugation. SPC-IF (∼2600 ng/mg protein). High levels of α-dicarbonyls
When comparing SPC and WPC, large soluble aggregates were observed in SPC than in WPC, which was suggested to be
were only present in WPC (marked by ★ in Figure 8A), while linked to the high lactose content of SPC. This was unexpected
both ingredients contained protein material which eluted when considering the gentle thermal processing of SPC, when
around 150 kDa. Compared to SPC, WPC had a significant compared to WPC, but quantification of α-dicarbonyls
lower area representing protein material smaller than 45 kDa suggests that differences in lactose to protein ratios are
(Figure 8B), which correlated with decreased α-LA and β-LG essential to consider when evaluating protein modifications in
peaks (Figure 8A). Absorbance was also measured at 280 nm, ingredients. Process-induced α-dicarbonyl levels in the IF
which confirmed the presence of cGMP in WPC (arrow in varied, but 3-DG formation was by far the most dominant α-
Figure 8A). Overall, SEC results were consistent with the SDS- dicarbonyl. Interestingly, SPC also contained higher levels of
PAGE results, showing that WPC contained more aggregated AGEs when compared to WPC, while WPC-IF contained the
material. In addition, it has previously been shown that heating highest AGE levels when compared to SPC-IF.
WPI with increasing temperatures caused α-LA and β-LG
peaks to decrease and peaks belonging to large aggregates to
increase.61

*
ASSOCIATED CONTENT
sı Supporting Information
IF processing had a significant impact on aggregation The Supporting Information is available free of charge at
behavior. In addition to the large precipitated aggregates, a https://pubs.acs.org/doi/10.1021/acs.jafc.1c05612.
significant increase in large soluble aggregates (>670 kDa) was
observed in both IFs, whereas the protein material with a Chromatograms of analytical methods (PDF)
molecular weight below 45 kDa decreased (Figure 8B).
However, when comparing the low-molecular-weight (<45
kDa) part of the chromatograms (Figure 8A), visual differences
■ AUTHOR INFORMATION
Corresponding Author
were detected between the two IFs. In addition to the presence Marianne N. Lund − Department of Food Science, Faculty of
of cGMP in WPC-IF (arrow in Figure 8A), lower α-LA and β- Science, University of Copenhagen, 1958 Frederiksberg C,
LG peaks were also observed, corresponding to decreased Denmark; Department of Biomedical Sciences, Faculty of
levels of monomeric α-LA and β-LG. However, it should be Health and Medical Sciences, University of Copenhagen,
kept in mind that WPC-IF had lower levels of α-LA and β-LG 2200 Copenhagen, Denmark; orcid.org/0000-0001-
when compared to SPC-IF (Figure 3). Batch differences were 8708-2210; Phone: +4535333547; Email: mnl@
observed for WPC-IF, where batch 2 had significantly lower food.ku.dk
levels of large soluble aggregates (>670 kDa). Only large
soluble aggregates or monomeric proteins were observed in Authors
IFs, with a lack of 150 kDa intermediates, which were present Pernille Lund − Department of Food Science, Faculty of
in ingredients. However, this trend has also been observed Science, University of Copenhagen, 1958 Frederiksberg C,
previously when heating whey proteins at increased temper- Denmark; orcid.org/0000-0001-8618-0943
atures.62,63 Mie Rostved Bechshøft − Arla Foods Ingredients Group P/S,
Differences in the Formation of Maillard-Related and 8260 Viby J, Denmark
Structural Protein Modifications. Only minor differences Colin A. Ray − Arla Foods Ingredients Group P/S, 8260 Viby
in Maillard-related and structural protein modifications were J, Denmark; orcid.org/0000-0002-1771-1840
observed between the casein ingredients, and the observed Complete contact information is available at:
differences between the IFs could be explained by the choice of https://pubs.acs.org/10.1021/acs.jafc.1c05612
whey protein ingredient. The processing differences between
the whey protein ingredients had some impact on protein Funding
modifications in the final IF, but the differences between This work was supported by the Green Development and
ingredients were in general minor when compared to the Demonstration Programme (GUDP) under the Ministry of
protein modification levels of the final IFs. Our study shows Environment and Food of Denmark [grant number 34009-17-
that if reduced levels of protein modifications are to be 1278] and Arla Foods Ingredients Group P/S.
achieved in the powdered IF, it is necessary to consider
processing parameters in addition to heat load in the Notes
production of whey protein ingredients as well as optimization The authors declare the following competing financial
of the IF process during manufacture. interest(s): M.R.B and C.A.R. are employed at Arla Foods
Ingredients Group P/S.


Different trends were observed when evaluating protein
modifications related to the Maillard reaction versus structural
changes. The increased processing of WPC led to increased ACKNOWLEDGMENTS
levels of disulfide-linked aggregates, LAN, LAL, and large Bente P. Danielsen is thanked for assistance with enzymatic
soluble aggregates. However, the choice of whey protein hydrolysis and sample preparation and Stine L. Gaarde is
ingredient only had a minor impact on structural changes of thanked for performing Kjeldahl analysis of SM. H. Gül
the IF, where the only significant difference was increased Akillioglu, Kasper Engholm-Keller, and Mahesha M. Poojary
levels of LAL in WPC-IF. Furosine formation followed the are thanked for helpful discussions on LC−MS/MS analyses.
330 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry


pubs.acs.org/JAFC Article

ABBREVIATIONS (12) Borad, S. G.; Kumar, A.; Singh, A. K. Effect of Processing on


Nutritive Values of Milk Protein. Crit. Rev. Food Sci. Nutr. 2017, 57,
ACN, acetonitrile; α-LA, α-lactalbumin; β-LG, β-lactoglobulin; 3690−3702.
CEL, N-ε-carboxyethyl-lysine; cGMP, caseinoglycomacropep- (13) Navis, M.; Muncan, V.; Sangild, P. T.; Møller Willumsen, L.;
tide; CML, N-ε-carboxymethyl-lysine; DETAPAC, pentetic Koelink, P. J.; Wildenberg, M. E.; Abrahamse, E.; Thymann, T.; van
acid; 1-DG, 1-deoxyglucosone; 3-DG, 3-deoxyglucosone; 3- Elburg, R. M.; Renes, I. B. Beneficial Effect of Mildly Pasteurized
DGal, 3-deoxygalactosone; 3,4-DGE, 3,4-dideoxyglucosone-3- Whey Protein on Intestinal Integrity and Innate Defense in Preterm
ene; 3-DG quinoxaline, 2-(2,3,4-trihydroxybutyl)-benzo[g]- and Near-Term Piglets. Nutrients 2020, 12, 1125.
quinoxaline; 1-DPs, 1-deoxypentosulose/1-deoxypentosone; (14) Cattaneo, S.; Masotti, F.; Pellegrino, L. Liquid Infant Formulas:
3-DPs, 3-deoxypentosulose/3-deoxypentosone; DTT, dithio- Technological Tools for Limiting Heat Damage. J. Agric. Food Chem.
threitol; FMC, 2-furylmethylketone; FOS, fructooligosacchar- 2009, 57, 10689−10694.
ides; GALA, glycolic acid-lysine-amide; GO-H1, glyoxal- (15) Contreras-Calderón, J.; Guerra-Hernández, E.; García-
hydroimidazolone 1; GOLA, glyoxal-lysine-amide; GOLD, Villanova, B. Indicators of Non-Enzymatic Browning in the Evaluation
glyoxal-lysine dimer; GOS, galactooligosaccharides; HMW, of Heat Damage of Ingredient Proteins Used in Manufactured Infant
Formulas. Eur. Food Res. Technol. 2008, 227, 117−124.
high molecular weight; HMF, 5-hydroxymethylfurfural; IF,
(16) Carter, B. G.; Cheng, N.; Kapoor, R.; Meletharayil, G. H.;
infant formula; LAN, lanthionine; LAL, lysinoalanine; LOD, Drake, M. A. Invited Review: Microfiltration-Derived Casein and
limit of detection; LOQ, limit of quantification; MF, 5-methyl- Whey Proteins from Milk. J. Dairy Sci. 2021, 104, 2465−2479.
2-furaldehyde; MG-H1 and MG-H3, methylglyoxal-hydro- (17) ISO/IDF. Milk, cream and evaporated milkDetermination of
imidazolone 1 and 3; MOLD, methylglyoxal-lysine dimer; total solids content. Retrieved from ISO 6731, IDF 21 2010 https://
MPC, milk protein concentrate; NCE, normalized collision www.iso.org/standard/56815.html.
energy; OPD, o-phenylenediamine; PRM, parallel reaction (18) ISO/IDF. MilkDetermination of nitrogen contentPart 3:
monitoring; SPC, serum protein concentrate; SDS-PAGE, Block-digestion method (Semi-micro rapid routine method).
sodium dodecyl sulfate polyacrylamide gel electrophoresis; Retrieved from ISO 8968-3, IDF 20-3 2004 https://www.iso.org/
SEC, size exclusion chromatography; SM, pasteurized skim standard/35214.html.
milk; TCA, trichloroacetic acid; TFA, trifluoroacetic acid; (19) ISO/IDF. Dried milk, dried ice-mixes and processed cheese
WPC, whey protein concentrate Determination of lactose contentPart 2: Enzymatic method


utilizing the galactose moiety of the lactose. Retrieved from ISO
5765-2, IDF 79-2. 2002 https://www.iso.org/standard/30566.html.
REFERENCES (20) ISO/IDF. Dried milk and dried milk productsDetermination
(1) Hellwig, M.; Gensberger-Reigl, S.; Henle, T.; Pischetsrieder, M. of fat contentGravimetric method (Reference method). Retrieved
Food-Derived 1,2-Dicarbonyl Compounds and Their Role in from ISO 1736, IDF 9 2008 https://www.iso.org/standard/51010.
Diseases. Semin. Cancer Biol. 2018, 49, 1−8. html.
(2) Nunes, L.; Martins, E.; Tuler Perrone, I.;́ Fernandes De (21) Lund, P.; Nielsen, S. B.; Nielsen, C. F.; Ray, C. A.; Lund, M. N.
Carvalho, A. The Maillard Reaction in Powdered Infant Formula. J. Impact of UHT Treatment and Storage on Liquid Infant Formula:
Food Nutr. Res. 2019, 7, 33−40. Complex Structural Changes Uncovered by Centrifugal Field-Flow
(3) Rabbani, N.; Thornalley, P. J. Dicarbonyl Stress in Cell and Fractionation with Multi-Angle Light Scattering. Food Chem. 2021,
Tissue Dysfunction Contributing to Ageing and Disease. Biochem. 348, 129145.
Biophys. Res. Commun. 2015, 458, 221−226. (22) Albalá-Hurtado, S.; Veciana-Nogués, M. T.; Izquierdo-Pulido,
(4) ALjahdali, N.; Carbonero, F. Impact of Maillard Reaction M.; Vidal-Carou, M. C. Determination of Free and Total Furfural
Products on Nutrition and Health: Current Knowledge and Need to Compounds in Infant Milk Formulas by High-Performance Liquid
Understand Their Fate in the Human Digestive System. Crit. Rev. Chromatography. J. Agric. Food Chem. 1997, 45, 2128−2133.
Food Sci. Nutr. 2019, 59, 474−487. (23) Moree-Testa, P.; Saint-Jalm, Y. Determination of α-Dicarbonyl
(5) Zhang, Q.; Wang, Y.; Fu, L. Dietary Advanced Glycation End- Compounds in Cigarette Smoke. J. Chromatogr. 1981, 217, 197−208.
Products: Perspectives Linking Food Processing with Health (24) Zhang, W.; Poojary, M. M.; Rauh, V.; Ray, C. A.; Olsen, K.;
Implications. Compr. Rev. Food Sci. Food Saf. 2020, 19, 2559−2587.
Lund, M. N. Quantitation of α-Dicarbonyls and Advanced Glycation
(6) Nowotny, K.; Schröter, D.; Schreiner, M.; Grune, T. Dietary
Endproducts in Conventional and Lactose-Hydrolyzed Ultrahigh
Advanced Glycation End Products and Their Relevance for Human
Temperature Milk during 1 Year of Storage. J. Agric. Food Chem.
Health. Ageing Res. Rev. 2018, 47, 55−66.
(7) Hellwig, M.; Bunzel, D.; Huch, M.; Franz, C. M. A. P.; Kulling, 2019, 67, 12863−12874.
S. E.; Henle, T. Stability of Individual Maillard Reaction Products in (25) Zhang, W.; Poojary, M. M.; Olsen, K.; Ray, C. A.; Lund, M. N.
the Presence of the Human Colonic Microbiota. J. Agric. Food Chem. Formation of α-Dicarbonyls from Dairy Related Carbohydrates with
2015, 63, 6723−6730. and without Nα-Acetyl-L-Lysine during Incubation at 40 and 50 °C. J.
(8) Š ebeková, K.; Klenovics, K. S.; Š ebeková, K. B. Advanced Agric. Food Chem. 2019, 67, 6350−6358.
Glycation End Products in Infant Formulas. In Handbook of Dietary (26) Frischmann, M.; Spitzer, J.; Fünfrocken, M.; Mittelmaier, S.;
and Nutritional Aspects of Bottle Feeding; Preedy, V. R., Watson, R. R., Deckert, M.; Fichert, T.; Pischetsrieder, M. Development and
Zibadi, S., Eds.; Wageningen Academic Publishers: Wageningen, Validation of an HPLC Method to Quantify 3,4-Dideoxyglucosone-
2014; pp 421−440. 3-Ene in Peritoneal Dialysis Fluids. Biomed. Chromatogr. 2009, 23,
(9) Hellwig, M.; Humpf, H.-U.; Hengstler, J.; Mally, A.; Vieths, S.; 843−851.
Henle, T. Quality Criteria for Studies on Dietary Glycation (27) Hellwig, M.; Degen, J.; Henle, T. 3-Deoxygalactosone, a “New”
Compounds and Human Health. J. Agric. Food Chem. 2019, 67, 1,2-Dicarbonyl Compound in Milk Products. J. Agric. Food Chem.
11307−11311. 2010, 58, 10752−10760.
(10) van Lieshout, G. A. A.; Lambers, T. T.; Bragt, M. C. E.; (28) Kocadağlı, T.; Gökmen, V. Effect of Sodium Chloride on α-
Hettinga, K. A. How Processing May Affect Milk Protein Digestion Dicarbonyl Compound and 5-Hydroxymethyl-2-Furfural Formations
and Overall Physiological Outcomes: A Systematic Review. Crit. Rev. from Glucose under Caramelization Conditions: A Multiresponse
Food Sci. Nutr. 2019, 60, 2422−2445. Kinetic Modeling Approach. J. Agric. Food Chem. 2016, 64, 6333−
(11) Friedman, M. Chemistry, Biochemistry, Nutrition, and 6342.
Microbiology of Lysinoalanine, Lanthionine, and Histidinoalanine in (29) Gensberger, S.; Mittelmaier, S.; Glomb, M. A.; Pischetsrieder,
Food and Other Proteins. J. Agric. Food Chem. 1999, 47, 1295−1319. M. Identification and Quantification of Six Major α-Dicarbonyl

331 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332
Journal of Agricultural and Food Chemistry pubs.acs.org/JAFC Article

Process Contaminants in High-Fructose Corn Syrup. Anal. Bioanal. (48) Hollnagel, A.; Kroh, L. W. Formation of α-Dicarbonyl
Chem. 2012, 403, 2923. Fragments from Mono- and Disaccharides under Caramelization
(30) Henning, C.; Liehr, K.; Girndt, M.; Ulrich, C.; Glomb, M. A. and Maillard Reaction Conditions. Eur. Food Res. Technol. 1998, 207,
Extending the Spectrum of α-Dicarbonyl Compounds in Vivo. J. Biol. 50−54.
Chem. 2014, 289, 28676−28688. (49) Fox, P. F.; Uniacke-Lowe, T.; McSweeney, P. L. H.; O’Mahony,
(31) Poojary, M. M.; Zhang, W.; Greco, I.; De Gobba, C.; Olsen, K.; J. A. Dairy Chemistry and Biochemistry, 2nd ed.; Springer International
Lund, M. N. Liquid Chromatography Quadrupole-Orbitrap Mass Publishing: Cham, 2015.
Spectrometry for the Simultaneous Analysis of Advanced Glycation (50) Pfeifer, Y. V.; Haase, P. T.; Kroh, L. W. Reactivity of Thermally
End Products and Protein-Derived Cross-Links in Food and Treated α-Dicarbonyl Compounds. J. Agric. Food Chem. 2013, 61,
Biological Matrices. J. Chromatogr. A 2020, 1615, 460767. 3090−3096.
(32) Poojary, M. M.; Tiwari, B. K.; Lund, M. N. Selective and (51) Rao, R. D.; Wendorff, W. L.; Smith, K. Changes in Galactose
Sensitive UHPLC-ESI-Orbitrap MS Method to Quantify Protein and Lactic Acid Content of Sweet Whey during Storage. J. Food Prot.
Oxidation Markers. Talanta 2021, 234, 122700. 2004, 67, 403−406.
(33) Akıllıoğlu, H. G.; Lund, M. N. Quantification of Advanced (52) Hollnagel, A.; Kroh, L. W. 3-Deoxypentosulose: An α-
Glycation End Products and Amino Acid Cross-Links in Foods by Dicarbonyl Compound Predominating in Nonenzymatic Browning
High-Resolution Mass Spectrometry. Food Chem. 2022, 366, 130601. of Oligosaccharides in Aqueous Solution. J. Agric. Food Chem. 2002,
(34) Antonova, K.; Vikhnina, M.; Soboleva, A.; Mehmood, T.; 50, 1659−1664.
Heymich, M.-L.; Leonova, T.; Bankin, M.; Lukasheva, E.; Gensberger- (53) Gobert, J.; Glomb, M. A. Degradation of Glucose:
Reigl, S.; Medvedev, S.; Smolikova, G.; Pischetsrieder, M.; Frolov, A. Reinvestigation of Reactive α-Dicarbonyl Compounds. J. Agric. Food
Analysis of Chemically Labile Glycation Adducts in Seed Proteins: Chem. 2009, 57, 8591−8597.
Case Study of Methylglyoxal-Derived Hydroimidazolone 1 (MG-H1). (54) Mehta, B. M.; Deeth, H. C. Blocked Lysine in Dairy Products:
Int. J. Mol. Sci. 2019, 20, 3659. Formation, Occurrence, Analysis, and Nutritional Implications.
(35) McCarthy, N. A.; Wijayanti, H. B.; Crowley, S. V.; O’Mahony, Compr. Rev. Food Sci. Food Saf. 2016, 15, 206−218.
J. A.; Fenelon, M. A. Pilot-Scale Ceramic Membrane Filtration of (55) Zenker, H. E.; Van Lieshout, G. A. A.; Van Gool, M. P.; Bragt,
Skim Milk for the Production of a Protein Base Ingredient for Use in M. C. E.; Hettinga, K. A. Lysine Blockage of Milk Proteins in Infant
Infant Milk Formula. Int. Dairy J. 2017, 73, 57−62. Formula Impairs Overall Protein Digestibility and Peptide Release.
(36) Rutherfurd, S. M.; Moughan, P. J. Development of a Novel Food Funct. 2020, 11, 358−369.
Bioassay for Determining the Available Lysine Contents of Foods and (56) Chen, Z.; Kondrashina, A.; Greco, I.; Gamon, L. F.; Lund, M.
Feedstuffs. Nutr. Res. Rev. 2007, 20, 3−16. N.; Giblin, L.; Davies, M. J. Effects of Protein-Derived Amino Acid
(37) Hegele, J.; Buetler, T.; Delatour, T. Comparative LC-MS/MS Modification Products Present in Infant Formula on Metabolic
Profiling of Free and Protein-Bound Early and Advanced Glycation- Function, Oxidative Stress, and Intestinal Permeability in Cell
Induced Lysine Modifications in Dairy Products. Anal. Chim. Acta Models. J. Agric. Food Chem. 2019, 67, 5634−5646.
2008, 617, 85−96. (57) Troise, A. D.; Fiore, A.; Wiltafsky, M.; Fogliano, V.
(38) Thorpe, S. R.; Baynes, J. W. CML: A Brief History. Int. Congr. Quantification of Nε-(2-Furoylmethyl)-l-Lysine (Furosine), Nε-
Ser. 2002, 1245, 91−99. (Carboxymethyl)-l-Lysine (CML), Nε-(Carboxyethyl)-l-Lysine
(39) Fenaille, F.; Parisod, V.; Visani, P.; Populaire, S.; Tabet, J.-C.; (CEL) and Total Lysine through Stable Isotope Dilution Assay and
Guy, P. A. Modifications of Milk Constituents during Processing: A Tandem Mass Spectrometry. Food Chem. 2015, 188, 357−364.
Preliminary Benchmarking Study. Int. Dairy J. 2006, 16, 728−739. (58) Pompei, C.; Ross, M.; Mar, F. Protein Quality in Commercial
(40) Delatour, T.; Hegele, J.; Parisod, V.; Richoz, J.; Maurer, S.; Milk-Based Infant Formulas. J. Food Qual. 1987, 10, 375−391.
Steven, M.; Buetler, T. Analysis of Advanced Glycation Endproducts (59) D’Agostina, A.; Boschin, G.; Rinaldi, A.; Arnoldi, A. Updating
in Dairy Products by Isotope Dilution Liquid Chromatography- on the Lysinoalanine Content of Commercial Infant Formulae and
Electrospray Tandem Mass Spectrometry. The Particular Case of Beicost Products. Food Chem. 2003, 80, 483−488.
Carboxymethyllysine. J. Chromatogr. A 2009, 1216, 2371−2381. (60) Havea, P.; Singh, H.; Creamer, L. K. Heat-Induced Aggregation
(41) Aktağ, I. G.; Hamzalıoğlu, A.; Gökmen, V. Lactose Hydrolysis of Whey Proteins: Comparison of Cheese WPC with Acid WPC and
and Protein Fortification Pose an Increased Risk for the Formation of Relevance of Mineral Composition. J. Agric. Food Chem. 2002, 50,
Maillard Reaction Products in UHT Treated Milk Products. J. Food 4674−4681.
Compos. Anal. 2019, 84, 103308. (61) Kazmierski, M.; Corredig, M. Characterization of Soluble
(42) Ritota, M.; Di Costanzo, M. G.; Mattera, M.; Manzi, P. New Aggregates from Whey Protein Isolate. Food Hydrocolloids 2003, 17,
Trends for the Evaluation of Heat Treatments of Milk. J. Anal. 685−692.
Methods Chem. 2017, 2017, 1864832. (62) Dalgleish, D. G.; Senaratne, V.; Francois, S. Interactions
(43) Ferrer, E.; Alegría, A.; Farré, R.; Abellán, P.; Romero, F. Effects between α-Lactalbumin and β-Lactoglobulin in the Early Stages of
of Thermal Processing and Storage on Available Lysine and Furfural Heat Denaturation. J. Agric. Food Chem. 1997, 45, 3459−3464.
(63) Schokker, E. P.; Singh, H.; Pinder, D. N.; Norris, G. E.;
Compounds Contents of Infant Formulas. J. Agric. Food Chem. 2000,
Creamer, L. K. Characterization of Intermediates Formed during
48, 1817−1822.
Heat-Induced Aggregation of β-Lactoglobulin AB at Neutral PH. Int.
(44) Chávez-Servín, J. L.; Castellote, A. I.; López-Sabater, M. C.
Dairy J. 1999, 9, 791−800.
Evolution of Potential and Free Furfural Compounds in Milk-Based
Infant Formula during Storage. Food Res. Int. 2006, 39, 536−543.
(45) Ferrer, E.; Alegría, A.; Farré, R.; Abellán, P.; Romero, F. High-
Performance Liquid Chromatographic Determination of Furfural
Compounds in Infant Formulas: Changes during Heat Treatment and
Storage. J. Chromatogr. A 2002, 947, 85−95.
(46) Sabater, C.; Montilla, A.; Ovejero, A.; Prodanov, M.; Olano, A.;
Corzo, N. Furosine and HMF Determination in Prebiotic-
Supplemented Infant Formula from Spanish Market. J. Food Compos.
Anal. 2018, 66, 65−73.
(47) Guerra-Hernandez, E.; Leon Gomez, C.; Garcia-Villanova, B.;
Corzo Sanchez, N.; Romera Gomez, J. M. Effect of Storage on Non-
Enzymatic Browning of Liquid Infant Milk Formulae. J. Sci. Food
Agric. 2002, 82, 587−592.

332 https://doi.org/10.1021/acs.jafc.1c05612
J. Agric. Food Chem. 2022, 70, 319−332

You might also like