plasmin determination based on enzymatic digestion of a

6
Colloids and Surfaces A: Physicochemical and Engineering Aspects 609 (2021) 125786 Available online 22 October 2020 0927-7757/© 2020 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). Plasmin determination based on enzymatic digestion of a β-casein layer at the air/water interface Lor´ and Rom´ anszki a, *, Tibor Hianik b , Zs´ ofia Keresztes a a Functional Interfaces Research Group, Institute of Materials and Environmental Chemistry, Research Centre for Natural Sciences address: Magyar tud´ osok k¨ orútja 2., 1117, Budapest, Hungary b Department of Nuclear Physics and Biophysics, Faculty of Mathematics, Physics and Informatics, Comenius University in Bratislava address: Mlynsk´ a dolina F1, 84248, Bratislava, Slovakia HIGHLIGHTS G R A P H I C A L ABSTRACT Tensiometry-based label-free determi- nation of proteolytic enzyme activity was developed. Change in surface pressure of interfacial protein layers upon reaction was determined. Assay linearity of plasmin enzyme against β-casein substrate of 0.7 nM was achieved. Sensitivity of the method can be enhanced by increasing the concentra- tion of β-casein. A R T I C L E INFO Keywords: Langmuir film Gibbs monolayer β-casein Proteolytic enzyme Tensiometry Surface pressure ABSTRACT The possibility of enzyme activity determination based on natural substrate digestion at the air/water interface has been investigated through the interaction of a β-casein layer with the protease plasmin. The protein layers were spread from aqueous solutions of β-casein in a 0.10.4 g/L concentration range to the air/water interface. 200 μL plasmin was introduced into the water subphase (15 mL) in various initial concentrations (251000 nM), corresponding to 0.3313.3 nM final concentration. The decrease of the surface pressure due to the enzymatic degradation of the β-casein monolayers was followed in time, and calibration curves were obtained by plotting either the total surface pressure change, or the maximum rate of the surface pressure change against the plasmin concentration. The method is suitable for label-free determination of plasmin, with assay linearity down to 0.7 nM as the final concentration of plasmin in the system, corresponding to 50 nM for the initial, undiluted plasmin sample in the given experimental setup. * Corresponding author. E-mail addresses: [email protected] (L. Rom´ anszki), [email protected] (T. Hianik), keresztes.zsofi[email protected] (Z. Keresztes). Contents lists available at ScienceDirect Colloids and Surfaces A: Physicochemical and Engineering Aspects journal homepage: www.elsevier.com/locate/colsurfa https://doi.org/10.1016/j.colsurfa.2020.125786 Received 4 September 2020; Received in revised form 8 October 2020; Accepted 17 October 2020

Upload: others

Post on 18-Apr-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Plasmin determination based on enzymatic digestion of a

Colloids and Surfaces A: Physicochemical and Engineering Aspects 609 (2021) 125786

Available online 22 October 20200927-7757/© 2020 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).

Plasmin determination based on enzymatic digestion of a β-casein layer at the air/water interface

Lorand Romanszki a,*, Tibor Hianik b, Zsofia Keresztes a

a Functional Interfaces Research Group, Institute of Materials and Environmental Chemistry, Research Centre for Natural Sciences address: Magyar tudosok korútja 2., 1117, Budapest, Hungary b Department of Nuclear Physics and Biophysics, Faculty of Mathematics, Physics and Informatics, Comenius University in Bratislava address: Mlynska dolina F1, 84248, Bratislava, Slovakia

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Tensiometry-based label-free determi-nation of proteolytic enzyme activity was developed.

• Change in surface pressure of interfacial protein layers upon reaction was determined.

• Assay linearity of plasmin enzyme against β-casein substrate of 0.7 nM was achieved.

• Sensitivity of the method can be enhanced by increasing the concentra-tion of β-casein.

A R T I C L E I N F O

Keywords: Langmuir film Gibbs monolayer β-casein Proteolytic enzyme Tensiometry Surface pressure

A B S T R A C T

The possibility of enzyme activity determination based on natural substrate digestion at the air/water interface has been investigated through the interaction of a β-casein layer with the protease plasmin. The protein layers were spread from aqueous solutions of β-casein in a 0.1–0.4 g/L concentration range to the air/water interface. 200 μL plasmin was introduced into the water subphase (15 mL) in various initial concentrations (25–1000 nM), corresponding to 0.33–13.3 nM final concentration. The decrease of the surface pressure due to the enzymatic degradation of the β-casein monolayers was followed in time, and calibration curves were obtained by plotting either the total surface pressure change, or the maximum rate of the surface pressure change against the plasmin concentration. The method is suitable for label-free determination of plasmin, with assay linearity down to 0.7 nM as the final concentration of plasmin in the system, corresponding to 50 nM for the initial, undiluted plasmin sample in the given experimental setup.

* Corresponding author. E-mail addresses: [email protected] (L. Romanszki), [email protected] (T. Hianik), [email protected] (Z. Keresztes).

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and Engineering Aspects

journal homepage: www.elsevier.com/locate/colsurfa

https://doi.org/10.1016/j.colsurfa.2020.125786 Received 4 September 2020; Received in revised form 8 October 2020; Accepted 17 October 2020

Page 2: Plasmin determination based on enzymatic digestion of a

Colloids and Surfaces A: Physicochemical and Engineering Aspects 609 (2021) 125786

2

1. Introduction

The need for detecting enzymes and determining enzyme activity is repeatedly encountered in biotechnology, food industry, health care etc. To fulfil these emerging needs, highly sensitive, highly specific, fast, simple, inexpensive, and environmentally harmless measurement plat-forms are desirable.

Plasmin (PL) is a serine protease of molecular mass of 83 kDa syn-thesized in the liver in its inactive, zymogen form plasminogen, and plays an important role in the blood clotting mechanism. In a cascade process involving a complex system of several activators and inhibitors, plasminogen in the blood is transformed into active plasmin, which dissolves the blood clot by hydrolysis of fibrin fibres. Under certain circumstances, plasminogen infiltrated from the blood into the milk may also be activated. In such a case, active plasmin may cleave the milk proteins, and several (about 23) of the produced peptides decrease the enjoyment value of the milk by conferring it a typical bitter taste [1]. Therefore, the detection and determination of active plasmin in the milk is highly justified in the dairy industry. Plasmin detection methods comprise chromatographic, optical, electrical, acoustic wave and ELISA-based methods, with a broad range of LoD, the lowest being in the order of pM (Table 1).

Langmuir film balance and tensiometry is a versatile and widely used

technique for the study of molecular organization, molecular in-teractions, (bio)analytical chemistry and sensor development, drug development, etc. The basic concept of the technique is that the very high surface tension of pure water (γwater=~72 mN/m) is effectively reduced drastically by only a fraction of a monolayer of surface-active molecules. The behaviour of this Langmuir monolayer can be studied under different conditions. Upon any interaction of the interfacial molecules with components of the subphase, the surface tension changes. By measuring the interfacial forces with a fine balance, the surface pressure (Π ≡ γwater – γmonolayer) can be monitored routinely with a better than 0.05 mN/m precision. Thus, small changes of material amount can be detected.

β-casein is a single-chained milk protein of molecular mass of 24 kDa that consists of 209 amino acids, without a pronounced secondary structure [2]. Comprising both hydrophilic and hydrophobic regions, it is an amphiphilic compound, capable of forming a monolayer at the air/water interface, with an equilibrium surface pressure of Πe = 21–22 mN/m, stable in a wide temperature range (5–40 ◦C) [3]. However, as a protein, it is also water-soluble, thus the stability of this monolayer differs from the stability of true Langmuir monolayers. The amphiphilic character of the molecule is exploited in the food industry, where β-casein is used as a common emulsifier. In the past decades, the inter-facial behaviour of the β-casein has been more and more understood. Mitchell et al. found that surface pressure – area (Π–A) and surface pressure – surface concentration (Π–Γ) isotherms of β-casein mono-layers at the air/water interface are practically identical, suggesting weak cohesion forces, and a disordered state [4]. Boyd et al. [5] showed that the surface viscosity of β-casein monolayers at the air/water interface is very low and practically does not change with surface con-centration. When the subphase is slightly acidic (pH = 5), the Π–A isotherms are shifted towards higher Π and lower A values as compared to a subphase of pH = 7. Practically no hysteresis of the isotherm is observed upon repeated compression–expansion cycles of the very same β-casein monolayer [3], however, a critical analysis of the available literature [3,4,6,7] reveals a very large discrepancy among the iso-therms of different β-casein monolayers, which will be the topic of a forthcoming paper. The monolayer consist of a densely packed, ~2 nm thick layer in the air side of the interface and a loose, ~6 nm thick layer extending into the aqueous phase [8]. Monolayers adsorbed from a so-lution follow a non-Gibbs adsorption isotherm [9].

However, despite of the relatively vast knowledge gathered, β-casein monolayers have not been used for studying the cleavage by proteases. The objective of this work was to study the proteolysis of β-casein by plasmin at the air/water interface with a view of method development for simple, label-free enzyme activity assay.

2. Experimental

As an enzyme substrate, β-casein (from bovine milk, BioUltra, ≥98 %, Sigma-Aldrich, M ≈ 24,000 g/mol) was used. The inactive zymogen, plasminogen from bovine plasma (lyophilized powder, ≥2.0 units/mg protein, Sigma-Aldrich) was transported in dry ice (− 78.5 ◦C) and stored in deep freezer (− 30 ◦C). The activation was done according to [28,29], using urokinase from human urine (EMD Millipore), yielding 200 μL 1000 nM for each aliquot, as described in [24] based on the spectro-photometric plasmin activity determination of [30]. The enzymatic degradation experiments were run in a home-built round PTFE vessel of 50 mm diameter and 8 mm depth (19.6 cm2 surface area). The surface pressure was monitored by a computer-controlled NIMA PS4 (Coventry, UK) surface pressure sensor, using chromatographic paper strips (w =10 mm, t =0.3 mm). The water used as, respectively for the preparation of, subphases was ultrapure water obtained by reverse osmosis (ELGA LabWater, Veolia Water Technologies, UK). β-casein layers were formed by spreading 2 × 25 μL of 0.100–0.411 g/L (= 4.16–17.13 μM) β-casein solutions in phosphate buffered saline (PBS; from tablets; Sigma-Aldrich; pH = 7.4, composition: 10 mM Na2HPO4, 2 mM KH2PO4,

Table 1 Plasmin detection methods reported in the literature, in decreasing order of the limit of detection. –: not reported.

Limit of detection /pM

Method Receptors/Substrate Reference

– spectrophotometry Nα-CBZ-L-Lys-p-nitrophenyl ester [10] – spectrofluorometry L-Val-Gly-L-Arg

β-naphthylamide; t-BOC-L-Val- Gly-L-Arg β-naphthylamide; Gly- L-Arg β-naphthylamide

[11]

111,100 affinity chromatography

peptidylmethylcoumarylamide [12]

43,029 spectrofluorometry benzyloxycarbonyl-Ala-Ala-Lys- 4-methoxy-2-naphthylamine

[13]

12,900 spectrophotometry D-Ala-L-Leu-L-Lys-7-amido-4- trifluoromethylcoumarin

[14]

3680 spectrophotometry Spectrozyme-PL [15] 780 ELISA antibodies [16] 650 acoustic wave KTFKGGGGGGC [17] 602 electrochemical ferrocenylpropionyl-

KTFKGGGGGGC [18]

560 electrochemical ferrocenylpropionyl- KTFKGGGGGGC

[19]

500 acoustic wave β-casein adsorbed layer [20] 2601 bioluminescence

resonance energy transfer

GFP2-LQGSKKYKVKEGSLQ- RLuc2

[21]

258 spectrofluorometry

D-Ala-L-Leu-L-Lys-7-amido-4- trifluoromethylcoumarin

[14]

D-Val-L-Leu-L-Lys-7-amido-4- methylcoumarin

[22]

200 acoustic wave κ-casein adsorbed layer [20] 181 spectrophotometry D-Val-L-Leu-L-Lys -4-nitroanilide [23] 32 acoustic wave β-casein adsorbed layer [21]

302 bioluminescence resonance energy transfer

GFP2-LQGSKKYKVKEGSLQ- RLuc2 [21]

8.6 spectrofluorometry 7-(N-succinoyl-Ala-Phe-Lys)-4- methylcoumarin trifluoroacetate

[25]

1.2

spectrofluorometry N-succinyl-AFK-7-amido-4- methylcoumarin [26]

sandwich ELISA

monoclonal antibodies specific for plasminogen, respectively plasminogen and plasmin; p- nitrophenyl phosphate

[27]

1for bovine plasmin. 2for human plasmin.

L. Romanszki et al.

Page 3: Plasmin determination based on enzymatic digestion of a

Colloids and Surfaces A: Physicochemical and Engineering Aspects 609 (2021) 125786

3

2.7 mM KCl and 137 mM NaCl) with a Hamilton microsyringe. This range of amount of spread β-casein was chosen so that the obtained nominal specific area (assuming all β-casein molecules being at the interface), 0.4–0.1 m2/mg (corresponding to an area per molecule of 1565–381 Å2) fall in the liquid condensed region of the surface pressure – area isotherms [7]. For the enzymatic degradation experiments, 15 mL subphase was used. After the surface pressure reached a constant value (in typically 3− 6 h), 200 μL undiluted or diluted enzyme solution was added to the subphase by a Hamilton microsyringe through a small channel. In the followings, ci

PL will be used to denote the initial con-centration of the injected plasmin solution, whereas cf

PL will be used for the corresponding final concentration calculated for the entire volume (15 mL), after injection. The temperature of air was 27 ◦C. After each experiment, the subphase was aspirated dry by the help of a vacuum pump (KNF Neuberger Laboport, Freiburg, Germany), the vessel was cleaned with lint-free cleaning wipers (Kimtech Pure 7624) soaked in chloroform. During the experiments, the subphase was continuously

stirred with a small magnetic bar positioned at maximum distance from both the vessel wall and the paper strip of the surface pressure sensor. The experimental setup is presented in Fig. 1. Data analysis, fitting and plotting was done in OriginPro (OriginLab Co., USA).

3. Results and discussion

In Fig. 2, typical kinetics of the β-casein monolayer formation and the change of surface pressure following the addition of plasmin is shown. The baseline is recorded and levelled in AB. At point B, the β-casein is added to the air/water interface. Consequently, the surface pressure rises, steeply at the beginning (BC, immediate spreading as a Langmuir monolayer), and slowly afterwards (CD, probably readsorption from the subphase, as a Gibbs monolayer), until an equilibrium surface pressure is reached. Then, in point D, the activated enzyme solution is introduced, which causes an exponential decrease of the surface pressure (DE), until the new stable surface pressure is reached. This exponential decrease is attributed to the reorganization of the interface following the dissolu-tion of some of the peptide products. This is possible because 1) the main cleavage sites of the β-casein for PL are located in the hydrophilic re-gions of the molecule, thus most likely facing the aqueous phase and the PL dissolved therein; and 2) some of the peptide fragments are fully hydrophilic [1], with high potential for dissolution in the aqueous subphase. However, the majority of the peptide residues are fully or predominantly hydrophobic, thus being expected to remain at the interface. This explains why the ΔΠ surface pressure change is always small relative to the surface pressure of the intact β-casein monolayer, ultimately limiting the sensitivity of the method. To obtain calibration curves, either the ΔΠ surface pressure difference, or the rmax opposite of the maximum rate of surface pressure decrease (1) is then plotted against the enzyme concentration.

rmax≝ − (dΠ/dt)max (1)

The calibration curves obtained by fitting the data points with (2) reveal a maximum sensitivity at low PL concentrations that levels off at higher PL concentrations (Fig. 3). The sensitivities and the plateau values are higher for higher β-casein concentrations. The three fitting parameters, A, κ and y0 are plotted in Fig. 4 against the β-casein concentration.

Fig. 1. Schematic representation of the experimental setup: a) Teflon vessel; b) subphase (water); c) surface pressure sensor with hanging filter paper strip; d) syringe with β-casein solution; e) syringe with plasmin solution; f) β-casein interfacial layer; g) plasmin; h) debris from digested β-casein; i) magnetic bar; j) stirrer; k) cabinet.

Fig. 2. A typical surface tension-based plasmin concentration measurement with β-casein monolayer, presented on the example of cPL

i =1 μM solution (cPLf =

13.3 nM). AB: baseline (pure water); BC: β-casein spreading; CD: β-casein spreading and readsorbing; D: adding the plasmin into the subphase; DE: β-casein layer digested, dissolution of reaction products into the subphase causing a ΔΠ drop with an rmax ≡ –(dΠ/dt)max maximum rate.

Fig. 3. Calibration data points and fitted exponential curves for plasmin con-centrations obtained with five different β-casein concentrations: cβ-casein =

0.100 (⬛), 0.167 ( ), 0.237 ( ), 0.325 ( ), and 0.411 ( ) g/L; –ΔΠ surface pressure changes as a function of the plasmin concentration: ci

PL – initial con-centration of the added plasmin solution, before injection; cf

PL – corresponding final concentration calculated for the entire volume, after injection.

L. Romanszki et al.

Page 4: Plasmin determination based on enzymatic digestion of a

Colloids and Surfaces A: Physicochemical and Engineering Aspects 609 (2021) 125786

4

− ΔΠ = Aexp(− ci

PL

/κ)+ y0 (2)

Interestingly, adding β-casein to the air/water interface in two different volumes and concentrations so that their product, i.e., the amount of substance is kept equal, produces two very different cali-bration curves. The one corresponding to the higher concentration and lower volume substrate shows higher sensitivity and higher plateau value than the one corresponding to the lower concentration and higher volume substrate (Fig. 5). The discrepancy is most likely related to the spreading mechanism of the β-casein at the air/water interface, which

partially resembles the spreading of a true, insoluble Langmuir film, and partially the spreading of a soluble, Gibbs monolayer.

As the other calibration possibility, rmax values, determined by derivation of the DE sections of the measurement curves (Fig. 2), were plotted against PL concentration (Fig. 6). In spite of the experimental difficulties (those related to the non-ideal spreading of the substrate, the eventual differences of the enzyme mass transport from experiment to experiment due to the possible irregularities of the convection, the high number of the cleavage sites, etc.), the conspicuously linear trends of the calibration data are impressive. Not surprisingly, as in the case of surface pressure change calibration plots, the sensitivity (the slope of the lines)

Fig. 4. Fitting parameters A (A), κ (B) and y0 (C) of (2) as a function of cβ-casein.

Fig. 5. Calibration data points and fitted exponential curves for plasmin con-centrations obtained with 1 × 25 μL 0.2 g/L (⬛), respectively 2 × 25 μL 0.1 g/L β-casein ( ); –ΔΠ surface pressure changes as a function of the plasmin con-centration: ci

PL – nominal concentration of the added plasmin solution, before injection; cf

PL – corresponding concentration calculated for the entire volume, after injection.

Fig. 6. Calibration lines for plasmin concentrations obtained with five different β-casein concentrations: cβ-casein = 0.100 (⬛), 0.167 ( ), 0.237 ( ), 0.325 ( ), and 0.411 ( ) g/L; rmax maximum rates of surface pressure decrease as a function of the plasmin concentration: ci

PL – nominal concentration of the added plasmin solution, before injection; cf

PL – corresponding concentration calculated for the entire volume, after injection. rmax in experiment-relevant units (left axis) and in basic SI units (right axis).

L. Romanszki et al.

Page 5: Plasmin determination based on enzymatic digestion of a

Colloids and Surfaces A: Physicochemical and Engineering Aspects 609 (2021) 125786

5

generally increases with increasing substrate concentration. Based on the data presented in Fig. 6, the linearity of the assay permits enzyme activity measurements down to 50 nM initial, corresponding to 0.7 nM final, concentration.

Fig. 7 is another representation of the same calibration data, where the rmax values are plotted against the substrate concentration, grouped for different enzyme concentrations. The curves, partially extrapolated, show a sigmoidal run, with maximum steepness at intermediate sub-strate concentrations, decreasing maximum steepness with decreasing PL concentration and maximum rmax for highest PL and substrate concentrations.

A final plot summarizing the sensitivity of the method is shown in Fig. 8. The rmax/cPL ratios, that are in fact the slopes of the calibration lines of Fig. 6, are plotted against the substrate concentration. The ob-tained points follow a parabolic trend that is very close to linearity, telling that the sensitivity of the method after all increases with substrate concentration. It should be noted, however, that the sensitivity most probably cannot be increased beyond a certain critical value by using increasingly higher substrate concentrations, which is expected to be near the critical micelle concentration of the β-casein, approx. 0.5 g/L [31].

4. Conclusions

β-casein monolayers at the air/water interface have been digested with plasmin, a proteolytic enzyme, injected into the subphase, and the surface pressure drop was observed in the function of applied enzyme concentration. The presented method allows the determination of plasmin with an assay linearity down to 50 nM sample concentration in the given experimental setup, corresponding to 0.7 nM final concen-tration. The advantage of the method consists in its simplicity. The several hours needed to reach the starting equilibrium surface pressure of the substrate interfacial film definitely constitutes a limitation, just as the measurement variant based on the surface pressure difference determination, the variant based on the maximum rate of surface pres-sure change being considerably faster. Ultimately, the source of all weaknesses of the method lies in the limited film-forming ability of the β-casein substrate. Although being amphiphilic, β-casein is also soluble in the aqueous subphase which is spread upon, consequently does not form a proper, i.e. insoluble, Langmuir monolayer at the interface, but it is partially dissolved in, and (again partially) readsorbed to the inter-face. Nevertheless, here we reported for the first time a plasmin deter-mination method based on tensiometric measurements of an interfacial layer of a natural substrate of the enzyme, β-casein.

Further ongoing work is aimed at optimising experimental variables, offering a substantial decrease of the detection limit of enzyme activity as well as enhancing the robustness of the method.

CRediT authorship contribution statement

Lorand Romanszki: Conceptualization, Methodology, Investiga-tion, Formal analysis, Writing - original draft, Writing - review & edit-ing. Tibor Hianik: Conceptualization, Methodology, Resources, Writing - review & editing, Project administration, Funding acquisition, Super-vision. Zsofia Keresztes: Conceptualization, Methodology, Resources, Writing - review & editing, Project administration, Funding acquisition.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

The research leading to these results has received funding from the European Commission within the project Innovative technology for the detection of enzyme activity in milk (FORMILK) under grant agreement number 690898/H2020-MSCA-RISE-2015, and the <GS1>BIONANO_GINOP-2.3.2-15-2016-00017 project. T.H. acknowl-edges funding from the Science Grant Agency VEGA, project No.: 1/ 0419/20.

References

[1] V.M. Rauh, L.B. Johansen, R. Ipsen, M. Paulsson, L.B. Larsen, M. Hammershøj, Plasmin activity in UHT milk: relationship between proteolysis, age gelation, and bitterness, J. Agric. Food Chem. 62 (2014) 6852–6860, https://doi.org/10.1021/ jf502088u.

[2] H.E. Swaisgood, The caseins, CRC Crit. Rev. Food Technol. 3 (1973) 375–414, https://doi.org/10.1080/10408397309527148.

[3] M.R. Rodríguez Nino, C.C. Sanchez, J.M. Rodríguez Patino, Interfacial characteristics of β-casein spread films at the air–water interface, Colloids Surf. B Biointerfaces 12 (1999) 161–173, https://doi.org/10.1016/S0927-7765(98) 00072-1.

[4] J. Mitchell, L. Irons, G.J. Palmer, A study of the spread and adsorbed films of milk proteins, Biochimica et Biophysica Acta (BBA) - Protein Structure 200 (1970) 138–150, https://doi.org/10.1016/0005-2795(70)90052-8.

[5] J.V. Boyd, J.R. Mitchell, L. Irons, R.P. Musselwhite, P. Sherman, The mechanical properties of milk protein films spread at the air-water interface, J. Colloid Interface Sci. 45 (1973) 478–486, https://doi.org/10.1016/0021-9797(73)90163- X.

Fig. 7. rmax maximum rates of surface pressure decrease as a function of the cβ-

casein for five different plasmin concentrations: cPLi / cPL

f = 1000 nM / 13.3 nM (⬛), 500 nM / 6.67 nM ( ), 333 nM / 4.44 nM ( ), 250 nM / 3.33 nM ( ), and 100 nM / 1.33 nM ( ), in experiment-relevant units (left axis) and in basic SI units (right axis).

Fig. 8. The sensitivity of the PL determination (slope of the lines in Fig. 6) increases with cβ-casein; linear (red) and parabolic (blue) fits. Left axis: sensitivity expressed in experiment-relevant units; right axis: sensitivity in basic SI units. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article).

L. Romanszki et al.

Page 6: Plasmin determination based on enzymatic digestion of a

Colloids and Surfaces A: Physicochemical and Engineering Aspects 609 (2021) 125786

6

[6] M. Mellema, D.C. Clark, F.A. Husband, A.R. Mackie, Properties of β-casein at the air/water interface as supported by surface rheological measurements, Langmuir 14 (1998) 1753–1758, https://doi.org/10.1021/la970651h.

[7] J.M. Rodríguez Patino, C.C. Sanchez, M.R. Rodríguez Nino, Structural and morphological characteristics of β-casein monolayers at the air–water interface, Food Hydrocoll. 13 (1999) 401–408, https://doi.org/10.1016/S0268-005X(99) 00025-9.

[8] E. Dickinson, D.S. Horne, J.S. Phipps, R.M. Richardson, A neutron reflectivity study of the adsorption of βcasein at fluid interfaces, Langmuir 9 (1993) 242–248, https://doi.org/10.1021/la00025a046.

[9] J. Benjamins, J.A. De Feijter, M.T.A. Evans, D.E. Graham, M.C. Phillips, Dynamic and static properties of proteins adsorbed at the air/water interface, Faraday Discuss. Chem. Soc. 59 (1975) 218–229, https://doi.org/10.1039/DC9755900218.

[10] R.M. Silverstein, The plasmin-catalyzed hydrolysis of Nα-CBZ-L-lysine-p- nitrophenyl ester, Thromb. Res. 3 (1973) 729–736, https://doi.org/10.1016/0049- 3848(73)90019-4.

[11] W. Nieuwenhuizen, G. Wijngaards, E. Groeneveld, Fluorogenic peptide amide substrates for the estimation of plasminogen activators and plasmin, Anal. Biochem. 83 (1977) 143–148, https://doi.org/10.1016/0003-2697(77)90519-X.

[12] H.K. Mousavi, Z.M. Nikougoftar, Preparation of plasminogen by affinity chromatography, Iran. J. Blood Cancer 6 (2014) 165–168.

[13] S.A. Clavin, J.L. Bobbitt, R.T. Shuman, E.L. Smithwick, Use of peptidyl-4-methoxy- 2-naphthylamides to assay plasmin, Anal. Biochem. 80 (1977) 355–365, https:// doi.org/10.1016/0003-2697(77)90656-X.

[14] R.E. Smith, E.R. Bissell, A.R. Mitchell, K.W. Pearson, Direct photometric or fluorometric assay of proteinases using substrates containing 7-amino-4- trifluoromethylcoumarin, Thromb. Res. 17 (1980) 393–402, https://doi.org/ 10.1016/0049-3848(80)90074-2.

[15] N.A. Andreani, L. Carraro, L. Fasolato, S. Balzan, R. Lucchini, E. Novelli, B. Cardazzo, Characterisation of the thermostable protease AprX in strains of Pseudomonas fluorescens and impact on the shelf-life of dairy products: preliminary results, Ital. J. Food Saf. 5 (2016) 239–244, https://doi.org/10.4081/ ijfs.2016.6175.

[16] J.C. Collin, P. Compagnone, I. Ryba, A. Baer, Determination of plasmin (alkaline milk proteinase) and chymosin in milk products by the ELISA assay, Le Lait 68 (1988) 235–240, https://doi.org/10.1051/lait:1988216.

[17] A. Poturnayova, I. Karpisova, G. Castillo, G. Mezo, L. Kocsis, A. Csampai, Z. Keresztes, T. Hianik, Detection of plasmin based on specific peptide substrate using acoustic transducer, Sens. Actuators B: Chem. 223 (2016) 591–598, https:// doi.org/10.1016/j.snb.2015.09.143.

[18] K. Ohtsuka, I. Maekawa, M. Waki, S. Takenaka, Electrochemical assay of plasmin activity and its kinetic analysis, Anal. Biochem. 385 (2009) 293–299, https://doi. org/10.1016/j.ab.2008.11.006.

[19] G. Castillo, K. Pribransky, G. Mezo, L. Kocsis, A. Csampai, K. Nemeth, Z. Keresztes, H. Hianik, Electrochemical and photometric detection of plasmin by specific peptide substrate, Electroanalysis 27 (2015) 789–798, https://doi.org/10.1002/ elan.201400622.

[20] M. Tatarko, E.S. Muckley, V. Subjakova, M. Goswami, B.G. Sumpter, T. Hianik, I. N. Ivanov, Machine learning enabled acoustic detection of sub-nanomolar concentration of trypsin and plasmin in solution, Sens. Actuators B: Chem. 272 (2018) 282–288, https://doi.org/10.1016/j.snb.2018.05.100.

[21] H. Dacres, J. Wang, A. Anderson, S.C. Trowell, A rapid and sensitive biosensor for measuring plasmin activity in milk, Sens. Actuators B: Chem. 301 (2019), 127141, https://doi.org/10.1016/j.snb.2019.127141.

[22] T. Tervo, N. Honkanen, G.B. Van Setten, T. Virtanen, A. Tarkkanen, M. Harkonen, A rapid fluorometric assay for tear fluid plasmin activity, Cornea 13 (1994) 148–155, https://doi.org/10.1111/j.1600-0420.1997.tb00109.x.

[23] H.S. Rollema, S. Visser, J.K. Poll, Spectrophotometric assay of plasmin and plasminogen in bovine-milk, Milchwissenschaft – Milk Sci. Int. 38 (1983) 214–217.

[24] L. Romanszki, M. Tatarko, M. Jiao, Z. Keresztes, T. Hianik, M. Thompson, Casein probe–based fast plasmin determination in the picomolar range by an ultra-high frequency acoustic wave biosensor, Sens. Actuators B: Chem. 275 (2018) 206–214, https://doi.org/10.1016/j.snb.2018.08.025.

[25] P.A. Pierzchala, C.P. Dorn, M. Zimmerman, A new fluorogenic substrate for plasmin, Biochem. J. 183 (1979) 555–559, https://dx.doi.org/10.1042% 2Fbj1830555.

[26] T. Saint-Denis, G. Humbert, J.L. Gaillard, Enzymatic assays for native plasmin, plasminogen and plasminogen activators in bovine milk, J. Dairy Res. 68 (2001) 437–449, https://doi.org/10.1017/S0022029901004939.

[27] D. Dupont, C. Bailly, J. Grosclaude, J.C. Collin, Differential titration of plasmin and plasminogen in milk using sandwich ELISA with monoclonal antibodies, J. Dairy Res. 64 (1997) 77–86, https://doi.org/10.1017/s0022029996001938.

[28] M.R. García-Risco, I. Recio, E. Molina, R. Lopez-Fandino, Plasmin activity in pressurized milk, J. Dairy Sci. 86 (2003) 728–734, https://doi.org/10.3168/jds. s0022-0302(03)73653-4.

[29] D.L. Schroeder, S.S. Nielsen, K.D. Hayes, The effect of raw milk storage temperature on plasmin activity and plasminogen activation in pasteurized milk, Int. Dairy J. 18 (2008) 114–119, https://doi.org/10.1016/j.idairyj.2007.08.003.

[30] K. Kolev, I. Lerant, K. Tenekejiev, R. Machovich, Regulation of fibrinolytic activity of neutrophil leukocyte elastase, plasmin, and miniplasmin by plasma protease inhibitors, J. Biol. Chem. 269 (1994) 17030–17034.

[31] D.G. Schmidt, T.A.J. Payens, The evaluation of positive and negative contributions to the second virial coefficient of some milk proteins, J. Colloid Interface Sci. 39 (1972) 655–662, https://doi.org/10.1016/0021-9797(72)90073-2.

L. Romanszki et al.