You are on page 1of 10

Innovative Food Science and Emerging Technologies 68 (2021) 102639

Contents lists available at ScienceDirect

Innovative Food Science and Emerging Technologies


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

Continuous production of tempe-based bioactive peptides using an


automated enzymatic membrane reactor
Azis Boing Sitanggang *, Julius Sumitra , Slamet Budijanto
Department of Food Science and Technology, IPB University, Kampus IPB Darmaga, Bogor 16680, Indonesia

A R T I C L E I N F O A B S T R A C T

Keywords: The utilization of bioactive peptides from food-grade raw materials has received increasing interest, especially
AICE inhibitor for the production of functional foods. Within this study, a combination of fermentation and in vitro papain
Antioxidant hydrolysis was devised to produce bioactive peptides exhibiting antioxidant property and angiotensin I-con­
Bioactive peptides
verting enzyme (AICE) inhibitory activity. The utilization of an automated membrane reactor system developed
Enzymatic membrane reactor
Papain
in this study could facilitate the continuous hydrolysis of tempe flour-rich in protein under constant flux, thus
Tempe constant residence time operation. The optimum operating conditions were: [E]/[S] = 10% (w/v) and τ = 9 h.
With a 10-kDa membrane filtration, the resulted antioxidant capacity and AICE inhibition were 0.033 mg AEAC/
mL and 50.9%, respectively. Further separation of permeate with a 2-kDa membrane cut-off leveled off the
antioxidant capacity but increased the AICE inhibitory activity (i.e., Reduction of IC50 of ~40%). By this, the
obtained IC50 for AICE was 0.08 mg protein/mL, and this value was comparable with the IC50 values reported in
the literature.
Industrial relevance: The production of bioactive peptides from soybean through a combination of fermentation
and in vitro enzymatic hydrolysis is limitedly reported. In this work, fermented soybean (tempe) was further
hydrolyzed continuously using an automated membrane reactor. The hydrolysis under optimum operating
conditions could increase the antioxidant activity and AICE inhibition of permeate. This work is expected to
increase the commercialization of tempe-based food products by using safer technological approaches (i.e.,
fermentation and membrane technology).

1. Introduction technology that can utilize tempe for producing functional ingredients.
Bioactive peptides are peptides having 2–20 amino acids within the
Tempe (Indonesian spelling), referred to as tempeh, is a soybean- structures of the parent proteins (Agyei, 2015; Sanjukta & Rai, 2016).
based fermented food widely consumed in Indonesia (Astuti, Meliala, These peptides must be released from parent proteins in order to have
Dalais, & Wahlqvist, 2000). Tempe is popular as nutritious non-meat health-beneficial activities, at least in three ways: (i) through in vivo
protein food, especially in the USA, Japan, and Europe (Nout & Kiers, hydrolysis by gastrointestinal proteolytic enzymes, (ii) using microbial
2005). Tempe has received increased interest, and numerous studies fermentation, and (iii) through in vitro hydrolysis by the actions of
have demonstrated tempe as an attractive fermented food related to the microorganism- or plant-based proteases (Banan-Mwine Daliri, Oh, &
prevention of various chronic diseases (Mani & Ming, 2017). Tempe has Lee, 2017; Korhonen & Pihlanto, 2006; Liu & Zhao, 2010). In such,
been reported to have a potent antioxidant capacity, antimicrobial ac­ bioactive peptides have physiological functionalities, especially for
tivity, and positive effects on intestinal microbiota, due to a high content cardiovascular system (antioxidative, antithrombotic, and antihyper­
of isoflavone, folate, vitamin B12, and other compounds (Cao, Green- tensive), nervous system, gastrointestinal system (mineral binding,
Johnson, Buckley, & Lin, 2019). On the other hand, however, tempe antimicrobial activity), and immune system (immunomodulatory)
still has several drawbacks in terms of consumer acceptance. The bitter (Chalamaiah, Keskin Ulug, Hong, & Wu, 2019; Korhonen & Pihlanto,
taste and unpleasant odor are the two main drawbacks of tempe. These 2006; Sanjukta & Rai, 2016; Tanzadehpanah, Asoodeh, Saberi, & Cha­
have caused tempe commercialization in the market, limited only as a mani, 2013).
whole food. Therefore, there is a need to provide an alternative The protein content in soybean is approximately 30–48% (dry basis,

* Corresponding author.
E-mail address: boing.lipan@apps.ipb.ac.id (A.B. Sitanggang).

https://doi.org/10.1016/j.ifset.2021.102639
Received 12 November 2020; Received in revised form 10 February 2021; Accepted 10 February 2021
Available online 16 February 2021
1466-8564/© 2021 Elsevier Ltd. All rights reserved.
A.B. Sitanggang et al. Innovative Food Science and Emerging Technologies 68 (2021) 102639

db) (Syah et al., 2015). With this amount, soybean is a good source of
parent proteins to produce bioactive peptides through different ap­
proaches mentioned previously. Singh, Vij, and Hati (2014) has
reviewed the functional significances of bioactive peptides derived from
soybean. These include antihypertensive, antioxidative, antiobesity,
immunomodulatory, anti-diabetic, hypocholesterolemic, and anti­
cancer. Gibbs, Zougman, Masse, and Mulligan (2004) reported angio­
tensin I-converting enzyme (AICE) inhibitory, anti-thrombotic, and
antioxidant properties of bioactive peptides isolated from soy-fermented
foods, like natto and tempe. Rho, Lee, Chung, Kim, and Lee (2009) also
demonstrated the production of AICE inhibitory peptide from fermented
soybean extract.
During soybean fermentation, the proteins are only partially hy­
drolyzed. It is due to the mass-transfer limitation of proteases to access
parent proteins (i.e., inside bean kernels). Additionally, most of the
extracellular proteases only cleave glycoproteins and phosphoproteins
(Gibbs et al., 2004). Herein, the fermentation of soybean (tempe), and
followed by in vitro hydrolysis, is considered as an apt method to pro­
duce small molecular weight bioactive peptides.
In our previous study, we reported a combined method between Fig. 1. Automated enzymatic membrane reactor (EMR) (No. 1 = Nitrogen (N2),
2 = pressure reducer, 3 = PPP-MPPES, 4 = substrate tank, 5 = reactor, 6 = UF
fermentation and in vitro hydrolysis for producing tempe-based bioac­
membrane, 7 = analytical balance, 8 = water bath system, 9 = PC with Lab­
tive peptides (Sitanggang, Lesmana, et al., 2020). The hydrolysis of
VIEW program).
tempe flour using papain and followed by 5-kDa membrane filtration
obtained the highest bioactivities (i.e., antioxidant activity, AICE-,
purchased from Microdyn-Nadir GmbH (Germany).
α-glucosidase-inhibitory activity). Additionally, the influence of mo­
lecular weight on the bioactivities of peptides was also demonstrated.
The smaller membrane cut-off used for sieving the peptides, the higher 2.2. Tempe production and preparation of tempe flour-rich in protein
the bioactivities obtained (Ajibola, Fashakin, Fagbemi, & Aluko, 2011;
Sonklin, Laohakunjit, & Kerdchoechuen, 2018). The limitation in our The fermentation of soybean was conducted according to our pre­
previous study was that tempe flour was hydrolyzed batch-wise vious study (Sitanggang, Lesmana, et al., 2020). Soybean (soybean-to-
(Sitanggang, Lesmana, et al., 2020). The batch reaction is often water ratio = 1:5 (w/w)) was boiled at 90–95 ◦ C for 90 min. The
considered unproductive due to a time-consuming intermittent stop fermentation was performed by inoculating the starter culture (also
between operation cycles (Sitanggang, Drews, & Kraume, 2016; Ubilla known as ‘laru’, composed mainly by Rhizopus spp., and growth media)
et al., 2020). To our best of knowledge, continuous production of as much as 2.0 g per kg soybean, at 32 ◦ C, relative humidity of 75% for
bioactive peptides in membrane reactors is mainly applied to milk six days. The resulted tempe was cut into dices (3 × 0.5 × 0.5 cm), dried
proteins (Korhonen & Pihlanto, 2006). It is due to the high solubility of in a tray dryer (Terara Seisakusho C. Ltd. No 4-60SP, Japan) at 70 ◦ C for
milk proteins. Within this study, an attempt to continuously hydrolyze 7 h. The dried tempe was then comminuted using FFC 23 pin disc-mill
tempe flour-rich in protein to produce bioactive peptides was realized. (Agowindo-Maksindo, Indonesia).
The automated membrane reactor system developed in this work could Tempe flour was defatted using technical grade hexane following
circumvent the limitations of batch-wise production of bioactive pep­ Rosset, Acquaro, and Beléia (2014). The defatted tempe flour was
tides (i.e., low productivity and enzyme utilization). Permeate exhibit­ dispersed in distilled water with a ratio of 1:10 (w/v), and the pH was
ing ACE inhibition and antioxidant activity is of interest because these adjusted to ~8 using 2.0 N NaOH. The solution was stirred at 200 rpm
bioactivities are dominating in tempe-based peptides (Sitanggang, Les­ for 2 h at 30 ◦ C. For facilitating phase separation, the mixture was
mana, et al., 2020). This work is expected to increase the commercial­ allowed to stand for 24 h. The upper phase containing dissolved proteins
ization of tempe-based functional ingredients by using safer was withdrawn, and its pH was brought down to the protein’s isoelectric
technological approaches (i.e., fermentation and membrane point (pI = ~ 4.5) using 1.0 N HCl. The slurry at the bottom of the
technology). mixture was obtained, neutralized with 2.0 N NaOH, and finally dried in
a freeze dryer (Liu et al., 2008). The resulted solid was milled and sieved
2. Materials and methods (Θ = 80 Tyler mesh) to obtain tempe flour-rich in protein.

2.1. Materials 2.3. Enzymatic membrane reactor (EMR)

Soybean was purchased from Koperasi Tempe Indonesia (KOPTI), The enzymatic membrane reactor (EMR) was constructed in parallel
Bogor, Indonesia. Technical grade hexane (80% v/v) and pure water (n = 2, Fig. 1). The upper and bottom part of the reactor (5) and sub­
(water one™) were purchased from Sentral Kimia and Jayamas Medica strate tank (4) were made of 304 stainless steel. The reactor body was
Industri, Indonesia, respectively. Folin–Ciocalteu reagent, methanol pro made of borosilicate glass (wall thickness = 3 mm, diameter = 48 mm,
analysis, 2,2-diphenyl-1-picrylhydrazyl (DPPH), bovine serum albumin height = 72 mm). The reactor volume was 9 × 10− 2 L, with a D/d ratio
(BSA), NaOH, HCl, Na2CO3 were purchased from Sigma-Aldrich, Inc. of 1.175 and the maximum pressure stability of 6.0 bar. The UF mem­
(Singapore). Hippuryl-histidine-leucine (HHL) was from Bachem AG brane (6) having an effective membrane area of 1.24 × 10− 3 m2 was
(Switzerland). The 4-(2-Hydroxyethyl)-1-piperazine ethanesulfonic acid placed at the bottom of the reactor. The nitrogen (N2) (1) was used to
(HEPES) was purchased from Shanghai Chaining Chemicals Co., Ltd. pump the substrate from the substrate tank to the reactor. The release of
(China). Papain (100,000 U/g), and angiotensin I-converting enzyme nitrogen was controlled by means of MPPES-3-1/4–6-010 proportional
(AICE) were purchased from, Xi’An Saiyang Bio-Technology Co., Ltd. pressure regulator (Festo SE & Co. KG, Germany) (3). Permeate was
(China), and Xi’An Saina Biological Technology Co., Ltd. (China), collected in a container that was located on the analytical balance (7).
respectively. The UF flat-sheet polyethersulfone (PES) membranes The automation developed in this study was to bring the process
(molecular weight cut-off/MWCO of 2-, 4-, 5-, and 10-kDa) were variable flux JPV as close as possible to a set flux JSV, thus resulted in a

2
A.B. Sitanggang et al. Innovative Food Science and Emerging Technologies 68 (2021) 102639

small control error (Sitanggang et al., 2016; Sitanggang, Drews, & 3):
Kraume, 2015). A program was realized using Laboratory Virtual In­

strument Engineering Workbench software (LabVIEW, National In­ J= (3)
struments, USA) where the proportional-integral-derivative (PID) Am
controller was embedded (Sitanggang, Drews, & Kraume, 2014a), and From Eq. (2) and Eq. (3), one could see that controlling flux invariant
tuned according to Kuhn (1995). To facilitate this automation, the data during the operation would also maintaining residence time constant
acquisition (DAq) was supported by compact DAQ USB chassis (cDAQ)- because the reactor volume and the effective membrane area did not
9174, NI 9263, and NI 9201 (National Instruments, USA). change respective to the time.

2.4. Membrane cleaning 2.7. Measurement of enzyme’s activity

The membrane was cut with a diameter of 48 mm. The cleaning was A total of 5.0 mL of casein (0.65% w/v) was poured in a 15-mL
done by soaking the membrane in 0.1% (w/v) NaOH for 15 min. The centrifuge tube and added with 1.0 mL of papain (0.5% w/v). The so­
membrane was rinsed and placed at the bottom of the reactor prior to lution was vortexed and incubated for 10 min at 37 ◦ C. Furthermore, 5.0
the filtration of pure water at a constant transmembrane pressure ∆P mL of 0.5 M trichloroacetic acid were added into the reacting solution to
(TMP) of 0.5 bar for 20 min. stop the hydrolysis. The mixture was centrifuged at 1000 xg (Hermle Z
383 K, Wehingen, Germany), at 25 ◦ C for 10 min. A 1.0-mL of the su­
2.5. Papain filtration pernatant was withdrawn and added with 5.0 mL of 0.5 M sodium
carbonate solution and 1.0 mL of Folin-Ciocalteu (1:5 v/v). The mixture
The tempe flour-rich in protein was continuously hydrolyzed using was vortexed and incubated for 30 min at 37 ◦ C. The absorbance was
freely-dispersed papain. Herein, the selection of membrane’s MWCO has monitored at a wavelength of 660 nm using a Genesys™ 140 UV–Vis
to ensure complete rejection of the enzyme molecules. The filtration of Spectrophotometer (USA). Tyrosine (0–60 ppm) was used to construct
papain was performed using two MWCOs (i.e., 5- and 10-kDa). Papain the standard curve (y = 0.0143× + 0.0079, R2 = 0.992). One unit of
(2.5% (w/v)) was dissolved in 0.01 M phosphate buffer (pH 6.8) and activity (U) was defined as the amount of tyrosine (in micromoles)
subject to the filtration at a constant flux. The enzyme’s activity in released during the proteolytic action of papain on the casein per minute
permeate was measured, and the rejection R was calculated as follows (Cupp-Enyard, 2008).
(Eq. 1):
( )
Ui− Up 2.8. Measurement of protein content
R = × 100% (1)
Ui
The protein content of tempe flour-rich in protein was measured
where Ui and Up were enzyme’s activity (U/g) at initial and in the according to the standard methods from the Association of Official
permeate, respectively. Agricultural Chemists (AOAC). In addition to this, the protein content of
permeate was monitored following Lowry’s method (Lowry et al., 1951)
2.6. Continuous production of tempe-based bioactive peptides in which bovine serum albumin (0–150 ppm) was used to construct the
standard curve (y = 0.0054× + 0.0384, R2 = 0.992).
The substrate was prepared by dissolving tempe flour-rich in protein
with 0.01 M phosphate buffer (pH 6.8) with a ratio of solid-to-buffer was 2.9. Analysis of total phenol
1:100 (w/v) in a beaker glass. The solution was rested for about 15 min,
and two-third of the substrate solution from the upper part was further A 0.2-mL sample was poured into the test tube and added with 1.8
utilized, thus leaving coarse suspension at the bottom of the beaker mL of fresh Folin-Ciocalteau reagent. The mixture was incubated at
glass. The substrate was then placed in the substrate tank (4) and the room temperature for 6 min. A total of 1.8 mL of Na2CO3 (6% w/v) was
reactor (5, see Fig. 1). The reaction was carried out at a temperature T of added to the solution prior to the incubation for 90 min in a dark place at
40 ◦ C, and agitation N of 300 rpm. Within this study, the influences of room temperature. The absorbance was measured at a wavelength of
enzyme-to-substrate ratio (i.e., [E]/[S] = 5, 10 and 15%) and residence 725 nm. Gallic acid (0–200 ppm) was used to construct the standard
time (i.e., τ = 5, 7 and 9 h) on the production of bioactive peptides were curve (y = 0.0047× + 0.009, R2 = 0.997). Total phenol was expressed as
investigated. For analyses, samples were withdrawn intermittently milligram gallic acid equivalent (mg GAE) per milliliter sample.
(every one hour) from the permeate pipe and also from the container
placed on the analytical balance. The sample from the container was 2.10. Measurement of antioxidant capacity
only taken once, at the end of the operation. This sample was considered
to represent the cumulative or overall reaction performance. In addition A 0.3-mL of the sample was added with 0.7 mL of distilled water, and
to this, permeate obtained from the optimum operating conditions was 3.0 mL of 120 μM 2,2-diphenyl-1-picrylhydrazyl (DPPH). The vortexed
filtered subsequently using 5-, 4- and 2-kDa PES membrane. The half- solution was incubated for 30 min in a dark room at room temperature.
maximal inhibitory concentration (IC50) of the resulted permeate was The absorbance was measured at 515 nm. The antioxidant capacity was
measured especially for the antioxidant activity and angiotensin I-con­ expressed in milligram ascorbic acid equivalent antioxidant capacity
verting enzyme inhibitory activity. (mg AEAC) per milliliter or gram sample (y = − 0.0083× + 0.9446, R2 =
For continuous operation where the density does not change in the 0.995) (Brand-Williams, Cuvelier, & Berset, 1995; Sitanggang, Lesmana,
system, the hydraulic residence time τ is calculated as (Eq. 2): et al., 2020; Sitanggang, Sinaga, Wie, Fernando, & Krusong, 2020).
VR
τ= (2)
V̇ 2.11. Measurement of angiotensin I-converting enzyme (AICE) inhibitory
activity
where VR = reactor volume (L) and V̇ = permeate volumetric flow rate
(L/h). During the continuous production of bioactive peptides, the A 200-μL of buffer (6.7 mM hipuryl-l-histidyl-l-leucine in 50 mM 4-
reactor was fully filled with substrate, and the reactor volume was (2-hydroxyethyl) -1-piperazineethanesulfonic acid with 300 mM NaCl,
maintained constant. Herein, permeate flux J (L/m2.h) was dependent pH 8.3) was added with 100 μL of sample, and pre-incubated at 37 ◦ C for
on the volumetric flow rate V̇ and effective membrane area Am (m2) (Eq. 5 min. A 30-μL AICE (0.33 U/mL) was added to the mixture before the

3
A.B. Sitanggang et al. Innovative Food Science and Emerging Technologies 68 (2021) 102639

Table 1
Several characteristics of tempe flour-rich in protein.
Parameters –

Moisture content (% wb) 12.25 ± 0.24


Protein content (% db) 51.60
Antioxidant activity (mg AEAC/g) 1.61 ± 0.04
Total phenol (mg GAE/g) 27.37 ± 0.10
AICE inhibition (%) 36.46 ± 0.82

incubation for 20 min at 37 ◦ C. The reaction was stopped by the addition


of 3.0 mL of 1.0 N HCl (Guo, Pan, & Tanokura, 2009). The hippuric acid
released due to the action of ACE was extracted using 2.4 mL ethyl ac­
etate, and vortexed for two minutes (Sitanggang, Lesmana, et al., 2020).
A 1.0-mL hippuric acid containing solution was withdrawn from the
upper layer, and evaporated at 120 ◦ C for 30 min. The remaining sedi­
ment was then dissolved in 1.0 mL of distilled water, and the absorbance
was measured at a wavelength of 228 nm. The AICE inhibition was
evaluated as follows (Eq. 4):
Fig. 2. SDS-PAGE analysis of (A) defatted soybean meal, (B) commercial SPI,
(a − b) − (c − d)
Inhibitory act.(%) = × 100% (4) (C) defatted tempe flour, and (D) tempe flour-rich in protein.
(a − b)

where a = absorbance of solution containing ACE without sample, b = addition to this, the total phenolic content was up to 27.37 mg GAE/g.
absorbance of solution containing ACE that has been inactivated with The analysis of SDS-PAGE revealed that soybean fermentation increases
HCl (without sample), c = absorbance of solution containing ACE and the protein degree of hydrolysis (Fig. 2). The distribution of low mo­
sample, and d = absorbance of solution containing ACE and sample that lecular weight peptides (i.e., <17 kDa) was higher in defatted tempe
has been inactivated with HCl. flour (lane C) and tempe-flour rich in protein (lane D) rather than in
defatted soybean meal (lane A) and commercial SPI (lane B). The protein
bands (electrophoretograms) of higher molecular weight peptides were
2.12. The calculation of papain (protein) charge at different physiological
also thinner in tempe flour-rich in protein and defatted tempe flour
pHs
compared to commercial soy isolate protein. Additionally, the higher
amount of free amino acids might present in both tempe derivatives that
The distribution of papain (protein) charge at different physiological
passed the SDS gel (Zinchenko, Muranova, Melanyina, & Miroshnikov,
pHs was performed using a software provided at http://protcalc.source
2019). The pooled extracellular proteases were considered to perform
forge.net (The Scripps Research Institute, USA). The complete amino
proteolysis on soybean proteins leading to a higher degree of hydrolysis
acid sequential composition was collected from Basic Local Alignment
(de Reu et al., 1995; Zhang et al., 2014)
Search Tool (BLAST) with P00784 entry (The Universal Protein
Resource, UniProt - www.uniprot.org) (Paul, Therrien, & Furrer, 2015;
Sitanggang et al., 2016). 3.2. Rejection of papain by UF membrane

2.13. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS- The initial activity of papain (2.5% w/v) in 0.01 M phosphate buffer
PAGE) analysis (pH 6.8) was about 79,265 U/g. There was small papain’s activity
recovered in permeate. These activities were approximately 0.2–0.3 U/g
A 50-mg of the sample was dissolved in Tris-HCl (pH 8.4). The so­ for both MWCOs (i.e., 5- and 10-kDa). Herein, the rejections of papain
lution was incubated at 80 ◦ C for one hour and subject to centrifugation by 5- and 10-kDa PES membrane were identical, more than 99% (see
at 16,000 xg for 25 min at 25 ◦ C. The final residue was dissolved with Fig. 3a). A complete rejection of papain by 5-kDa PES membrane and its
distilled water to obtain a concentration of 4.0 mg/mL (Sitanggang, utilization for the production of sesame protein hydrolysate has also
Lesmana, et al., 2020). To run the gel separation, the detailed proced­ been reported elsewhere (Das, Bhattacherjee, & Ghosh, 2009). As
ures were adopted from literature published elsewhere (Sitanggang, common to the constant flux filtration, to compensate membrane
Alexander, et al., 2020). The electrophoresis apparatus (Bio-Rad, fouling, TMPs increased over time during papain filtrations using both
Singapore) was used according to the manufacturer’s guide and the MWCOs. In addition to this, since papain has a molecular weight of
analysis was started by giving a constant voltage of 70 V for approxi­ approximately 23 kDa (Muharram & Abdel-Kader, 2017), a higher
mately 3 h. fouling propensity with a smaller membrane cut-off was evident. TMP
by 5-kDa filtration was still higher than 10-kDa regardless of the dif­
2.14. Statistical analysis ference in set flux value (JSV). Herein, 10-kDa PES membrane was
selected for further investigations.
The evaluation of the results statistically was done using SPSS Soft­
ware version 23 (IBM, US). The analysis of variance (ANOVA) was 3.3. Robustness of the control system
performed with a confidence level of 95%.
The PID controller introduced in the reactor system was tuned
3. Results and discussion following Kuhn (1995) under the normal setting (Sitanggang et al.,
2016). The controller gain K was about 0.036, while the integral TI and
3.1. Characteristics of tempe flour-rich in protein derivative time TD were 0.468 and 0.118 min, respectively. This setting
was tested to filter the substrate solution (1% w/v) at varying fluxes (JSV
The characteristics of tempe flour-rich in protein can be seen in = 10 ➔ 25 ➔ 15 L/m2.h) using 10-kDa PES membrane. As seen in Fig. 4,
Table 1. The protein content was 51.60% (db). The antioxidant activity although an overshoot was found in process variable flux JPV as an initial
and AICE inhibition were 1.61 mg AEAC/g and 36.46%, respectively. In response of the controller, the control error could be minimized over

4
A.B. Sitanggang et al. Innovative Food Science and Emerging Technologies 68 (2021) 102639

Fig. 3. (a) Rejection of papain by 5- and 10-kDa PES membrane, and (b) filtration behavior of papain (2.5% w/v) at constant fluxes, thus residence times (τ for 5-kDa
membrane = 3 h and JSV = 22.73 L/m2.h, τ for 10-kDa membrane = 2 h and JSV = 34.09 L/m2.h).

time by less than 5%. In addition to this, the controller output (i.e., TMP)
was not jerky. A similar investigation has been reported elsewhere,
where the open-loop tuning was utilized to tune the PID setting, and also
resulted in a control error of less than 5% (Sitanggang, Drews, &
Kraume, 2014b). Conclusively, the developed PID controller in this
study could be used to facilitate a constant flux operation during the
continuous hydrolysis of tempe flour-rich in protein for producing
bioactive peptides.

3.4. Influence of enzyme-to-substrate ratio for producing bioactive


peptides

The influence of enzyme-to-substrate ratio [E]/[S] was investigated


at three different levels (i.e., 5, 10, and 15%). The total phenol and
protein concentration insignificantly changed in every hour of sampling
for the three levels of [E]/[S]. The antioxidant activity and AICE inhi­
bition of permeate at 10% [E]/[S] were slightly higher than that of 5 and
Fig. 4. Robustness of PID controller during the filtration of tempe flour-rich in 15% (see Fig. 5).
protein (15% w/v) at varying fluxes (JSV: 10 ➔ 25 ➔ 15 L/m2.h). PID setting: K The total phenols, protein concentrations, antioxidant activities, and
= 0.036, TI = 0.468 min and TD= 0.118 min, 10-kDa PES membrane. AICE inhibitions between the initial substrate and cumulative permeate
are shown in Table 2. The total phenol and protein concentration of the

Fig. 5. The influence of enzyme-to-substrate ratio [E]/[S] on total phenol, protein content, antioxidant activity and AICE inhibition during continuous hydrolysis of
tempe flour-rich in protein. Reacting conditions: [S] = 1% (w/v), τ = 5 h (JSV = 13.64 L/m2.h), pH 6.8, N = 300 rpm, T = 40 ◦ C, 10-kDa PES membrane.

5
A.B. Sitanggang et al. Innovative Food Science and Emerging Technologies 68 (2021) 102639

Table 2 substrate were significantly higher than that of cumulative permeates at


Phenol, protein, antioxidant capacity, and AICE inhibition between initial re­ a confidence level of 95%. The rejection of soybean isoflavones from soy
action and cumulative permeate taken in the container. Reacting conditions for residue by 20-kDa membrane by more than 40% has been reported (Li,
different treatments given in Fig. 5 and Fig. 7. Wu, Dong, & Li, 2012). Within this study, with a 10-kDa membrane,
Treatment Phenol Protein Antioxidant AICE such phenol rejection was inevitable to yield smaller total phenol in
(mg GAE/ (mg/mL) (mg AEAC/ inhibition permeate compared to the substrate. From Fig. 2, the electrophoreto­
mL) mL) (%)
gram of tempe flour-rich in protein still had parent proteins with mo­
Substrate 0.151 ± 0.161 ± 0.028 ± 36.61 ± lecular weights of more than 17 kDa. The incomplete proteolysis on the
0.001d 0.001c 0.002b 0.26a parent proteins (i.e., <100% of protein degree of hydrolysis) might lead
Permeate, 5% 0.133 ± 0.114 ± 0.026 ± 43.04 ±
[E]/[S] 0.002a 0.004b 0.001a 1.71b
to the protein rejection by 10-kDa membrane (Sun, 2011). Thus, the
10% 0.136 ± 0.115 ± 0.030 ± 46.02 ± protein concentrations in the three cumulative permeates (i.e., [E]/[S]:
0.002b 0.003b 0.001b 1.10c 5, 10, and 15%) were lower than that of the substrate.
15% 0.141 ± 0.108 ± 0.027 ± 43.25 ± The trend of antioxidant activities between substrate and the three
0.001c 0.002a 0.001c 2.40b
cumulative permeates were different from that of total phenol and
Substrate 0.153 ± 0.161 ± 0.027 ± 36.60 ±
0.002d 0.001c 0.001a 0.16a protein content. The compounds acting as antioxidants were presumably
Permeate, τ 5h 0.136 ± 0.115 ± 0.030 ± 46.02 ± not limited from the isoflavones as the hydrolysis of soy protein also
0.002a 0.003a 0.001b 1.10b produced bioactive peptides exhibiting antioxidant activity (Park, Lee,
7h 0.140 ± 0.126 ± 0.032 ± 48.43 ± Baek, & Lee, 2010). Amongst the three permeates, increasing [E]/[S] to
0.002b 0.002b 0.001c 0.86c
15% was detrimental for antioxidant activity. It could be due to exces­
9h 0.142 ± 0.128 ± 0.033 ± 50.88 ±
0.002b 0.002b 0.001c 0.95d sive hydrolysis that led to the breakdown of peptides displaying anti­
oxidant activity (Xu, Shen, Chen, Bean, & Li, 2019; Zhao, Zhao, Ahn, Jin,
Note: Different superscript letters in a column show significant difference with a
& Huang, 2019). The substrate AICE inhibition was significantly lower
confidence level of 95%.
compared to the three cumulative permeates (P < 0.05). This indicated
that hydrolysis of parent proteins by papain could release bioactive

Fig. 6. Profiles of ∆P, JSV and JPV during continuous hydrolysis of tempe flour-rich in protein at different levels of [E]/[S]: (a) 5%, (b) 10%, and (c) 15%.

Fig. 7. The influence of residence time τ on total phenol, protein content, antioxidant activity and AICE inhibition during continuous hydrolysis of tempe flour-rich
in protein. Reacting conditions: [S] = 1% (w/v), [E]/[S] = 10%, pH 6.8, N = 300 rpm, T = 40 ◦ C, 10-kDa PES membrane.

6
A.B. Sitanggang et al. Innovative Food Science and Emerging Technologies 68 (2021) 102639

Fig. 8. Profiles of ∆P, JSV and JPV during continuous hydrolysis of tempe flour-rich in protein at different levels of τ: (a) 5 h (JSV = 13.64 L/m2.h), (b) 7 h (JSV = 9.74
L/m2.h), and (c) 9 h (JSV = 7.58 L/m2.h).

peptides exhibiting both antioxidant capacity and AICE inhibitory ac­ The total phenol and protein concentration in the substrate were
tivity (Lo & Li-Chan, 2005; Park et al., 2010). Amongst the three cu­ higher significantly than that of the three permeates (i.e., treated with
mulative permeates, the highest total phenol, protein concentration, different residence times) with a confidence level of 95%. Amongst
antioxidant activity, and AICE inhibition were obtained by 10% [E]/[S] permeates, longer residence time obtained higher total phenol and
treatment. Chiang, Tsou, Tsai, and Tsai (2006) reported an optimum protein content. Higher protein content at τ = 9 h (i.e., 0.128 mg/mL)
[E]/[S] of 10% during the hydrolysis of soy protein isolate (SPI) using was due to a longer time for papain to break down the peptide bonds in
alcalase for producing peptides exhibiting AICE inhibitory activity. tempe parent proteins. These small molecular weight peptides and free
Other studies also used [E]/[S] by 2 to 10% for hydrolyzing soybean amino acids passed through 10-kDa membrane, thus yielded in a higher
protein using different enzymes to obtain AICE inhibitory peptides protein concentration in permeate (Wei & Chiang, 2009). As mentioned
(Chen, Okada, Muramoto, Suetsuna, & Yang, 2002; Gouda, Gowda, Rao, previously, the compounds acting as antioxidants were presumably not
& Prakash, 2006; Li, Xia, Zhang, & Li, 2018; Wu & Ding, 2001). limited from the isoflavones. The hydrolysis of tempe protein might also
Fig. 6 shows the influence of [E]/[S] on TMP during the continuous produce bioactive peptides exhibiting antioxidant activity. The antiox­
hydrolysis of tempe flour-rich in protein. A higher [E]/[S] given in the idant activities for the three permeates were significantly higher than
reacting solution caused higher increase in pressure. The UF membrane that of the substrate (P < 0.05). The highest antioxidant capacity was
made of PES (Microdyn-Nadir GmbH, Germany) is negatively charged also found at τ = 9 h, approximately 0.032 mg AEAC/mL. As a cysteine
(de la Torre, Harff, Lesjean, Drews, & Kraume, 2009; Puro et al., 2010; protease-endopeptidase, papain has been reported to cleave at (hydro­
Suárez, Díez, García, & Riera, 2012). The calculated charge of papain at phobic amino acid)-(R/K/E/H/G/Y) in protein structures (Fernández-
pH 6.8 (0.01 M phosphate buffer) was approximately +1. The electro­ Lucas, Castañeda, & Hormigo, 2017; Nath et al., 2020). The release of
static interaction between papain molecules and PES membrane surface low molecular peptides containing a high portion of hydrophobic amino
was toward adhesion, leading to the adsorption of protein molecules acids (Q, A, V, and L), aromatic amino acids (Y, H, W, and F), and
onto the membrane surface (Sitanggang et al., 2016). Therefore, a cysteine (C) has been reported to exhibit high antioxidant activity
higher fouling rate was pronounced at [E]/[S] of 15%, thus having a (Sarmadi & Ismail, 2010; Zou et al., 2016). Similar to antioxidant ca­
higher total resistance at the end of the filtration. pacity, the AICE inhibitions for the three permeates were also higher
than the substrate. The highest inhibition of 50.9% was obtained at τ =
9 h. The production of peptides having hydrophobic amino acids (W, F,
3.5. Influence of residence time for producing bioactive peptides
A, I, and V) at the C-terminal position and positively charged amino
acids (K, R) is considered to display higher AICE inhibitory activity
Fig. 7 shows the influence of three residence times (τ = 5, 7, and 9 h)
(Sanjukta & Rai, 2016; Sarmadi & Ismail, 2010). The AICE inhibition of
on total phenol, protein content, antioxidant activity, and AICE inhibi­
50.9% was the highest one found in this study by varying two operating
tion of protein hydrolysates. The total phenol, protein content, and
conditions ([E]/[S] and τ). In our previous study, the combination of
antioxidant capacity of permeate were relatively constant during 5 h
papain hydrolysis under batch operation followed by 10-kDa membrane
reaction, in which τ of 9 h prevailed as compared to the other treat­
filtration could obtain permeate having AICE inhibition up to 57.3%
ments. For AICE inhibition, the fluctuations were observed, especially
(Sitanggang, Lesmana, et al., 2020). The batch hydrolysis was per­
when the inhibition peaked at 4 h of reaction. Similar to the other three
formed under [E]/[S] = 3.3% (w/v), [E] = 100,000 U/g solid, [S] = 10%
observed parameters, the longer residence time could obtain permeate
(w/v) and reaction time t = 4 h. The notable difference between our
with a higher AICE inhibition.

0.8 0.3
(a) (b)
Antioxidant IC50 (mg/mL)

d
c
AICE IC50 (mg/mL)

0.6
0.2

b
0.4 c
a a a b b
0.1 a
0.2

0.0 0.0
Substrate10 kDa 5 kDa 4 kDa 2 kDa Substrate10 kDa 5 kDa 4 kDa 2 kDa

Fig. 9. The half-maximal inhibitory concentration (IC50) of the substrate and resulted permeates (with 10-, 5-, 4-, and 2-kDa membrane cut-off): (a) antioxidant and
(b) AICE. Different letters indicate a significant difference with a confidence level of 95%.

7
A.B. Sitanggang et al. Innovative Food Science and Emerging Technologies 68 (2021) 102639

Table 3 IC50 for AICE in the substrate was significantly higher than that of
IC50 values of AICE from soy-based bioactive peptides prepared with different hydrolysates filtered by different membrane cut-offs with a confidence
methods. level of 95%. The lowest cut-off (2-kDa membrane) obtained better ac­
No. Type of Reacting conditions IC50 (mg Ref. tivity to inhibit AICE (IC50 = 0.08 mg protein/mL). Permeate resulted
soybean meal protein/ from filtration using 2-kDa membrane had IC50 ~ 40% smaller than that
mL) of 10-kDa membrane. Other studies also demonstrated the increase in
1 Soybean flour Batch-wise, pepsin, 37 ◦ C, 0.01a Chen et al., AICE inhibition using a lower membrane cut-off (Chiang et al., 2006;
24 h, pH 2, [E]/[S] = 1% 2002 Sitanggang, Lesmana, et al., 2020; Yao et al., 2019). Table 3 shows soy-
2 Batch-wise, M. purpureus 0.126 Kuba, Tana,
β-conglycinin
based bioactive peptides IC50 values for AICE prepared with different
3 Glycinin acid proteinase, 37 ◦ C, 10 0.148 Tawata, &
h, pH 3.3, [E]/[S] = 100 Yasuda, methods. The types of soybean meal utilized were soybean flour, soy­
U/g 2005 bean storage protein (β-conglycinin and glycinin), and SPI. The reported
4 Glycinin Batch-wise, Aspergillus sp. 0.001b Gouda et al., IC50 values were in a range of 0.001–0.28 mg protein/mL. The AICE IC50
Amano-P (Protease P), 2006 of tempe hydrolysate filtered with 2-kDa membrane, in this study, was
37 ◦ C, 18 h, pH 8.2, [E]/
[S] = 2%
comparable with the reported IC50 values found in the literature.
5 SPI Batch-wise, pepsin, 37 ◦ C, 0.28 Lo & Li- Furthermore, most of the reported studies were performed under batch
1 h, pH 2.0, [E]/[S] = 4% Chan, 2005 operation, except that of Chiang et al. (2006). Higher productivity for an
continued with enzyme-catalyzed reaction is always one of the considerations for
hydrolysis by pancreatin,
selecting continuous operation over batch one (Chiang et al., 2006;
37 ◦ C, 2 h, pH 7.5
6 SPI Batch-wise, E.coli JM109 0.018 Kodera & Sitanggang et al., 2014b).
protease D3, 37–40 ◦ C, Nio, 2006
24 h, pH 4.5, [E]/[S] = 4. Conclusions
0.2%
7 SPI Batch-wise, alcalase, 0.078 Chiang et al.,
50 ◦ C, 24 h, pH 9, [E]/[S] 2006 Within this study, a combination of fermentation and in vitro pro­
= 2.5%, 10-kDa PES teolysis was realized to produce soy-based bioactive peptides exhibiting
membrane antioxidant acitivity and AICE inhibitory activity. The operating con­
8 SPI Continuous, alcalase, 0.09
ditions, such as [E]/[S] = 10% (w/v), and τ = 9 h were selected as the
50 ◦ C, 8 h, pH 9, [E]/[S]
= 2.5%, τ = n.a (JSV = n. optimum operating conditions for hydrolyzing tempe flour-rich in pro­
a) 10-kDa PES membrane tein using membrane reactor. With a 10-kDa membrane filtration, the
9 Tempe-flour Continuous, papain, 0.08 This study resulted antioxidant capacity and AICE inhibition were 0.033 mg AEAC/
rich in protein 40 ◦ C, pH 6.8, [E]/[S] = mL and 50.9%, respectively. Further permeate filtration with a smaller
10%, τ = 9 h (JSV = 7.58
membrane cut-off (i.e., 2-kDa membrane) reduced the antioxidant ca­
L/m2.h), N = 300 rpm, 2-
kDa PES membrane pacity. In contrast, with a 2-kDa membrane cut-off, AICE inhibition of
permeate increased which was indicated by a reduction of IC50
Note: “a” and “b” were calculated based on molecular weight of Y-L-A-G-N-Q
approximately 40%. Conclusively, in vitro hydrolysis of tempe to pro­
and V-L-I-V-P, respectively; n.a = not available.
duce functional ingredients such as bioactive peptides is amenable. This
is expected to increase the commercialization of tempe-based food
current study with the previous one (Sitanggang, Lesmana, et al., 2020) products.
was that the current substrate concentration was one-tenth of the pre­
vious study.
Fig. 8 depicts the influence of τ on TMP. In general, a shorter resi­ Declaration of Competing Interest
dence time, thus a higher convective transport, causes a higher rate of
fouling. The TMP almost reached 2.0 bar during the hydrolysis at τ = 5 h The authors declare that the research was conducted without a po­
(JSV = 13.64 L/m2.h). When the flux reduced down to 7.58 L/m2.h tential conflict of interest.
(equivalent to τ = 9 h), the final TMP was below 1.0 bar. With such
increase in TMP within 5 h of reaction (i.e., TMP < 1.0 bar), it is still
possible to perform longer hydrolysis of tempe flour-rich in protein since Acknowledgement
the pressure stability of the reactor system is up to 6.0 bar.
The authors gratefully acknowledge the Ministry of Research and
3.6. IC50 of antioxidant and AICE with different membrane cut-offs Technology/National Agency for Research and Innovation, Indonesia,
for supporting this research through the schemes of Penelitian Dasar
The protein hydrolysate produced by employing the optimum Unggulan Perguruan Tinggi (PDUPT) 4053/IT3.L1/PN/2020 and
operating conditions (i.e., [E]/[S] = 10% and τ = 9 h, 10-kDa mem­ (partially) WCU Program managed by Institut Teknologi Bandung. We
brane) was subject to further separations using 5-, 4-, and 2-kDa PES also thank Dr. Simson Tarigan (Indonesian Research Center for Veteri­
membrane. IC50 values for the antioxidant and AICE of the corre­ nary Sciences) and Ir. Oentojo Tjiptobusono for their technical support.
sponding filtrates were measured and presented in Fig. 9. IC50 for
antioxidant activity in the substrate was higher significantly compared References
to permeate with 10-, 5-, 4-, and 2-kDa membrane filtration (P < 0.05,
Agyei, D. (2015). Bioactive proteins and peptides from soybeans. Recent Patents on Food,
Fig. 9a). The antioxidant IC50 values for permeates resulted from 10-, 5-, Nutrition & Agriculture, 7(2), 100–107. https://doi.org/10.2174/
and 4-kDa membrane filtration were not significantly different at 95% 2212798407666150629134141
confidence level, approximately 0.25 mg protein/mL. The rejection of Ajibola, C. F., Fashakin, J. B., Fagbemi, T. N., & Aluko, R. E. (2011). Effect of peptide size
on antioxidant properties of African yam bean seed (Sphenostylis stenocarpa) protein
isoflavones might contribute to the reduction of antioxidant capacity at
hydrolysate fractions. International Journal of Molecular Sciences, 12(10), 6685–6702.
2-kDa membrane filtration (Li et al., 2012). The excessive breakdown of https://doi.org/10.3390/ijms12106685
peptides by the enzyme, leading to the release of smaller molecular Astuti, M., Meliala, A., Dalais, F. S., & Wahlqvist, M. L. (2000). Tempe, a nutritious and
weight peptides and free amino acids having a less antioxidative prop­ healthy food from Indonesia. Asia Pacific Journal of Clinical Nutrition, 9(4), 322–325.
https://doi.org/10.1046/j.1440-6047.2000.00176.x
erty, could also relate to the reduced antioxidant capacity of the filtrate Banan-Mwine Daliri, E., Oh, D. H., & Lee, B. H. (2017). Bioactive peptides - review.
(Xu et al., 2019). Foods, 6(32), 1–21. https://doi.org/10.3390/foods6050032

8
A.B. Sitanggang et al. Innovative Food Science and Emerging Technologies 68 (2021) 102639

Brand-Williams, W., Cuvelier, M. E., & Berset, C. (1995). Use of a free radical method to Park, S. Y., Lee, J. S., Baek, H. H., & Lee, H. G. (2010). Purification and characterization
evaluate antioxidant activity. LWT - Food Science and Technology, 28(1), 25–30. of antioxidant peptides from soy protein hydrolysate. Journal of Food Biochemistry,
https://doi.org/10.1016/S0023-6438(95)80008-5 34(Suppl. 1), 120–132. https://doi.org/10.1111/j.1745-4514.2009.00313.x
Cao, Z.-H., Green-Johnson, J. M., Buckley, N. D., & Lin, Q.-Y. (2019). Bioactivity of soy- Paul, L. E. H., Therrien, B., & Furrer, J. (2015). Interactions of arene ruthenium
based fermented foods: A review. Biotechnology Advances, 37(1), 223–238. https:// metallaprisms with human proteins. Organic & Biomolecular Chemistry, 13(3),
doi.org/10.1016/j.biotechadv.2018.12.001 946–953. https://doi.org/10.1039/C4OB02194K
Chalamaiah, M., Keskin Ulug, S., Hong, H., & Wu, J. (2019). Regulatory requirements of Puro, L., Kallioinen, M., Mänttäri, M., Natarajan, G., Cameron, D. C., & Nyström, M.
bioactive peptides (protein hydrolysates) from food proteins. Journal of Functional (2010). Performance of RC and PES ultrafiltration membranes in filtration of pulp
Foods, 58(June), 123–129. https://doi.org/10.1016/j.jff.2019.04.050 mill process waters. Desalination, 264(3), 249–255. https://doi.org/10.1016/j.
Chen, J. R., Okada, T., Muramoto, K., Suetsuna, K., & Yang, S. C. (2002). Identification of desal.2010.06.034
angiotensin I-converting enzyme inhibitory peptides derived from the peptic digest de Reu, J. C., ten Wolde, R. M., de Groot, J., Nout, M. J. R., Rombouts, F. M., &
of soybean protein. Journal of Food Biochemistry, 26(6), 543–554. https://doi.org/ Gruppen, H. (1995). Protein hydrolysis during soybean tempe fermentation with
10.1111/j.1745-4514.2002.tb00772.x Rhizopus oligosporus. Journal of Agricultural and Food Chemistry, 43(8), 2235–2239.
Chiang, W. D., Tsou, M. J., Tsai, Z. Y., & Tsai, T. C. (2006). Angiotensin I-converting https://doi.org/10.1021/jf00056a050
enzyme inhibitor derived from soy protein hydrolysate and produced by using Rho, S. J., Lee, J. S., Chung, Y. I., Kim, Y. W., & Lee, H. G. (2009). Purification and
membrane reactor. Food Chemistry, 98(4), 725–732. https://doi.org/10.1016/j. identification of an angiotensin I-converting enzyme inhibitory peptide from
foodchem.2005.06.038 fermented soybean extract. Process Biochemistry, 44(4), 490–493. https://doi.org/
Cupp-Enyard, C. (2008). Sigma’s non-specific protease activity assay - casein as a 10.1016/j.procbio.2008.12.017
substrate. Journal of Visualized Experiments, 19, 4–5. https://doi.org/10.3791/899 Rosset, M., Acquaro, V. R., & Beléia, A. D. P. (2014). Protein extraction from defatted
Das, R., Bhattacherjee, C., & Ghosh, S. (2009). Effects of operating parameters and nature soybean flour with Viscozyme L pretreatment. Journal of Food Processing and
of fouling behavior in ultrafiltration of sesame protein hydrolysate. Desalination, 237 Preservation, 38(3), 784–790. https://doi.org/10.1111/jfpp.12030
(1–3), 268–276. https://doi.org/10.1016/j.desal.2008.01.020 Sanjukta, S., & Rai, A. K. (2016). Production of bioactive peptides during soybean
Fernández-Lucas, J., Castañeda, D., & Hormigo, D. (2017). New trends for a classical fermentation and their potential health benefits. Trends in Food Science and
enzyme: Papain, a biotechnological success story in the food industry. Trends in Food Technology, 50, 1–10. https://doi.org/10.1016/j.tifs.2016.01.010
Science & Technology, 68, 91–101. https://doi.org/10.1016/j.tifs.2017.08.017 Sarmadi, B. H., & Ismail, A. (2010). Antioxidative peptides from food proteins: A review.
Gibbs, B. F., Zougman, A., Masse, R., & Mulligan, C. (2004). Production and Peptides, 31(10), 1949–1956. https://doi.org/10.1016/j.peptides.2010.06.020
characterization of bioactive peptides from soy hydrolysate and soy-fermented food. Singh, B. P., Vij, S., & Hati, S. (2014). Functional significance of bioactive peptides
Food Research International, 37(2), 123–131. https://doi.org/10.1016/j. derived from soybean. Peptides, 54, 171–179. https://doi.org/10.1016/j.
foodres.2003.09.010 peptides.2014.01.022
Gouda, K. G. M., Gowda, L. R., Rao, A. G. A., & Prakash, V. (2006). Angiotensin I- Sitanggang, A. B., Alexander, R., & Budijanto, S. (2020). The utilization of bilimbi
converting enzyme inhibitory peptide derived from glycinin, the 11S globulin of (Averrhoa bilimbi) and lime (Citrus aurantifolia) juices as natural acid coagulants for
soybean (Glycine max). Journal of Agricultural and Food Chemistry, 54(13), tofu production. Journal of Food Science and Technology. https://doi.org/10.1007/
4568–4573. https://doi.org/10.1021/jf060264q s13197-020-04503-5
Guo, Y., Pan, D., & Tanokura, M. (2009). Optimisation of hydrolysis conditions for the Sitanggang, A. B., Drews, A., & Kraume, M. (2014a). Rapid transgalactosylation towards
production of the angiotensin-I converting enzyme (ACE) inhibitory peptides from lactulose synthesis in a small-scale enzymatic membrane reactor (EMR). Chemical
whey protein using response surface methodology. Food Chemistry, 114(1), 328–333. Engineering Transactions, 38(JUNE), 19–24. https://doi.org/10.3303/CET1438004
https://doi.org/10.1016/j.foodchem.2008.09.041 Sitanggang, A. B., Drews, A., & Kraume, M. (2014b). Continuous synthesis of lactulose in
Kodera, T., & Nio, N. (2006). Identification of an angiotensin I-converting enzyme an enzymatic membrane reactor reduces lactulose secondary hydrolysis. Bioresource
inhibitory peptides from protein hydrolysates by a soybean protease and the Technology. https://doi.org/10.1016/j.biortech.2014.05.124
antihypertensive effects of hydrolysates in 4 spontaneously hypertensive model rats. Sitanggang, A. B., Drews, A., & Kraume, M. (2015). Influences of operating conditions on
Journal of Food Science, 71(3), C164–C173. https://doi.org/10.1111/j.1365- continuous lactulose synthesis in an enzymatic membrane reactor system: A basis
2621.2006.tb15612.x prior to long-term operation. Journal of Biotechnology, 203, 89–96. https://doi.org/
Korhonen, H., & Pihlanto, A. (2006). Bioactive peptides: Production and functionality. 10.1016/j.jbiotec.2015.03.016
International Dairy Journal, 16(9), 945–960. https://doi.org/10.1016/j. Sitanggang, A. B., Drews, A., & Kraume, M. (2016). Development of a continuous
idairyj.2005.10.012 membrane reactor process for enzyme-catalyzed lactulose synthesis. Biochemical
Kuba, M., Tana, C., Tawata, S., & Yasuda, M. (2005). Production of angiotensin I- Engineering Journal, 109, 65–80. https://doi.org/10.1016/j.bej.2016.01.006
converting enzyme inhibitory peptides from soybean protein with Monascus Sitanggang, A. B., Lesmana, M., & Budijanto, S. (2020). Membrane-based preparative
purpureus acid proteinase. Process Biochemistry, 40(6), 2191–2196. https://doi.org/ methods and bioactivities mapping of Tempe-based peptides. Food Chemistry, 329,
10.1016/j.procbio.2004.08.010 127193. https://doi.org/10.1016/j.foodchem.2020.127193
Kuhn, U. (1995). Eine praxisnahe Einstellregel fuer PID-Regler: Die T-Summen-Regel. Sitanggang, A. B., Sinaga, W. S. L., Wie, F., Fernando, F., & Krusong, W. (2020).
Automatisierungstechnische Praxis, 37(5), 10–17. Enhanced antioxidant activity of okara through solid state fermentation of GRAS
Li, B., Wu, Z. Q., Dong, L. Y., & Li, L. (2012). Purification of soybean isoflavones from soy fungi. Food Science and Technology, 40(1), 178–186. https://doi.org/10.1590/
sauce residue by ultrafiltration. Advanced Materials Research, 581–582(1), fst.37218
1184–1188. https://doi.org/10.4028/www.scientific.net/AMR.581-582.1184 Sonklin, C., Laohakunjit, N., & Kerdchoechuen, O. (2018). Assessment of antioxidant
Li, M., Xia, S., Zhang, Y., & Li, X. (2018). Optimization of ACE inhibitory peptides from properties of membrane ultrafiltration peptides from mungbean meal protein
black soybean by microwave-assisted enzymatic method and study on its stability. hydrolysates. PeerJ, 2018(7). https://doi.org/10.7717/peerj.5337
Lwt, 98, 358–365. https://doi.org/10.1016/j.lwt.2018.08.045 Suárez, L., Díez, M.a., García, R., & Riera, F.a. (2012). Membrane technology for the
Liu, C., Wang, X., Ma, H., Zhang, Z., Gao, W., & Xiao, L. (2008). Functional properties of recovery of detergent compounds: A review. Journal of Industrial and Engineering
protein isolates from soybeans stored under various conditions. Food Chemistry, 111 Chemistry, 18(6), 1859–1873. https://doi.org/10.1016/j.jiec.2012.05.015
(1), 29–37. https://doi.org/10.1016/j.foodchem.2008.03.040 Sun, X. D. (2011). Enzymatic hydrolysis of soy proteins and the hydrolysates utilisation.
Liu, T.-X., & Zhao, M. (2010). Physical and chemical modification of SPI as a potential International Journal of Food Science and Technology, 46(12), 2447–2459. https://doi.
means to enhance small peptide contents and antioxidant activity found in org/10.1111/j.1365-2621.2011.02785.x
hydrolysates. Innovative Food Science & Emerging Technologies, 11(4), 677–683. Syah, D., Sitanggang, A. B., Faradilla, R. F., Trisna, V., Karsono, Y., & Septianita, D. A.
https://doi.org/10.1016/j.ifset.2010.08.004 (2015). The influences of coagulation conditions and storage proteins on the textural
Lo, W. M. Y., & Li-Chan, E. C. Y. (2005). Angiotensin I converting enzyme inhibitory properties of soy-curd (tofu). CyTA Journal of Food, 13(2), 259–263. https://doi.org/
peptides from in vitro pepsin− pancreatin digestion of soy protein. Journal of 10.1080/19476337.2014.948071
Agricultural and Food Chemistry, 53(9), 3369–3376. https://doi.org/10.1021/ Tanzadehpanah, H., Asoodeh, A., Saberi, M. R., & Chamani, J. (2013). Identification of a
jf048174d novel angiotensin-I converting enzyme inhibitory peptide from ostrich egg white and
Lowry, O. H., N, J. R., Farr, L. A., & Randall, R. J. (1951). Protein measurement with the studying its interactions with the enzyme. Innovative Food Science & Emerging
Folin phenol reagent. Readings, 193(1), 265–275. https://doi.org/10.1016/0304- Technologies, 18, 212–219. https://doi.org/10.1016/j.ifset.2013.02.002
3894(92)87011-4 de la Torre, T., Harff, M., Lesjean, B., Drews, A., & Kraume, M. (2009). Characterisation
Mani, V., & Ming, L. C. (2017). Tempeh and other fermented soybean products rich in of polysaccharide fouling of an ultrafiltration membrane using model solutions.
isoflavones. Fermented Foods in Health and Disease Prevention, 453–474. Elsevier Desalination and Water Treatment, 8(1–3), 17–23. https://doi.org/10.5004/
https://doi.org/10.1016/B978-0-12-802309-9.00019-4. dwt.2009.685
Muharram, M. M., & Abdel-Kader, M. S. (2017). Utilization of gel electrophoreses for the Ubilla, C., Ramírez, N., Valdivia, F., Vera, C., Illanes, A., & Guerrero, C. (2020). Synthesis
quantitative estimation of digestive enzyme papain. Saudi Pharmaceutical Journal, 25 of lactulose in continuous stirred tank reactor with β-galactosidase of Apergillus
(3), 359–364. https://doi.org/10.1016/j.jsps.2016.09.002 oryzae immobilized in monofunctional glyoxyl agarose support. Frontiers in
Nath, A., Kailo, G. G., Mednyánszky, Z., Kiskó, G., Csehi, B., Pásztorné-Huszár, K., … Bioengineering and Biotechnology, 8. https://doi.org/10.3389/fbioe.2020.00699
Vatai, G. (2020). Antioxidant and antibacterial peptides from soybean milk through Wei, J. T., & Chiang, B. H. (2009). Bioactive peptide production by hydrolysis of porcine
enzymatic-and membrane-based technologies. Bioengineering, 7(1), 1–18. https:// blood proteins in a continuous enzymatic membrane reactor. Journal of the Science of
doi.org/10.3390/bioengineering7010005 Food and Agriculture, 89(3), 372–378. https://doi.org/10.1002/jsfa.3451
Nout, M. J. R., & Kiers, J. L. (2005). Tempe fermentation, innovation and functionality: Wu, J., & Ding, X. (2001). Hypotensive and physiological effect of angiotensin converting
Update into the third millenium. Journal of Applied Microbiology, 98(4), 789–805. enzyme inhibitory peptides derived from soy protein on spontaneously hypertensive
https://doi.org/10.1111/j.1365-2672.2004.02471.x rats. Journal of Agricultural and Food Chemistry, 49(1), 501–506. https://doi.org/
10.1021/jf000695n

9
A.B. Sitanggang et al. Innovative Food Science and Emerging Technologies 68 (2021) 102639

Xu, S., Shen, Y., Chen, G., Bean, S., & Li, Y. (2019). Antioxidant characteristics and Zhao, Q. C., Zhao, J. Y., Ahn, D. U., Jin, Y. G., & Huang, X. (2019). Separation and
identification of peptides from sorghum Kafirin hydrolysates. Journal of Food Science, identification of highly efficient antioxidant peptides from eggshell membrane.
84(8), 2065–2076. https://doi.org/10.1111/1750-3841.14704 Antioxidants, 8(10), 1–16. https://doi.org/10.3390/antiox8100495
Yao, G. L., He, W., Wu, Y. G., Chen, J., Hu, X. W., & Yu, J. (2019). Purification of Zinchenko, D. V., Muranova, T. A., Melanyina, L. A., & Miroshnikov, A. I. (2019).
angiotensin-I-converting enzyme inhibitory peptides derived from Camellia oleifera Hydrolysis of soybean and rapeseed proteins with enzyme complex extracted from
abel seed meal hydrolysate. Journal of Food Quality, 2019. https://doi.org/10.1155/ the pyloric caeca of the cod. Applied Biochemistry and Microbiology, 55(2), 165–172.
2019/7364213 https://doi.org/10.1134/S0003683819020182
Zhang, S. T., Shi, Y., Zhang, S. L., Shang, W., Gao, X. Q., & Wang, H. K. (2014). Whole Zou, T. B., He, T. P., Li, H. B., Tang, H. W., & Xia, E. Q. (2016). The structure-activity
soybean as probiotic lactic acid bacteria carrier food in solid-state fermentation. relationship of the antioxidant peptides from natural proteins. Molecules, 21(1),
Food Control, 41(1), 1–6. https://doi.org/10.1016/j.foodcont.2013.12.026 1–14. https://doi.org/10.3390/molecules21010072

10

You might also like