Sie sind auf Seite 1von 30

FO06CH11-Aluko

ARI

14 March 2015

ANNUAL
REVIEWS

Further

12:22

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Click here for quick links to


Annual Reviews content online,
including:
Other articles in this volume
Top cited articles
Top downloaded articles
Our comprehensive search

Antihypertensive Peptides from


Food Proteins
Rotimi E. Aluko
Department of Human Nutritional Sciences, University of Manitoba, Winnipeg, Manitoba,
Canada R3T 2N2; email: rotimi.aluko@umanitoba.ca

Annu. Rev. Food Sci. Technol. 2015. 6:23562

Keywords

The Annual Review of Food Science and Technology is


online at food.annualreviews.org

antihypertensive peptides, food proteins, protein hydrolysates,


angiotensin-converting enzyme, renin, spontaneously hypertensive rats

This articles doi:


10.1146/annurev-food-022814-015520
c 2015 by Annual Reviews.
Copyright 
All rights reserved

Abstract
Bioactive peptides are encrypted within the primary structure of food
proteins where they remain inactive until released by enzymatic hydrolysis.
Once released from the parent protein, certain peptides have the ability
to modulate the renin-angiotensin system (RAS) because they decrease
activities of renin or angiotensin-converting enzyme (ACE), the two main
enzymes that regulate mammalian blood pressure. These antihypertensive
peptides can also enhance the endothelial nitric oxide synthase (eNOS)
pathway to increase nitric oxide (NO) levels within vascular walls and
promote vasodilation. The peptides can block the interactions between
angiotensin II (vasoconstrictor) and angiotensin receptors, which can
contribute to reduced blood pressure. This review focuses on the methods
that are involved in antihypertensive peptide production from food sources,
including fractionation protocols that are used to enrich bioactive peptide
content and enhance peptide activity. It also discusses mechanisms that are
believed to be involved in the antihypertensive activity of these peptides.

235

FO06CH11-Aluko

ARI

14 March 2015

12:22

1. INTRODUCTION

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Food proteins normally contain several specic peptide sequences that remain inactive as long
as they remain bonded to other amino acids within the primary structure. Treatment of proteins
with an enzyme or a combination of enzymes can lead to proteolysis that releases these peptide
sequences. The free forms of these peptides can then be used as therapeutic tools against human
chronic diseases such as hypertension. Following proteolysis, the product mixture contains a
mixture of peptides and undigested proteins. The product mixture is then centrifuged to precipitate
the heavier undigested proteins while the smaller (lighter) peptides remain in the supernatant,
which can then be freeze-dried or spray-dried into a powder. This peptide mixture is called
a protein hydrolysate and can be used as is to determine bioactive properties or can undergo
additional treatments that will separate the component peptides into different fractions. The
protein hydrolysate can be separated into various fractions on the basis of peptide size, charge, or
hydrophobicity. Active fractions can also be subjected to column chromatography-based peptide
purication to isolate homogeneous preparations that contain a single peptide (Li et al. 2014).
Protein hydrolysate, peptide fractions, and isolated homogeneous peptides can be analyzed to
determine amino acid composition or amino acid sequence (Herregods et al. 2011, Li et al. 2014),
which provides information on the structural composition of the products.
Hypertension is a chronic disease that is usually manifested as excessive high blood pressure
(systolic and diastolic values of 140 mmHg and 90 mmHg, respectively), which is characterized
by insufcient relaxation of blood vessels and reduced blood ow. If left untreated, hypertension
can lead to insufcient blood ow to vital organs, which can cause stroke and eventually death.
Human blood pressure is mainly regulated by the renin-angiotensin system (RAS), which is
maintained by two critically important proteases, renin and angiotensin-converting enzyme
(ACE). Renin (EC 3.4.23.15) is a 37-kDa enzyme that hydrolyses the Leu10-Val11 peptide
bond of angiotensinogen (a 55-kDa protein synthesized in the liver) to produce a physiologically
inactive angiotensin I, a decapeptide (Imai et al. 1983, Acharya et al. 2003). ACE (EC 3.4.15.1)
then cleaves off a dipeptide to convert angiotensin I into a vasoactive angiotensin II (an octapeptide) that binds with receptors on the vascular wall to cause blood vessel contractions. In disease
conditions, there is increased activity of renin and/or ACE, which causes abnormally high blood
levels of angiotensin II and leads to excessive blood vessel contraction but insufcient relaxation.
ACE also cleaves and inactivates bradykinin (a vasodilatory peptide), which contributes to blood
vessel relaxation insufciency. As with the RAS pathway, the production of nitric oxide (NO)
by endothelial nitric oxide synthase (eNOS) is an important mechanism for maintaining normal
blood pressure. NO is a vasodilatory agent that helps maintain regular blood vessel dilation and
normal blood pressure; therefore, when there are insufcient levels of NO (especially under
high levels of angiotensin II), there is less dilation and more contraction. On the basis of these
hypertension-inducing factors, traditional drug treatments have involved the use of compounds
that inhibit activities of renin and/or ACE as well as those that reduce/inhibit binding of
angiotensin II to its receptors or those that increase blood NO levels. However, it is well known
that antihypertensive drugs cause unpleasant negative side effects in some patients and could lead
to reduced compliance with prescribed drug treatment (Flack et al. 1997, Bremmer 2003). For
example, ACE-inhibitory drugs cause dry cough in 520% of patients in addition to other negative
side effects such as loss of taste, hyperkalemia (high levels of blood potassium), skin rashes,
hematological disorders, and angioneurotic edema (Antonios & MacGregor 1995, Semple 1995,
Gunkel et al. 1996, Tenenbaum et al. 2000). Similarly, the use of Aliskiren, a recently approved
renin-inhibitory antihypertensive drug, has been associated with gastrointestinal disorders (Abassi
et al. 2009). Therefore, there has been increased interest in developing natural compounds

236

Aluko

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

ARI

14 March 2015

12:22

that can serve as alternatives to antihypertensive drugs. This is because the use of such natural
products is associated with reduced incidence of negative side effects. Moreover, natural peptides
have the advantages of specicity (limits negative side effects), potency (therapeutic efciency),
and low toxicity such that they can be administered at high doses (Ishida et al. 2011, Thayer
2011). In terms of bioavailability, dipeptides and tripeptides can be absorbed intact by crossing the
intestinal membrane with the help of a peptide-specic transport system, e.g., peptide transporter
1 (PepT1; a proton-dependent transporter). PepT1 uses transmembrane electrochemical proton
gradient as a broad specicity transport force to facilitate exit of small, hydrolysis-resistant
peptides from enterocytes into the surrounding bloodstream (Yang et al. 1999, Satake et al.
2002). In contrast, absorption of peptides with more than three amino acid residues by epithelial
cells occurs through receptor-mediated or nonreceptor-mediated endocytosis, referred to as
pinocytosis when the material is completely soluble (Lee 2002). A third alternative absorption
pathway is through paracellular (aqueous transport through the intercellular space between
adjacent cells) and transcellular routes, both of which allow uptake of intact peptides (Stevenson
& Keon 1998). The high hydrophilic nature of some short-chain naturally occurring peptides
enhances absorption and transport (by passive diffusion) through the paracellular route (Noach
et al. 1994). Absorption through the paracellular route has been demonstrated for an octapeptide
(Conradi et al. 1993) and a tetrapeptide (Satake et al. 2002). The transcellular route also facilitates
peptide uptake through the apical membrane brush border from where peptides are moved
through the enterocyte to the basolateral membrane (Pappenheimer & Michel 2003). Absorption
of hydrophobic peptides is facilitated through the transcellular route (Vermeirssen et al. 2004).
Food protein-derived peptides represent a suitable group of natural compounds that could
serve as alternative antihypertensive agents with less negative side effects. Research in the past
20 years has shown that peptide sequences obtained from food protein proteolysis can act similarly
to drugs as antihypertensive agents by inhibiting in vitro and in vivo renin, ACE, and angiotensin
II receptor activities in addition to enhancing blood NO levels. For example, oral administration
of sour milk to spontaneously hypertensive rats (SHRs) led to detection of active peptides in the
aorta, which was accompanied by decreased ACE activity in the aorta and lungs (Masuda et al.
1996). Matsui et al. (2004) provided evidence that an antihypertensive dipeptide (VY) absorbed
from the gastrointestinal tract (GIT) reduced SHR systolic blood pressure (SBP) by approximately
22% and can be detected in the plasma and various organs such as liver, lung, heart, and kidney.
Recent work also showed reduced renin and ACE activity levels during long-term (48 weeks) oral
administration of a hemp seed protein hydrolysate to SHRs (Girgih et al. 2014a). An egg white
peptide (RVPSL) has also been shown to reduce plasma levels of renin and angiotensin II after oral
administration to SHRs (Yu et al. 2014). Thus, there is evidence that food protein-derived peptides
can be absorbed from the GIT into the blood circulatory system. The following sections discuss
the production method, efcacy testing, and mechanism of action of antihypertensive peptides.

2. METHODS FOR PRODUCING ANTIHYPERTENSIVE PEPTIDES


2.1. Enzymatic Protein Hydrolysis
This involves subjecting protein materials to one or more enzymes, usually at the optimum temperature and pH conditions for each protease (Pihlanto et al. 2008, Makinen et al. 2012, MemarpoorYazdi et al. 2012, Ruiz-Ruiz et al. 2013, Boschin et al. 2014, Garcia-Mora et al. 2014, Ghassem
et al. 2014, Morais et al. 2014). More than one enzyme being used to produce a single protein
hydrolysate can be added simultaneously if optimum hydrolysis conditions are similar (Wang
et al. 2011, Memarpoor-Yazdi et al. 2012, Yamada et al. 2013) or sequentially (Yang et al. 2003,
www.annualreviews.org Antihypertensive Peptides

237

ARI

14 March 2015

12:22

Vastag et al. 2011, You & Wu 2011, Rao et al. 2012a, Gu & Wu 2013, Marrufo-Estrada et al.
2013, Boschin et al. 2014, Li et al. 2014) if different. These proteases can be obtained as puried
commercially available forms (Pihlanto et al. 2008, Kazuki et al. 2009, Ruiz-Ruiz et al. 2013,
Girgih et al. 2011a, He et al. 2013a) or prepared as crude forms (Banerjee & Shanthi 2012, Nasri
et al. 2013, de Gobba et al. 2014, Garca-Tejedor et al. 2014, Ktari et al. 2014). In general, there
are no specic proteins or enzymes needed to generate antihypertensive peptides; however, an
easier-to-digest protein will generate higher peptide yield, which is compatible with potential
commercialization. Thus, various protein materials have been used to generate antihypertensive
peptides from plant, animal, and marine-based proteins. Moreover, enzymes that generate smallsize peptides (<10 amino acid residues) during protein hydrolysis seem to be more desirable
proteolysis tools because low-molecular-weight peptides may be more effective antihypertensive
agents when compared to high-molecular-weight peptides. For example, ACE-inhibitory activities (IC50 ) of <1-, 13-, 35-, 510-, and >10-kDa peptide fractions prepared from a bean protein
hydrolysate obtained by alcalase + avourzyme digestions were 0.001, 0.049, 0.107, 0.268, and
0.268 g/ml, respectively (Ruiz-Ruiz et al. 2013). Lower IC50 values indicate better inhibitory
activity when compared to higher values. The protein content of the starting raw material is irrelevant for producing desirable protein hydrolysates; however, a higher protein content of the
raw material will ensure less contamination of the hydrolysate with nonprotein materials. The
presence of nonpeptide but biologically active materials in the protein hydrolysate may confound
results obtained from the activity tests. For example, phenolic compounds are also known to have
antihypertensive properties; therefore, their removal from proteins prior to proteolysis will reduce
or eliminate this confounding effect. This aspect is illustrated by a previous study, which involved
the use of acetone to remove polyphenols from a hemp seed protein isolate prior to enzymatic
hydrolysis (Girgih et al. 2011a).
During proteolysis, it is necessary to maintain the optimum conditions required for enzyme
activity to ensure efcient release of peptides. Therefore, a constant temperature can be maintained
using a thermostated instrument, whereas a constant pH is usually maintained through the addition
of alkali or acid solution. This is because during proteolysis (peptide bond hydrolysis), protons that
are released will cause the reaction mixture pH to lower, which if not corrected by adding alkali
can lead to low pH values and enzyme inactivation. Adding alkali during proteolysis is usually one
of the reasons the salt content of the hydrolysate increases. The need to add alkali or acid solutions
may be obviated if the proteolysis is performed in a buffer solution, but this could also lead to
high salt content in the hydrolysate. A pH-Stat instrument can be used to avoid manual addition
of alkali solutions during proteolysis; the instrument constantly measures pH during hydrolysis
and automatically adds alkali when deviations to the set pH value occur. In most cases, the protein
material is simply dispersed in water to perform enzymatic hydrolysis; in other cases, the protein
material is dispersed in buffers (Zhang et al. 2013). If a single enzyme is used, the reaction is
maintained to completion at the optimum conditions and hydrolysis durations allowed, which
could range from 124 h depending on the intended product characteristics. The hydrolysis
duration has a substantial inverse relationship with peptide size, but in most cases a plateau is
reached during which further increases in hydrolysis time do not produce any effect on peptide
size or activity. For example, hydrolysis of lentil proteins with different enzymes showed that ACEinhibitory activity was time dependent and reached a maximum after which further prolongation
of proteolysis actually led to reduced peptide activity (Garcia-Mora et al. 2014). This can be
interpreted as meaning that more antihypertensive peptides can be released as protease action
progresses but a maximum number of active peptides will eventually be released. Increasing the
reaction time beyond this point could lead to a decrease in the activity of the protein hydrolysate

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

238

Aluko

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

ARI

14 March 2015

12:22

because of the possibility that some of the already generated active peptides can undergo further
proteolysis to produce inactive fragments.
Garcia-Mora et al. (2014) showed that alcalase hydrolysis produced a lentil hydrolysate with a
maximum ACE-inhibitory activity of 72% at 3 h in comparison to savinase (63%, 2 h), protamex
(47%, 1 h), and corolase (51%, 2 h). Therefore, longer hydrolysis periods are more suitable than
shorter periods to generate peptides with higher levels of potential antihypertensive activities. The
enzyme type is also important; however, this is dependent on the type of protein substrate. Thus, it
is important to perform initial optimization experiments that include varying the proteolysis duration and testing hydrolysate activity to determine the digestion period associated with the highest
peptide activity. Upon completion of enzyme hydrolysis, the reaction mixture is centrifuged to
separate the materials into a supernatant (contains desired peptides), which is collected, and a
discarded precipitate (contains undigested proteins). The supernatant can be desalted, subjected
to further separations such as membrane ultraltration or column chromatography, or simply
freeze-dried. If high salt content is a concern, the protein hydrolysate can be passed through
a nanoltration membrane to obtain a retentate with minimal salt content (Picot et al. 2010).
However, because most nanoltration membranes have close to 300-Da molecular weight cut-off
(MWCO) sizes, it is likely that some small peptides, especially dipeptides, could be lost in the
permeate. Apart from customized in vitro enzymatic hydrolysis, antihypertensive peptides have
also been isolated from naturally produced muscle peptides during meat curing (Escudero et al.
2012a, 2013, 2014). This type of extraction takes advantage of the natural peptides present in
muscles or those that are released by endogenous enzymes during meat processing. For example,
acidied water was used to obtain a peptide extract from Spanish dry-cure ham; after desalting,
the peptides at 10 mg/ml concentration showed up to 60% in vitro inhibition of ACE activity
(Escudero et al. 2014).

2.2. Fermentation
This method is based on some bacteria or yeast species secreting enzymes (including proteases)
into the extracellular medium during growth. Therefore, inoculating protein materials with this
type of bacteria cell can lead to proteolysis and peptide production. In a typical experiment, an
inoculum is prepared by growing the bacterial cells on nutrient agar during incubation at 37 C
for 24 h. After harvesting the cells in sterile distilled water that contains fermentable sugars
(usually glucose), the suspension can then be used as a starter to inoculate sterile protein raw
material for fermentation. Fermentation is allowed to proceed for a period ranging from a few
hours to several days, depending on the type of fermenting microorganism and the desired
peptide product. At the end of fermentation, the product can be used as is, such as sour milk
(Nakamura et al. 1995) for blood pressure reduction, or a known quantity of the fermented
mixture can be mixed with aqueous reagents and shaken thoroughly to extract the peptides
present (Hernandez-Ledesma et al. 2005, Hayes et al. 2007, Ahn et al. 2009, Inoue et al. 2009,
Garca-Tejedor et al. 2013). The extracted solution is then centrifuged to isolate the peptides in
the supernatant, and the precipitate is discarded. The supernatant can be subjected to additional
processing such as indicated in Section 2.1. Fermented peptides can also be further treated with
in vitro enzyme digestion to enhance release of antihypertensive peptides ( Jakubczyk et al. 2013).

3. PEPTIDE FRACTIONATION AND PURIFICATION


Generally, enzymatic protein hydrolysis leads to the production of a hydrolysate that contains
peptides with a wide range of chain lengths, amino acid compositions, and antihypertensive

www.annualreviews.org Antihypertensive Peptides

239

FO06CH11-Aluko

ARI

14 March 2015

12:22

efcacies. Therefore, fractionation is used to separate the hydrolysate peptides into groups based
on chain length, hydrophobicity, or net charge. Depending on the type of protein hydrolysate,
fractionation may increase (Zhang et al. 2009, Li et al. 2014) or decrease (Girgih et al. 2011b)
peptide antihypertensive potency. Peptide purication is required to isolate homogeneous
fractions, which offer better studies of structure-function properties by enhancing amino acid
sequence determination in conjunction with an antihypertensive assay. The main peptide
fractionation protocols are described below.

3.1. Membrane Ultrafiltration

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

The importance of peptide size is related to absorption potential into the blood from the GIT
as well as ability to interact with target enzymes involved in blood pressure regulation. In vitro
studies have shown that smaller-size peptides may be more active than the bigger peptides (Zhang
et al. 2009, Seguro Campos et al. 2010, Ko et al. 2012a, Perez-Vega et al. 2013). However, oral
administration of low molecular weight (<3 kDa) and high molecular weight (>3 kDa) salmon
protein hydrolysate fractions to SHRs showed similar SBP-reducing effects with the latter producing a slightly better reduction (Ewart et al. 2009). Membrane ultraltration offers the possibility
of fractionating peptides into a wide range of peptide sizes using the most common membranes
with the standard MWCO sizes 1, 3, 5, and 10 kDa. However, other membrane types with 2or 6-kDa MWCO sizes have also been used for peptide separation and isolation (Lin et al. 2012,
Liu et al. 2012). Membrane fractionation can be performed to obtain a single or two products by
using a single membrane with a desired MWCO size (Fujita et al. 2001, del Castillo et al. 2007,
Hayes et al. 2007, Li et al. 2011, Liu et al. 2012, Memarpoor-Yazdi et al. 2012, Garca-Tejedor
et al. 2014) or different fractions with various membranes (Girgih et al. 2011b, Wang et al. 2011,
Ruiz-Ruiz et al. 2013). To obtain fractions that have distinct differences in peptide size, the
hydrolysate is rst passed through a membrane with the smallest MWCO size, such as the 1-kDa
membrane; the ow-through solution (permeate) is then collected as the <1-kDa peptide fraction.
The retained solution (retentate) can be passed through the 3-kDa membrane, and permeate
will contain peptides with sizes bigger than 1 kDa but less than 3 kDa (13-kDa fraction). The
retentate from the 3-kDa membrane can be passed through a 5-kDa membrane to obtain a
35-kDa permeate fraction. Finally, the 5-kDa membrane retentate is then passed through the
10-kDa membrane to obtain a permeate with 510-kDa peptides. The fractionation process can
be reversed by starting with the highest membrane MWCO size and nishing with the lowest
( Jung et al. 2006, Zhang et al. 2009). For example, the protein hydrolysate can be rst passed
through a 10-kDa membrane and the permeate collected and passed through a 5-kDa (or 6-kDa)
membrane; the retentate will be collected as the 510-kDa (or 610-kDa) fraction (Lin et al.
2012, Puchalska et al. 2014). The permeate from the 5-kDa (or 6-kDa) membrane is then passed
through a 3-kDa (or 2-kDa) membrane whose permeate is passed through a 1-kDa membrane
to obtain retentates that are 35- and 13-kDa peptide fractions. The nal permeate obtained
from the 1-kDa membrane will be taken as the <1-kDa peptide fraction. During ultraltration,
it is common to add distilled water at intervals to reduce solution viscosity and improve the rate
of peptide permeation of the membrane; this process is called dialtration and can be used to
enhance the yield and speed of the membrane ultraltration process. However, the membranes do
not have 100% efciency, and it is possible that a small amount of peptides with sizes higher than
the MWCO size could pass through into the permeate. To reduce fouling caused by peptides that
stick to the membrane during ultraltration, it is better to use membranes that have been designed
specically for protein separation. This is because membrane fouling can reduce efciency of the
ltration process and lead to low peptide yields in addition to extended ltration periods.
240

Aluko

FO06CH11-Aluko

ARI

14 March 2015

12:22

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

3.2. Column Chromatography


Two main column chromatography methods, medium-pressure (fast protein liquid chromatography, FPLC) and high-pressure liquid chromatography (HPLC), are used for peptide separation.
FPLC separation of peptides uses mostly gel permeation or ion-exchange (cation or anion) columns
(Sheih et al. 2009, Lee et al. 2010, Wang et al. 2011, You & Wu 2011, Banerjee & Shanthi 2012,
Intarasirisawat et al. 2013, Tomatsu et al. 2013). Protein hydrolysate fractionation on an FPLC
column is designed principally for initial peptide separation followed by further purication on
an HPLC column (Murakami et al. 2004, Zhang et al. 2006, Wang et al. 2008, Balti et al. 2010,
Jang et al. 2011, Wang et al. 2011, Lee et al. 2012, He et al. 2013c, Intarasirisawat et al. 2013, Liu
et al. 2013, de Gobba et al. 2014). Therefore, fractions within each peak are pooled, freeze-dried,
and assayed for potential antihypertensive activity using ACE or renin-inhibition assay. The peak
that shows the highest in vitro enzyme inhibition activity is then subjected to HPLC separation
using a reverse-phase column (RP-HPLC) during which fractions are also collected and assayed
for activity (Wang et al. 2011; Memarpoor-Yazdi et al. 2012; He et al. 2013b,c; de Gobba et al.
2014; Girgih et al. 2014c). In most instances, a single round of RP-HPLC may not be adequate to
produce homogeneous peptide fractions; therefore, the most active fraction is usually subjected to
a second or third round of RP-HPLC separation to obtain pure peptides (Chen et al. 2013a, He
et al. 2013b). Pure peptides have also been isolated through initial ion-exchange chromatography
followed by fractionation and refractionation of the most active peaks by gel permeation chromatography (Majumder & Wu 2009, Banerjee & Shanthi 2012). Mass spectrometry can check
the purity of each peptide peak, the point at which a single specie dominates (>90%) the mass
spectrum. The mass of each peptide peak can be used to check for matching amino acid sequences
if the primary structure of the parent protein is available on public online protein databases (Hayes
et al. 2007, Gu & Wu 2013, de Gobba et al. 2014, Udenigwe et al. 2012a). Alternatively, in situ
amino acid sequencing can be performed for each peptide fraction using tandem mass spectrometry (Qian et al. 2007; Majumder & Wu 2009; Balti et al. 2010; Jang et al. 2011; Ahn et al. 2012;
Lee et al. 2012; Du et al. 2013; He et al. 2013b,c; Tanzadehpanah et al. 2013; Garca-Tejedor
et al. 2014; Girgih et al. 2014c; Norris et al. 2014; Puchalska et al. 2014) or automated Edman
degradation (Marczak et al. 2003, Jang & Lee 2005, Jung et al. 2006, Lee et al. 2006, Wang et al.
2008, Chen et al. 2013a). However, gel permeation chromatography can also be used to separate the protein hydrolysate into peptide fractions of known molecular sizes followed by assay of
each fraction for potential in vitro or in vivo antihypertensive properties (Escudero et al. 2012a,
Ruiz-Ruiz et al. 2013). Similar to ultraltration fractionation, the results from the gel permeation
chromatography separation can be used to estimate the relationship between peptide size and in
vitro or in vivo antihypertensive property.

4. ACTIVITY ASSAY
4.1. In Vitro Methods
Several methods have been devised for estimating potential antihypertensive properties of peptides;
generally, these methods involve ACE or renin activity assays. The rst ACE-inhibitory assay was
developed using hippuryl-histidyl-leucine (HHL) as a substrate; cleavage of the hippuryl-histidine
bond releases free hippuric acid, which can then be extracted into ethyl acetate. The solvent is
evaporated and the hippuric acid residue dissolved in distilled water followed by spectrophotometric measurement at 228 nm (Cushman & Cheung 1971). Alternatively, the reaction mixture can
be separated on a reverse-phase HPLC column to quantify the hippuric acid peak (Wu et al. 2002,
Wang et al. 2011). In the presence of peptide inhibitors, ACE-mediated production of hippuric
www.annualreviews.org Antihypertensive Peptides

241

ARI

14 March 2015

12:22

acid is reduced, and the percentage ratio to the value in the absence of peptide can be calculated as
the percent inhibition. A major drawback of the spectrophotometric assay is that if excessive heat
is used for the ethyl acetate evaporation, some of the hippuric acid may not be dissolved in water,
which causes underestimation of enzyme activity and overestimation of peptide inhibitory capacity. The ethyl acetate may be evaporated with nitrogen gas but if solvent residues remain, there is
overestimation of enzyme activity and underestimation of peptide inhibitory activity because the
ethyl acetate also absorbs ultra-violet (UV) radiation at 228 nm, where hippuric acid is normally
measured. Moreover, the ethyl acetate extract can sometimes be contaminated with the substrate
(HHL), which also absorbs at 228 nm and can lead to overestimation of ACE activity and underestimation of peptide inhibitory activity (Wu et al. 2002). A recent report has conrmed the higher
sensitivity and precision of the HPLC method when compared to the spectrophotometric method
(Chen et al. 2013b). An alternative ACE assay involves uorimetry and uses ortho-aminobenzoic
acid-phenylalanine-arginine-lysine-dinitrophenyl-proline [abz-FRK(dnp)P-OH] as the substrate
(van Elswijk et al. 2003). ACE cleaves the arginine-lysine (RK) bond, which removes the DNP
quenching moiety and causes enhanced uorescence (excitation at 320 nm and emission at
420 nm) of the ABZ moiety. Therefore, ACE-inhibitory peptides will reduce the rate of substrate
cleavage, which is measured as less DNP release and hence reduced uorescence emission values.
Another commonly used ACE assay involves N-(3-[2-furyl]acryloyl)-phenylalanylglycylglycine
(FAPGG) as the substrate (Holmquist et al. 1979). ACE hydrolyzes FAPGG into FAP and GG;
the reaction can then be followed by continuous UV absorption measurement to determine the
decrease in absorbance at 345 nm as a result of the Phe-Gly peptide bond cleavage (Udenigwe et al.
2009). The rate of absorption decrease in the presence of peptide inhibitors will be less and can be
subtracted from the reaction rate in the absence of the inhibitor to obtain percentage inhibition.
Alternatively, the released FAP can be quantied by injecting the reaction mixture directly onto
a reverse-phase HPLC column with detection at 305 nm (Lahogue et al. 2010). In general, the
FAPGG method is simpler and faster to complete and is therefore better suited to routine analysis
of several samples when compared to the HHL assay (Shalaby et al. 2006).
The most widely used renin-inhibitory assay is based on Wang et al.s (1993) method and is
specic for human renin. This method uses an internally quenched uorogenic substrate ArgGlu(EDANS)-Ile-His-Pro-Phe-His-Leu-Val-Ile-His-Thr-Lys(Dabcyl)-Arg; renin hydrolyzes
the Leu-Val bond to yield a highly uorescent Arg-Glu(EDANS)-Ile-His-Pro-Phe-His-Leu
(called peptide-EDANS) product. The amount of peptide-EDANS released can be measured at
335345-nm excitation and 485510-nm emission wavelengths; in the presence of renin-inhibitory
peptides, the emission values are smaller and percentage inhibition can be calculated in relation
to the uninhibited reaction (Udenigwe et al. 2009, Li & Aluko 2010, Girgih et al. 2011b, He et al.
2013c, Fitzgerald et al. 2014). Another uorescence method uses substrates specic for human or
porcine renin (Takahashi et al. 2008). The human renin substrate is N-methylanthranyl (Nma)-IleHis-Pro-Phe-His-Leu-Val-Ile-His-Thr-Lys-2,4 dinitrophenyl (dnp)-D-Arg(r)-r-NH2 , and the
porcine renin substrate is Nma-His-Pro-Phe-His-Leu-Leu-Val-Tyr-Lys(dnp)-r-r-NH2 . Hydrolysis occurs at the Leu-Val and Leu-Leu peptide bonds, respectively, for the human and porcine
substrates (Takahashi et al. 2008). The uorescent fragment (contains the Nma moiety) is then
measured at excitation and emission wavelengths of 340 and 440 nm, respectively. The high sensitivity of these uorescence methods enables identication of a wide range of renin-inhibitory
compounds because minor changes in enzyme activity can still be detected.

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

4.2. In Vivo Methods


Historically, the SHR has been the most used genetically hypertensive rat model for testing
the in vivo antihypertensive effects of food protein-derived protein hydrolysates and peptides.
242

Aluko

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

ARI

14 March 2015

12:22

This is because the SHR is the closest model of human arterial hypertension and has been used
extensively for antihypertensive research in safety and efcacy testing (Feld et al. 1981, Tschudi
et al. 1994, Curtis et al. 1995, Zheng et al. 1997, Kvam et al. 1998, Sipola et al. 2001, Sato
et al. 2002, Suetsuna et al. 2004, Wood et al. 2005, Katayama et al. 2007, Hiwatashi et al. 2010,
Sanchez et al. 2010, Herregods et al. 2011, Lu et al. 2011). Due to growth differences, male SHRs
are more commonly used than females to test efcacy of antihypertensive compounds. This is
because male SHRs can grow up to 300 g within a 15-week period in comparison to a maximum
weight of approximately 200 g for females (http://www.criver.com/products-services/basicresearch/find-a-model/spontaneously-hypertensive-%28shr%29-rat). The males can develop systolic and diastolic blood pressure values of up to 203 and 176 mmHg, respectively;
similar values for females are 191 and 154 mmHg, respectively (http://www.criver.com/
products-services/basic-research/find-a-model/spontaneously-hypertensive-%28shr%29rat). SHRs can survive up to 66 weeks independent of whether they receive antihypertensive
treatments (Feld et al. 1981) and can therefore serve as useful models to determine long-term
effects of antihypertensive peptides. A recent work has shown that the plasma renin activity in
SHRs is almost twice that of the normotensive equivalent (Girgih et al. 2014a); therefore, the
SHR provides an excellent model to test renin-inhibitory peptides. Moreover, the SHRs have
been shown to respond positively, which was measured as decreased blood pressure when treated
with antihypertensive agents that inhibit renin (Wood et al. 2005; Hiwatashi et al. 2010; Chou
et al. 2013; Girgih et al. 2014a,c) and ACE (Curtis et al. 1995; Miguel et al. 2010; Sanchez et al.
2010; Fernandez-Musoles et al. 2013; Girgih et al. 2014a,c) activities. Previous works have also
shown that the SHR is a good model for testing ability of antihypertensive peptides to cross the
gastrointestinal barrier and be present within the blood circulatory system (Masuda et al. 1996,
Matsui et al. 2004). Recent works have shown that ACE-inhibitory peptides (VPP, IPP, AHIII)
not only reduced the blood pressure of SHRs but induced endothelium relaxation through
increased NO production (Hirota et al. 2011, Ko et al. 2012b). Therefore, the SHR also serves
as a suitable animal model to determine the NO-enhancing and vasorelaxation properties of
antihypertensive peptides. Most tests involve oral administration of the peptides but intravenous
injection has also been reported (Lee et al. 2012). However, a polycystic kidney disease rat model
was also recently used to show effective blood pressure reduction by a pea protein hydrolysate
during an 8-week oral feeding trial (Li et al. 2011).

5. MECHANISM OF ACTION
In most cases, the exact cause of human systemic hypertension is unknown but the disease is
characterized by excessive RAS enzyme activities, which lead to abnormally high levels of plasma
angiotensin II. The high plasma angiotensin II level causes excessive blood vessel constriction with
reduced vasorelaxation and enhanced superoxide radical production, which lead to increased NO
degradation. The net effect is high blood pressure and the pathological condition of hypertension.
Bioactive peptides have been shown to produce antihypertensive effects mainly as a result of
actions that involve inhibition of RAS enzyme activities (ACE and renin); however, enhanced
production of NO has also been demonstrated. For example, in long-term experiments, it has
been shown that plasma, lung, kidney, and aorta renin and/or ACE activities as well as angiotensin
II levels can be downregulated by antihypertensive protein hydrolysates (Wang et al. 2012,
Fernandez-Musoles et al. 2013, Girgih et al. 2014a). However, long-term oral administration of
antihypertensive corn peptides reduced ACE activity in lungs but not in the plasma (Huang et al.
2011). The antihypertensive effect of an egg protein-derived tripeptide (IRW) was associated with
upregulation of the NO synthesis pathway in addition to ACE inhibition (Majumder et al. 2013).
www.annualreviews.org Antihypertensive Peptides

243

FO06CH11-Aluko

ARI

14 March 2015

12:22

5.1. ACE Inhibition

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Protein hydrolysates have been shown to inhibit in vitro and in vivo activities of ACE, but activity
is highly dependent on the type of protease(s) used or the proteolysis conditions, in addition to
the nature of the raw material (Ewart et al. 2009, Ko et al. 2012a, Marrufo-Estrada et al. 2013, He
et al. 2013a, Alashi et al. 2014). For example, hydrolysis of whey protein concentrate (WPC) with
alcalase was found to produce protein hydrolysates with higher levels of ACE-inhibitory activities
than similar hydrolysates produced from avourzyme and pancreatin (Morais et al. 2014). Li
et al. (2006) also showed that the antihypertensive activity of an alcalase mung bean hydrolysate
was greater than that of the neutrase hydrolysate. The differences in product activity are due
to peptide bond hydrolysis specicity differing among enzymes and, as such, each hydrolysate
containing peptide mixtures that differ in amino acid composition and chain length. Even for
the same enzyme and starting protein material, varying the enzyme-to-substrate ratio (E:S) can
produce hydrolysates with different ACE-inhibitory properties, further illustrating the importance
of peptide characteristics. For hydrolysis of WPC, there was increased ACE-inhibitory activity of
the hydrolysates at high E:S values, which indicates that the presence of more enzyme molecules
could enhance production of active peptides. However, the relationship between the enzyme
concentration and ACE-inhibitory activity of the WPC protein hydrolysates was not exactly linear.
The exact reason for this nonlinearity is unknown but may be related to inactivation (degradation)
of active peptides at some enzyme levels. In general, E:S values used for food protein hydrolysate
production can range from 0.1 to 10%, but hydrolysis should be optimized so as to not use excessive
enzyme levels that will not translate to better activity of the products.
Modes of action of certain protein hydrolysates and peptides have shown that ACE inhibition
takes place through competitive, noncompetitive, uncompetitive, or even mixed-type peptideenzyme interactions. For example, isolated peptides from collagen hydrolysates were shown to
inhibit ACE activity in a competitive manner, which indicates interaction with the active site
and competition with the substrate to block proteolysis (Banerjee & Shanthi 2012). In contrast,
Memarpoor-Yazdi et al. (2012) showed that egg white lysozyme-derived peptides inhibited ACE
activity in a noncompetitive manner, which suggests interaction with enzyme proteins at locations
other than the active site. Memarpoor-Yazdi et al. (2012) also showed that interaction of the
lysozyme peptides with ACE led to a reduction in -helix, whereas unordered conformation
increased. The increased level of disordered fraction could have led to the reduced catalytic ability
of the ACE molecule. An almond protein hydrolysate containing <1-kDa peptides was also found
to inhibit ACE activity in a noncompetitive manner (Wang et al. 2011).
Several studies have examined the role of peptide chain amino acid position on the potential for higher ACE inhibition. Even though most of these studies involve statistical modeling
coupled with in vitro experiments, they are relevant to our ability to design peptides with high
antihypertensive potency. This is because knowledge of critical amino acid residues required on
a peptide chain can allow selection of specic enzymes that will liberate peptides with desirable
amino acid contents or sequences. For example, if arginine and/or lysine are important for ACE
inhibition, then trypsin could be an effective proteolysis tool. Similarly, thermolysin has a high
propensity to hydrolyze peptide bonds that involve branched-chain amino acids (BCAAs), and its
hydrolysate will contain higher levels of these amino acids. BCAAs, especially valine and isoleucine,
are components of the famous milk tripeptides (VPP and IPP) that were identied as potent ACE
inhibitors and antihypertensive agents very early in bioactive peptide research (Nakamura et al.
1995). Proline, especially when present at the C-terminal, has also been shown to potentiate
high ACE inhibition; therefore, prolyl endopeptidases and other proteases that generate prolinecontaining peptides may be useful protein hydrolysis tools for producing strong antihypertensive

244

Aluko

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

ARI

14 March 2015

12:22

peptides. Aromatic (or bulky) amino acids (AAAs) can also potentiate high ACE-inhibitory activity
of peptides and enhance antihypertensive effects. Thus, enzymes such as pepsin, chymotrypsin,
and alcalase provide important protein cutting tools for producing AAA-enriched peptides with
potentially strong ACE-inhibitory and antihypertensive properties. In terms of specic amino acid
positions and enhanced ACE-inhibitory activity of peptides, it has been shown that for dipeptides,
the presence of AAAs or hydrophobic side chains is important (Wu et al. 2006b). For tripeptides, the presence of an AAA at the C-terminal coupled with a positively charged residue in the
middle position and a hydrophobic residue at the N-terminal was shown to be important for
strong ACE inhibition (Wu et al. 2006b). Wu et al. (2006b) then used their statistical model to
predict various peptide sequences, which were then validated through in vitro ACE-inhibition
assay; LRW, IKP, and FW were shown to have experimental IC50 values of 0.23, 2.75, and
5.89 M, respectively. For longer peptides, it was shown that the ACE-inhibitory activity was
highly dependent on the amino acid sequence C-terminal tetra residue section (Wu et al. 2006a).
Thus, for a strong ACE-inhibitory tetrapeptide, the following amino acids are critical: C1 (tyrosine, proline, phenylalanine), C2 (phenylalanine), C3 (arginine, histidine, tryptophan, phenylalanine), and C4 (valine, isoleucine, methionine). To enhance ACE-inhibitory potency of pentapeptides and longer chains, these amino acid residues are critical at the four C-terminal positions:
C1 (tyrosine, cysteine), C2 (histidine, tryptophan, methionine), C3 (valine, leucine, isoleucine,
methionine), and C4 (tryptophan). However, a recent work also determined the favorable amino
acid residues for the C-terminal pentapeptide residues of strong ACE-inhibitory peptides, which
were reported as follows (Sagardia et al. 2013): C1 (glycine, leucine, alanine, valine, isoleucine),
C2 (arginine, valine, threonine), C3 (aspartic acid, asparagine, lysine), C4 (tryptophan, tyrosine,
cysteine), and C5 (valine, isoleucine). All the reports agree that the C-terminal sequence of a peptide contains the critical amino acid residues necessary for strong ACE inhibition. There is also
some agreement with regard to specic amino acid requirements, as all the models reported the
importance of a positively charged amino acid at the C2 and tryptophan at the C4 of long-chain
(>4 residues) peptides. However, there are still major differences in the reported required amino
acids for other C-terminal positions, which may be due to variations in modeling tools and peptide
databases. A major agreement between these peptide structure-activity modeling studies is that
proline did not appear important for C-terminal activity, which is contrary to previous reports for
milk peptides. The reason for this discrepancy is unknown but it could be that the physicochemical
properties of individual amino acids on which the statistical modeling is based are not enough to
provide accurate predictions for all possible potent ACE-inhibitory amino acid combinations.

5.2. Renin Inhibition


Unlike ACE inhibition, there have been only very few reports on the renin-inhibitory properties of food protein-derived peptides. The reason for this areas slow development could be due
to renin being a more difcult enzyme to inhibit and the costs associated with the assay being
signicantly higher than those associated with ACE. One of the earliest works identied three
dipeptides from an alcalase digest of yellow pea seed proteins; IR, KF, and EF had 9.2, 17.8, and
22.7-mM IC50 values, respectively (Li & Aluko 2010). On the basis of these three peptides, it
was suggested that a hydrophobic residue at the N-terminus combined with a bulky amino acid
residue at the C-terminus could enhance renin inhibition by dipeptides. Udenigwe et al. (2009)
showed that a axseed protein hydrolysate fraction with a more cationic character inhibited renin
activity more than the fraction with less cationic properties. Moreover, kinetic studies showed the
cationic fraction inhibited renin through an uncompetitive mode of action, which suggests lack
of afnity for the enzyme active site. More recent works have shown the in vitro renin-inhibitory
www.annualreviews.org Antihypertensive Peptides

245

ARI

14 March 2015

12:22

capacities for various food protein hydrolysates such as hemp seed (Girgih et al. 2011b, 2014a),
rapeseed (He et al. 2013a,b,c), canola (Alashi et al. 2014), macroalga (Fitzgerald et al. 2012), kidney
bean (Mundi & Aluko 2014), African yam seed (Ajibola et al. 2013), and chicken skin (Onuh et al.
2013). A quantitative structure-activity relationship (QSAR) study of dipeptides showed that the
presence of a bulky amino acid at the C-terminal with a hydrophobic amino acid at the N-terminal
favored higher levels of renin-inhibitory activity than peptides without C-terminal hydrophobic
residues (Udenigwe et al. 2012b). Using the QSAR model, four tryptophan-containing antihypertensive dipeptides (IW, LW, VW, and AW) were predicted to have the highest potential as renin
inhibitors. However, in vitro validation of their renin-inhibitory activity showed IW to be most
active (IC50 = 2.3 mM), followed by LW (IC50 = 3.2 mM); VW and AW were inactive. Thus,
the structural similarity between isoleucine and leucine may be responsible for this distinguishing inhibitory capacity when compared to other tryptophan-containing dipeptides. Using pure
peptide sequences found in rapeseed protein hydrolysates, it has been shown that higher renin
inhibition is dependent on the ability of the peptide to reduce the -helix and -sheet fractions of
the enzyme in addition to forming several hydrogen bond interactions with the active site residues
(He et al. 2014). Recent work showed that hemp seed peptides (WYT, SVYT, WVYY, IPAGV)
also interacted with the active site of renin; specically, WYT formed electrostatic bonds with the
aspartic acid residues that participate in enzyme catalysis (Girgih et al. 2014b). Therefore, these
works have shown that certain food protein-derived peptides can bind to active or nonactive sites
of renin to modulate catalytic activity. In vivo conrmation of activity was also obtained for the
hemp seed protein hydrolysate, which was shown to lower plasma renin activity in SHRs during
long-term oral feeding (Girgih et al. 2014a). However, it should be stressed that previous works
have not been able to establish a correlation between the renin-inhibitory activities and ACEinhibitory activities of food protein-derived peptides. Therefore, it seems the mode of action of
these peptides against ACE is different from that against renin.

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

5.3. Nitric Oxide Production


By inhibiting excessive accumulation of angiotensin II in the blood, antihypertensive peptides
will reduce superoxide free radical (an NO degradation agent) levels and contribute to increased
plasma NO levels. Moreover, as part of their multifunctional properties, food protein-derived
peptides have been shown to enhance the eNOS pathway, which is responsible for producing
NO. For example, the antihypertensive peptide IRW that was isolated from egg ovotransferrin
was shown to restore eNOS expression in the SHR mesenteric artery and aorta (Majumder et al.
2013). Using cultured vascular endothelial cells, it was demonstrated that addition of the milk
protein-derived ACE-inhibitory tripeptides (VPP and IPP) led to increased NO production and
better endothelium-dependent relaxation of aortic rings (Hirota et al. 2011). Increased production
of eNOS leads to higher levels of NO in the vascular system, which can enhance blood vessel
relaxation and lead to blood pressure reduction. The effects of VPP and IPP were inhibited when
coincubated with NOS inhibitors, which conrmed the production of NO as contributing to the
vasorelaxative action of these antihypertensive peptides. Similarly, oral administration of a silk
broin hydrolysate to SHRs led to 8, 13, and 22% increases in plasma NO content at daily doses
of 100, 600, and 1,200 mg/kg body weight (bw), respectively, which was accompanied by reduced
levels of vasoconstrictive endothelin (Yuan et al. 2012).

5.4. Angiotensin II Receptor Blockage


There is scant information on the angiotensin II receptor-blocking effects of food protein-derived
peptides. AT1 is the main angiotensin receptor in vascular walls, and binding of angiotensin II
246

Aluko

FO06CH11-Aluko

ARI

14 March 2015

12:22

leads to blood vessel contractions. The antihypertensive drug valsartan prevents interaction of
angiotensin II with AT1 , which contributes to vascular relaxation. Similarly, RPYL, a lactoferrinderived peptide, has been shown to have concentration-dependent inhibition (up to 62% at
300 M) of interaction between angiotensin II and human angiotensin AT1 receptors in an in vitro
model (Fernandez-Musoles et al. 2013). RPYL also inhibited angiotensin I-dependent vascular
contraction, which suggests ACE inhibition as an additional mechanism of its antihypertensive
activity.

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

6. BLOOD PRESSUREREDUCING EFFECTS


The ability of protein hydrolysates to reduce blood pressure is highly dependent on several factors,
which include resistance to structural degradation and functional inactivation by gastrointestinal
enzymes. Because the antihypertensive activity of peptides is dependent on amino acid composition
and sequence, the loss of amino acid residues as a result of proteolysis within the GIT can lead to
reduced or increased potency or complete loss of activity. Peptides must also be absorbed from
the GIT and transported through the blood circulatory system to elicit antihypertensive effects.
To conrm GIT stability, the protein hydrolysates or peptides can be treated in vitro with pepsin
followed by pancreatin (or trypsin + chymotrypsin) at 37 C and under appropriate pH conditions;
this is referred to as simulated GIT digestion (Yang et al. 2003, Jang et al. 2011, Escudero et al.
2014, Garcia-Mora et al. 2014, Puchalska et al. 2014). A comparison of the ACE or renin-inhibitory
activity before and after treatment can then be made. A reduction or loss of in vitro inhibitory
activity after the simulated GIT digestion (Banerjee & Shanthi 2012, Puchalska et al. 2014) signies
potential inactivity (no reduction in blood pressure) if the protein hydrolysate is provided orally.
However, if the in vitro activity is unchanged (Katayama et al. 2007, Wang et al. 2008, Jimsheena
& Gowda 2010, Lee et al. 2010, Vercruysse et al. 2010, Ko et al. 2012a, Escudero et al. 2014)
or increases (Ewart et al. 2009, Jang et al. 2011, Aleman et al. 2013, Garcia-Mora et al. 2014)
after the simulated GIT digestion, the protein hydrolysate has potential in vivo activity and can be
tested in a suitable animal hypertension model. For peptides, the original sequence could undergo
fragmentation into shorter peptides that are less or more active. In addition, within the blood
vessels the peptides must not become substrates for renin and ACE to avoid structural degradation
and loss of activity. Similar to the simulated GIT treatment, the protein hydrolysate or peptide can
also be incubated with ACE (Yang et al. 2003, Rao et al. 2012b) or renin followed by determination
of in vitro activity (ACE or renin inhibition). The results obtained will be interpreted similarly to
those explained above for simulated GIT treatment. For example, incubation of spinach Rubisco
peptides with ACE showed that peptide MRWRD was fragmented into a more potent MRW
and a less potent RW, whereas peptide MRW was unaffected (Yang et al. 2003). Lee et al. (2012)
also showed that the ACE-inhibitory activity of peptides VAW and MKR was not affected by
preincubation with ACE. To the contrary, incubation of peptide LRIPVA with ACE led to LR
and IP fragments with 120 times less activity. Peptide IAYKPAG was also fragmented by ACE
into IAY and KP with reduced ACE-inhibitory activity (Yang et al. 2003). Similarly, the ACEinhibitory activity of peptide RGY was decreased after incubation with ACE (Rao et al. 2012b).
On the basis of these principles, Fujita et al. (2000) classied protein hydrolysates and peptides
into the following three main categories. In category one, true inhibitors are the peptides that
maintain their ACE-inhibitory values. Category two comprises peptides that exhibit reduced ACE
inhibition after incubation with ACE and are called substrates. Prodrug peptides, to the contrary,
are converted into true inhibitors after incubation with ACE or GIT proteases and belong to
category three. Therefore, only true inhibitors and prodrug peptides will likely exert in vivo blood
pressurelowering effects.
www.annualreviews.org Antihypertensive Peptides

247

FO06CH11-Aluko

ARI

14 March 2015

12:22

6.1. Protein Hydrolysates

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Initial testing of potential antihypertensive activity of protein hydrolysates involves in vitro inhibition of renin and/or ACE activities. Hydrolysates that show highest levels of in vitro activities
are then usually orally administered to SHRs to determine blood pressurereducing ability (Fujita
et al. 2001, Herregods et al. 2011); however, in some instances all the hydrolysates are orally
administered (Li et al. 2006, Girgih et al. 2011b, He et al. 2013a, Alashi et al. 2014). To determine
the effect of peptide size, membrane ultraltration fractions with dened peptide size ranges are
also orally administered for comparison with the whole (unfractionated) hydrolysate (Girgih et al.
2011b). SHR tests are usually carried out either on a short-term (024 h) basis with blood pressure measurements at 12 h intervals or on a long-term basis consisting of 28 weeks of feeding
experiments. The short-term trials provide useful information on daily efciency of the protein
hydrolysate (Li et al. 2006, Nakano et al. 2006, Wang et al. 2008, Girgih et al. 2011b, Wang et al.
2011, Ko et al. 2012a, Alashi et al. 2014), whereas long-term trials are used to determine either
preventive or treatment effects (Yang et al. 2004, Nakano et al. 2006, Wang et al. 2008, Huang
et al. 2011, Wang et al. 2011, Lin et al. 2012, Liu et al. 2012, Fernandez-Musoles et al. 2013,
Yamada et al. 2013, Girgih et al. 2014a). To determine preventive effects, the protein hydrolysate
is given to young (growing) SHRs. Blood pressure is measured weekly and compared with that of
SHRs that consumed the placebo diet (Yang et al. 2004, Nakano et al. 2006, Girgih et al. 2014a).
Treatment potential can be determined by feeding the adult rats (with full-blown hypertension)
the protein hydrolysate coupled with weekly blood pressure measurements (Girgih et al. 2014a).
Direct use of protein hydrolysates as antihypertensive agents has the benet of reduced costs
given the absence of the need for peptide separation and purication. Moreover, the synergistic
effects of different peptides present in protein hydrolysates could produce better blood pressure
reducing effects than when the peptides have been separated into less heterogeneous mixtures.
For example, separation of a hemp seed protein hydrolysate into <1-kDa or 13-kDa peptide
fractions resulted in reduced efcacy as a blood pressurereducing agent when compared to the
unfractionated hydrolysate (Girgih et al. 2011b). The peptide fractions showed a maximum of
18 mmHg reduction after 6 h of oral administration to SHRs when compared to 30 mmHg
after 8 h for the unfractionated hydrolysate. More importantly, the unfractionated hydrolysate
was still active (19 mmHg), whereas the peptide fractions were inactive after 24 h of oral administration. The demonstrated greater antihypertensive effectiveness of unfractionated hydrolysate
is consistent with synergistic effects of varying peptide sizes. Table 1 shows the blood pressure
reducing effects of different food protein hydrolysates after oral administration to SHRs. The data
show the different protein sources, peptide production tools, and protein hydrolysate doses. More
importantly, Table 1 shows that the antihypertensive effect can be directly dose dependent, like
the silk broin, or not, like the soybean and almond. Table 1 also shows different rates of blood
pressurelowering activity; the mushroom peptide extract produces a maximum decrease within
30 min ( Jang et al. 2011), whereas protein hydrolysates such as whey (Wang et al. 2012), sesame
(Nakano et al. 2006), and rapeseed (He et al. 2013a) take 824 h. These results demonstrate
that potency of antihypertensive protein hydrolysates is highly dependent not only on starting
raw material but also the type of enzyme used for proteolysis. Commercial products showed that
gelatins had no effect on the blood pressure of SHRs after oral administration (Herregods et al.
2011), which suggests lack of absorption, minimal presence of bioactive peptides, or structural
degradation of active peptides in the GIT. However, subsequent hydrolysis of the gelatins with
thermolysin produced a protein hydrolysate that reduced SHR SBP by almost 17 mmHg after
6 h; thus, the gelatin was broken down by thermolysin into antihypertensive peptides (Herregods
et al. 2011). The thermolysin-induced antihypertensive gelatin peptides acted through inhibition

248

Aluko

FO06CH11-Aluko

ARI

14 March 2015

12:22

Table 1 Effects of oral administration of food protein-derived protein hydrolysates on short-term (24 h) changes in the
systolic blood pressure (SBP) of spontaneously hypertensive rats
SBP reduction

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Protein source

Protease(s)

Dose (mg/kg bw)

(mmHg)a

References

300

17 after 6 h

Herregods et al. 2011

Pepsin

1,000

10 after 6 h

Li et al. 2014

Trypsin

1,000

22 after 6 h

Li et al. 2014

Pistachio

Pepsin + Trypsin

1,000

22 after 6 h

Li et al. 2014

Fish

Thermolysin

500

15 after 4 h

Fujita et al. 2001

Sesame seed

Thermolysin

30 after 8 h

Nakano et al. 2006

Canola meal

Pancreatin

200

15 after 6 h

Alashi et al. 2014

Canola meal

Pepsin

200

24 after 4 h

Alashi et al. 2014

Canola meal

Trypsin

200

5 after 8 h

Alashi et al. 2014

Canola meal

Alcalase

200

34 after 4 h

Alashi et al. 2014

Canola meal

Chymotrypsin

200

15 after 6 h

Alashi et al. 2014

Rapeseed meal

Proteinase K

100

5 after 8 h

He et al. 2013a

Rapeseed meal

Thermolysin K

100

9 after 8 h

He et al. 2013a

Rapeseed meal

Alcalase

100

24 after 8 h

He et al. 2013a

Rapeseed meal

Flavourzyme

100

17 after 6 h

He et al. 2013a

Rapeseed meal

Pepsin + Pancreatin

100

20 after 24 h

He et al. 2013a

Almond

Neutrase + N120P

400

17 after 2 h

Wang et al. 2011

Almond

Neutrase + N120P

800

21 after 2 h

Wang et al. 2011

Bamboo shoot

None (aqueous extract)

20

11 after 4 h

Liu et al. 2013

Bamboo shoot

None (aqueous extract)

50

18 after 4 h

Liu et al. 2013

Bamboo shoot

None (aqueous extract)

100

27 after 5 h

Liu et al. 2013

Mung bean

Alcalase

600

30 after 6 h

Li et al. 2006

Spanish ham

None (aqueous extract)

4.56

38 after 6 h

Escudero et al. 2012a

Spanish ham

None (aqueous extract)

1.48

28 after 6 h

Escudero et al. 2012a

Spanish ham

None (aqueous extract)

8.7

24 after 6 h

Escudero et al. 2012a

Dairy whey

Crude enzyme extract

400

22 after 4 h

Tavares et al. 2012

Oyster

Pepsin

20

16 after 4 h

Wang et al. 2008

Gelatin

Thermolysin

Pistachio
Pistachio

Mushroom

None (aqueous extract)

600

48 after 0.5 h

Jang et al. 2011

Rapeseed

Pepsin

500

7 after 4 h

Marczak et al. 2003

Rapeseed

Alcalase

500

16 after 4 h

Marczak et al. 2003

Soybean

Protease D3

50

25 after 2 h

Kodera & Nio 2006

Soybean

Protease D3

500

40 after 2 h

Kodera & Nio 2006

Soybean

Protease D3

1,000

50 after 2 h

Kodera & Nio 2006

Jellysh

Pepsin + Papain

200

23 after 2 h

Liu et al. 2012

Jellysh

Pepsin + Papain

400

28 after 4 h

Liu et al. 2012

Jellysh

Pepsin + Papain

800

31 after 2 h

Liu et al. 2012

Shrimp

Pepsin

300

17 after 4 h

Zhang et al. 2009

Shrimp

Pepsin

600

27 after 6 h

Zhang et al. 2009

Shrimp

Pepsin

900

35 after 6 h

Zhang et al. 2009

Bovine casein

A combination of 3 enzymes

20 after 2 h

Yamada et al. 2013

Bovine casein

A combination of 3 enzymes

10

28 after 2 h

Yamada et al. 2013


(Continued )

www.annualreviews.org Antihypertensive Peptides

249

FO06CH11-Aluko

ARI

14 March 2015

12:22

Table 1 (Continued )

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Protein source

SBP reduction
(mmHg)a

Protease(s)

Dose (mg/kg bw)

Bovine casein

A combination of 3 enzymes

100

Bovine casein

Yeast (Debaryomyces hansenii )

200

20 after 2 h

Garca-Tejedor et al. 2013

Bovine casein

Yeast (Kluyveromyces lactis)

200

12 after 1 h

Garca-Tejedor et al. 2013

Bovine casein

Yeast (Kluyveromyces
marxianus)

200

20 after 2 h

Garca-Tejedor et al. 2013

Bovine
lactoferrin

Yeast (Debaryomyces hansenii )

200

18 after 1 h

Garca-Tejedor et al. 2013

Bovine
lactoferrin

Yeast (Kluyveromyces lactis)

200

10 after 1 h

Garca-Tejedor et al. 2013

Bovine
lactoferrin

Yeast (Kluyveromyces
marxianus)

200

23 after 1 h

Garca-Tejedor et al. 2013

Dairy whey

Alcalase

240

32 after 8 h

Wang et al. 2012

Silk broin

Alcalase

100

11 after 4 h

Zhou et al. 2010

Silk broin

Alcalase

600

20 after 4 h

Zhou et al. 2010

Silk broin

Alcalase

1,200

42 after 4 h

Zhou et al. 2010

Hemp seed

Pepsin + Pancreatin

200

30 after 8 h

Girgih et al. 2011b

Flaxseed

Thermoase

200

29 after 4 h

Nwachukwu et al. 2014

42 after 2 h

References
Yamada et al. 2013

Yellow eld pea

Thermolysin

200

19 after 4 h

Li et al. 2011

Flaxseed

Trypsin + Pronase

200

18 after 2 h

Udenigwe et al. 2012a

Palmaria palmata

Papain

50

34 after 2 h

Fitzgerald et al. 2014

Some values are estimates obtained from graphs in the respective publications.

of ACE activity, as shown by the inhibition of angiotensin I-evoked contraction of isolated rat
aortic rings. The lack of contraction in the presence of the thermolysin gelatin peptides indicates
that ACE molecules were blocked from converting angiotensin I into angiotensin II, the latter
being the contractile agent.
Peptide structural characteristics such as type and position of specic amino acids on the peptide chain have been shown to be important for ACE or renin inhibition and ultimately blood
pressurereducing effects. Pistachio protein hydrolysates produced through pepsin, trypsin, or
pepsin + trypsin hydrolyses were shown to reduce blood pressure in SHRs after oral administration; however, the effect was stronger for the trypsin and pepsin + trypsin hydrolysates with a
22-mmHg decrease in SBP when compared to the maximum decrease of 10 mmHg for the pepsin
hydrolysate (Li et al. 2014). The pepsin + trypsin hydrolysate was shown to contain ACKEP as
the most active peptide. ACKEP shares some structural similarity with antihypertensive drugs
such as Lisinopril and Enalapril, as they each contain proline at the C-terminal. Thus, the blood
pressurereducing ability of pepsin + trypsin hydrolysates may be attributed to the presence of
proline, which enhanced binding of the peptides to the ACE active site and resulted in substrate
binding blocking.

6.2. Purified Peptides


Using consecutive HPLC separation techniques, it is possible to obtain homogeneous peptide
preparations that can be subjected to amino acid sequence determination. These peptides can be
250

Aluko

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

ARI

14 March 2015

12:22

either synthesized or isolated during the HPLC separations and then used to determine blood
pressurelowering effects in SHRs. Table 2 shows some of the food protein-derived peptides that
have been identied and tested for in vivo efcacy. With a few exceptions, the results conrm that
shorter-chain peptides could provide better antihypertensive effects than the longer-chain peptides; this may be due to faster rates of absorption and resistance to physiological clearance (longer
persistence in the body). As shown for IRW, the antihypertensive effect may also be dose dependent. One of the early reports was the blood pressurereducing effect of milk tripeptides (VPP
and IPP), which were shown to be active components of sour milk (Table 2). Since then, various
peptide doses have been reported to be effective blood pressurereducing agents. For example,
Kontani et al. (2014) showed that the peptide IHRF reduced SHR blood pressure up to a maximum
of approximately 18 mmHg at 5 mg/kg bw 7 h after oral administration, whereas a 15-mg/kg bw
dose led to a 39-mmHg decrease (Table 2). A recent report showed 3, 7, and 10 mg/kg bw peptides produced dose-dependent decreases in SBP for two lactoferrin-derived peptides (DPYKLRP
and LRP) after oral administration to SHRs (Garca-Tejedor et al. 2014). The mechanism of
action was conrmed to be through in vivo inhibition of plasma ACE activity, which decreased
in SHRs that were given the two peptides. Various casein-derived peptides at 410 mg/kg bw
were also shown to be effective antihypertensive peptides when tested in SHRs; results suggested
that shorter peptides were more effective than longer peptides, as shown in Table 2 (Miguel
et al. 2010). This may be due to the possibility that longer peptides suffer structural degradation
in the GIT, leading to absorption of fragments with lower antihypertensive potency. For some
peptides, there seems to be a direct relationship between the measured in vitro activity and blood
pressurereducing effects in SHRs, as shown in Table 2 for the spinach Rubisco peptides. Peptides
VNP and VWP isolated from rice protein showed, however, very similar in vitro ACE-inhibitory
activities, but VWP produced a greater decrease in SBP (38 mmHg) when compared to VNP
(29 mmHg), also shown in Table 2. Thus, it is possible that ACE inhibition alone may not be
responsible for the blood pressurelowering effects of peptides; other in vivo effects such as renin
inhibition and enhanced NO production could be contributing to the observed antihypertensive
results.
Although short-chain peptides have been proposed as being more effective blood pressure
reducing agents, a tuna dark musclederived peptide containing 13 amino acids (WPEAAELMMEVDP) was shown to reduce SHR SBP by 17 mmHg (Qian et al. 2007). A 21-amino acid
residue peptide (GDLGKTTTVSNWSPPKWKDTP) was isolated from tuna frame protein and
also shown to reduce SBP in SHRs (Lee et al. 2010). In vitro analysis showed both peptides
inhibited ACE activity in a noncompetitive manner, which suggests binding to nonactive sites.
A recent report also showed that a tridecapeptide (IRLIIVLMPILMA) obtained from a papain
digest of the red seaweed (Palmaria palmata) protein had SBP-reducing effects (up to 34 mmHg)
in SHR (Fitzgerald et al. 2014). However, due to the long chain length it is possible that the active forms of these peptides are shorter sequences produced through proteolytic cleavage during
passage through the GIT or during in vivo interactions with ACE or renin. For example, in vitro
simulated gastrointestinal digestion of IRLIIVLMPILMA resulted in two peptide products, IR
and LIIVLMPILMA (Fitzgerald et al. 2014). Similarly, the concentration of peptide VKKVLGNP (isolated from porcine skeletal muscle peptic hydrolysate) decreased with time during in
vitro incubation with ACE (Katayama et al. 2007). However, the ACE-inhibitory activity did not
change during the incubation period, which suggests that VKKVLGNP was converted into active
peptide fragments. In general, the currently available data do not provide denitive information
on type or sequence of amino acids required for effective peptide antihypertensive activity when
provided orally.

www.annualreviews.org Antihypertensive Peptides

251

FO06CH11-Aluko

ARI

14 March 2015

12:22

Table 2 Effects of oral administration of food protein-derived peptides on systolic blood pressure (SBP) of spontaneously
hypertensive rats

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Max. SBP
Protease(s)

Dose
(mg/kg bw)

reduction
(mmHg)a

Peptide sequence

Protein source

IHRF

Rice glutelin

Chymotrypsin

18

Kontani et al. 2014

References

IHRF

Rice glutelin

Chymotrypsin

15

39

Kontani et al. 2014

DPYKLRP

Lactoferrin

Yeast extract

10

27

Garca-Tejedor et al. 2014

PYKLRP

Lactoferrin

Yeast extract

10

21

Garca-Tejedor et al. 2014

YKLRP

Lactoferrin

Yeast extract

10

20

Garca-Tejedor et al. 2014

KLRP

Lactoferrin

Yeast extract

10

12

Garca-Tejedor et al. 2014

LRP

Lactoferrin

Yeast extract

10

27

Garca-Tejedor et al. 2014

GILRP

Lactoferrin

Yeast extract

10

20

Garca-Tejedor et al. 2014

IAK

Casein

Synthesized

21

Miguel et al. 2010

HLPLP

Casein

Synthesized

24

Miguel et al. 2010

YAKPVA

Casein

Synthesized

23

Miguel et al. 2010

HPHPHLSF

Casein

Synthesized

10

16

Miguel et al. 2010

KKYNVPQL

Casein

Synthesized

10

12

Miguel et al. 2010

LVYPFTGPIPN

Casein

Synthesized

10

28

Miguel et al. 2010

WQVLPNAVPAK

Casein

Synthesized

18

Miguel et al. 2010

VPP

Sour milk

Fermentation

20

Nakamura et al. 1995

IPP

Sour milk

Fermentation

18

Nakamura et al. 1995

MRW

Spinach Rubisco

Pepsin + Pancreatin

30

20

Yang et al. 2003

MRWRD

Spinach Rubisco

Pepsin + Pancreatin

30

14

Yang et al. 2003

IAYKPAG

Spinach Rubisco

Pepsin + Pancreatin

100

15

Yang et al. 2003

DY

Bamboo shoot

None (aqueous extract)

10

18

Liu et al. 2013

AAATP

Spanish ham

None (aqueous extract)

26

Escudero et al. 2013

VNP

Rice

Alcalase + Trypsin

29

Chen et al. 2013a

VWP

Rice

Alcalase + Trypsin

38

Chen et al. 2013a

IRW

Egg
ovotransferrin

Thermolysin + Pepsin

10

Majumder et al. 2013

IRW

Egg
ovotransferrin

Thermolysin + Pepsin

15

22

Majumder et al. 2013

AVF

Insect

Pepsin + Trypsin +
Chymotrypsin

13

Vercruysse et al. 2010

VF

Insect

Pepsin + Trypsin +
Chymotrypsin

19

Vercruysse et al. 2010

GQP

Mushroom

None (aqueous extract)

27

Lee et al. 2004

DKVGINYW

Dairy whey

Synthesized

15

Tavares et al. 2012

DAQSAPLRVY

Dairy whey

Synthesized

10

Tavares et al. 2012

KGYGGVSLPEW

Dairy whey

Synthesized

20

Tavares et al. 2012

YFP

Yellown sole

Chymotrypsin

10

22

Jung et al. 2006

IY

Rapeseed

Alcalase

7.5

10

Marczak et al. 2003

VW

Rapeseed

Alcalase

7.5

11

Marczak et al. 2003

RIY

Rapeseed

Alcalase

7.5

11

Marczak et al. 2003

VWIS

Rapeseed

Alcalase

12.5

13

Marczak et al. 2003


(Continued )

252

Aluko

FO06CH11-Aluko

ARI

14 March 2015

12:22

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Table 2 (Continued )

Protease(s)

Dose

Max. SBP
reduction

(mg/kg bw)

(mmHg)a

Peptide sequence

Protein source

VKKVLGNP

Porcine muscle

Pepsin

10

24

Katayama et al. 2007

References

RPR

Pork meat

Synthesized

33

Escudero et al. 2012b

PTPVP

Pork meat

Synthesized

26

Escudero et al. 2012b

KAPVA

Pork meat

Synthesized

34

Escudero et al. 2012b

KRVIQY

Porcine myosin

Pepsin

10

23

Muguruma et al. 2009

VKAGF

Porcine myosin

Pepsin

10

17

Muguruma et al. 2009

MKP

Bovine casein

Alcalase + Trypsin +
Protease N

0.1

35

Yamada et al. 2013

VEGY

Microalgae

Alcalase

10

23

Ko et al. 2012a

YH

Edible seaweed

Synthesized

50

50

Suetsuna et al. 2004

KY

Edible seaweed

Synthesized

50

45

Suetsuna et al. 2004

WY

Edible seaweed

Synthesized

50

46

Suetsuna et al. 2004

IY

Edible seaweed

Synthesized

50

33

Suetsuna et al. 2004

FQ

Buckwheat
sprout

Synthesized

0.1

42

Koyama et al. 2013

VAE

Buckwheat
sprout

Synthesized

0.1

38

Koyama et al. 2013

VVG

Buckwheat
sprout

Synthesized

0.1

38

Koyama et al. 2013

DVWY

Buckwheat
sprout

Synthesized

0.1

59

Koyama et al. 2013

FDART

Buckwheat
sprout

Synthesized

0.1

32

Koyama et al. 2013

WTFR

Buckwheat
sprout

Synthesized

0.1

36

Koyama et al. 2013

MAW

Cuttlesh

Synthesized

10

13

Balti et al. 2012

AHSY

Cuttlesh

Synthesized

10

14

Balti et al. 2012

VYAP

Cuttlesh

Synthesized

10

22

Balti et al. 2012

VIIF

Cuttlesh

Synthesized

10

19

Balti et al. 2012

WYT

Hemp seed

Pepsin + Pancreatin

30

13

Girgih et al. 2014c

WVYY

Hemp seed

Pepsin + Pancreatin

30

34

Girgih et al. 2014c

SVYT

Hemp seed

Pepsin + Pancreatin

30

24

Girgih et al. 2014c

PSLPA

Hemp seed

Pepsin + Pancreatin

30

40

Girgih et al. 2014c

IPAGV

Hemp seed

Pepsin + Pancreatin

30

36

Girgih et al. 2014c

LY

Rapeseed

Alcalase

30

26

He et al. 2013b

TF

Rapeseed

Alcalase

30

12

He et al. 2013b

RALP

Rapeseed

Alcalase

30

16

He et al. 2013b

GHS

Rapeseed

Pepsin + Pancreatin

30

17

He et al. 2013c

WMP

Yellow eld pea

Thermolysin

30

39

Aluko et al. 2014

ADMFPF

Yellow eld pea

Thermolysin

30

25

Aluko et al. 2014

IRLIIVLMPILMA

Palmaria palmata

Papain

50

33

Fitzgerald et al. 2014

Some values are estimates obtained from graphs in the respective publications.

www.annualreviews.org Antihypertensive Peptides

253

FO06CH11-Aluko

ARI

14 March 2015

12:22

7. FUTURE PERSPECTIVES

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Research into production, characterization, and efcacy testing of antihypertensive peptides continues to evolve rapidly. Current efforts include the use of in silico methods to perform virtual
enzyme hydrolysis of proteins for which the released peptide sequences can be synthesized and
tested for activity. If in vitro activity (usually against renin and ACE) is conrmed, actual enzyme
hydrolysis can be performed to obtain peptides that can then undergo in vitro and in vivo testing.
This in silico approach could obviate the current time-consuming (and expensive) hit-and-miss
approach whereby various enzymes are thrown at proteins followed by testing of each hydrolysate
to identify active products. However, availability of the primary structure of several more proteins
is required to expand the in silico method beyond the currently limited structurally characterized food proteins. Additional information on structure-function properties of antihypertensive
peptides is also needed to develop better enzyme digestion tools that will release peptides with
desirable amino acid sequences. Finally, more human intervention trials are necessary to provide efcacy data that meet current regulatory requirements in order to allow a health claim of
antihypertensive activity on peptide-formulated food products or nutraceuticals.

DISCLOSURE STATEMENT
The author is not aware of any afliations, memberships, funding, or nancial holdings that might
be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS
The research program of Dr. Aluko is funded by the Natural Sciences and Engineering Research
Council of Canada (NSERC) and the Manitoba Agri-Food Research & Development Initiative
(ARDI).

LITERATURE CITED
Abassi Z, Winaver J, Feuerstein GZ. 2009. The biochemical pharmacology of renin inhibitors: implications for
translational medicine in hypertension, diabetic nephropathy and heart failure: expectations and reality.
Biochem. Pharmacol. 78:93340
Acharya KR, Sturrock ED, Riordan JF, Ehlers MRW. 2003. ACE revisited: a new target for structure-based
drug design. Nat. Rev. 2:891902
Ahn CB, Jeon Y-J, Kim Y-T, Je J-Y. 2012. Angiotensin I converting enzyme (ACE) inhibitory peptides from
salmon byproduct protein hydrolysate by alcalase hydrolysis. Process Biochem. 47:224045
Ahn JE, Park SY, Atwal A, Gibbs BF, Lee BH. 2009. Angiotensin I-converting enzyme (ACE) inhibitory
peptides from whey fermented by Lactobacillus species. J. Food Biochem. 33:587602
Ajibola CF, Fashakin JB, Fagbemi TN, Aluko RE. 2013. Renin and angiotensin converting enzyme inhibition
with antioxidant properties of African yam bean protein hydrolysate and reverse-phase HPLC-separated
peptide fractions. Food Res. Int. 52:43744
Alashi AM, Blanchard CL, Mailer RJ, Agboola SO, Mawson AJ, et al. 2014. Blood pressure lowering effects
of Australian canola protein hydrolysates in spontaneously hypertensive rats. Food Res. Int. 55:28187
Aleman A, Gomez-Guillen MC, Montero P. 2013. Identication of ace-inhibitory peptides from squid collagen
after in vitro gastrointestinal digestion. Food Res. Int. 54:79095
Aluko RE, Wu J, Aukema HM. 2014. Yellow eld pea seed protein-derived peptides. US Patent No. 8,815,806
Antonios TFT, MacGregor GA. 1995. Angiotensin converting enzyme inhibitors in hypertension: potential
problems. J. Hypertens. 13(Suppl. 3):S1116
254

Aluko

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

ARI

14 March 2015

12:22

Balti R, Bougatef A, Guillochon D, Dhulster P, Nasri M, Nedjar-Arroume N. 2012. Changes in arterial


blood pressure after single oral administration of cuttlesh (Sepia ofcinalis) muscle derived peptides in
spontaneously hypertensive rats. J. Funct. Foods 4:61117
Balti R, Nedjar-Arroume N, Bougatef A, Guillochon D, Nasri M. 2010. Three novel angiotensin I-converting
enzyme (ACE) inhibitory peptides from cuttlesh (Sepia ofcinalis) using digestive proteases. Food Res. Int.
43:113643
Banerjee P, Shanthi C. 2012. Isolation of novel bioactive regions from bovine Achilles tendon collagen having
angiotensin I-converting enzyme inhibitory properties. Process Biochem. 47:233546
Boschin G, Scigliuolo GM, Resta D, Arnoldi A. 2014. Optimization of the enzymatic hydrolysis of lupin
(Lupinus) proteins for producing ACE-inhibitory peptides. J. Agric. Food Chem. 62:184651
Bremmer AD. 2003. Antihypertensive medication and quality of lifesilent treatment of a silent killer?
Cardiovasc. Drugs Ther. 16:35364
Chen J, Liu S, Ye R, Cai G, Ji B, Wu Y. 2013a. Angiotensin-I converting enzyme (ACE) inhibitory tripeptides
from rice protein hydrolysate: purication and characterization. J. Funct. Foods 5:168492
Chen J, Wang Y, Ye R, Wua Y, Xia W. 2013b. Comparison of analytical methods to assay inhibitors of
angiotensin I-converting enzyme. Food Chem. 141:332934
Chou C-L, Pang C-Y, Lee TJF, Fang T-C. 2013. Direct renin inhibitor prevents and ameliorates insulin resistance, aortic endothelial dysfunction and vascular remodeling in fructose-fed hypertensive rats. Hypertens.
Res. 36:12328
Conradi RA, Wilkinson KF, Rush BD, Hilgers AR, Ruwart MJ, Burton PS. 1993. In-vitro/in-vivo models for
peptide oral absorption: comparison of Caco-2 cell permeability with rat intestinal absorption of renin
inhibitory peptides. Pharm. Res. 10:179092
Curtis KK, Li P, Jackson EK. 1995. Blood pressure after captopril withdrawal from spontaneously hypertensive
rats. Hypertension 25:8287
Cushman DW, Cheung HS. 1971. Spectrophotometric assay and properties of the angiotensin-converting
enzyme of rabbit lung. Biochem. Pharmacol. 20:163748
de Gobba C, Tompa G, Otte J. 2014. Bioactive peptides from caseins released by cold active proteolytic
enzymes from Arsukibacterium ikkense. Food Chem. 165:20515
del Castillo MD, Ferrigno A, Acampa I, Borrelli RC, Olano A, et al. 2007. In vitro release of angiotensinconverting enzyme inhibitors, peroxyl-radical scavengers and antibacterial compounds by enzymatic hydrolysis of glycated gluten. J. Cereal Sci. 45:32734
Du L, Fang M, Wu H, Xie J, Wu Y, et al. 2013. A novel angiotensin I-converting enzyme inhibitory peptide
from Phascolosoma esculenta water-soluble protein hydrolysate. J. Funct. Foods 5:47583
Escudero E, Aristoy MC, Hitoshi N, Arihara K, Toldra F. 2012a. Antihypertensive effect and antioxidant
activity of peptide fractions extracted from Spanish dry-cure ham. Meat Sci. 91:30611
Escudero E, Mora L, Fraser PD, Aristoy MC, Arihara K, Toldra F. 2013. Purication and identication of
antihypertensive peptides in Spanish dry-cure ham. J. Proteomics 78:499507
Escudero E, Mora L, Toldra F. 2014. Stability of ACE inhibitory ham peptides against heat treatment and in
vitro digestion. Food Chem. 161:30511
Escudero E, Toldra F, Sentandreu MA, Nishimura H, Arihara K. 2012b. Antihypertensive activity of peptides
identied in the in vitro gastrointestinal digest of pork meat. Meat Sci. 91:38284
Ewart HS, Dennis D, Potvin M, Tiller C, Fang L-H, et al. 2009. Development of a salmon protein hydrolysate
that lowers blood pressure. Eur. Food Res. Technol. 229:56169
Feld LG, Van Liew B, Brentjens JR, Boylan JW. 1981. Renal lesions and proteinuria in the spontaneously
hypertensive rats made normotensive by treatment. Kidney Int. 20:60614
Fernandez-Musoles R, Manzanares P, Burguete MC, Alborch E, Salom JB. 2013. In vivo angiotensin Iconverting enzyme inhibition by long-term intake of antihypertensive lactoferrin hydrolysate in spontaneously hypertensive rats. Food Res. Int. 54:62732
Fitzgerald C, Aluko RE, Hossain M, Rai DK, Hayes M. 2014. The potential of a renin inhibitory peptide from
the red seaweed Palmaria palmata as functional food ingredient following conrmation and characterization of a hypotensive effect in spontaneously hypertensive rats (SHRs). J. Agric. Food Chem. 62:835256
www.annualreviews.org Antihypertensive Peptides

255

ARI

14 March 2015

12:22

Fitzgerald C, Mora-Soler L, Gallagher E, OConnor P, Prieto J, et al. 2012. Isolation and characterization
of bioactive pro-peptides with in vitro renin inhibitory activities from the macroalga Palmaria palmata. J.
Agric. Food Chem. 60:742127
Flack JM, Novikov SV, Ferrario CM. 1997. Benets of adherence to anti-hypertensive therapy. Blood Press.
1(Suppl.):4751
Fujita H, Yamagami T, Ohshima K. 2001. Effects of an ace-inhibitory agent, katsuobushi oligopeptide, in the
spontaneously hypertensive rat and in borderline and mildly hypertensive subjects. Nutr. Res. 21:114958
Fujita H, Yokoyama K, Yoshikawa M. 2000. Classication and antihypertensive activity of angiotensin Iconverting enzyme inhibitory peptides derived from food proteins. J. Food Sci. 65:56469
Garcia-Mora P, Penas E, Frias J, Martinez-Villaluenga C. 2014. Savinase, the most suitable enzyme for
releasing peptides from lentil (Lens culinaris var. Castellana) protein concentrates with multifunctional
properties. J. Agric. Food Chem. 62:416674
Garca-Tejedor A, Padilla B, Salom JB, Belloch C, Manzanares P. 2013. Dairy yeasts produce milk proteinderived antihypertensive hydrolysates. Food Res. Int. 53:2038
Garca-Tejedor A, Sanchez-Rivera L, Castello-Ruiz
M, Recio I, Salom JB, Manzanares P. 2014. Novel anti
hypertensive lactoferrin-derived peptides produced by Kluyveromyces marxianus: gastrointestinal stability
prole and in vivo angiotensin I-converting enzyme (ACE) inhibition. J. Agric. Food Chem. 62:160916
Ghassem M, Babji AS, Said M, Mahmoodani F, Arihara K. 2014. Angiotensin I-converting enzyme inhibitory
peptides from snakehead sh sarcoplasmic protein hydrolysate. J. Food Biochem. 38:14049
Girgih AT, Alashi AM, He R, Malomo SA, Aluko RE. 2014a. Preventive and treatment effects of hemp seed
(Cannabis sativa L.) meal protein hydrolysate against high blood pressure in spontaneously hypertensive
rats. Eur. J. Nutr. 53:123746
Girgih AT, He R, Aluko RE. 2014b. Kinetics and molecular docking studies of the inhibitions of angiotensin
converting enzyme and renin activities by hemp seed (Cannabis sativa L.) peptides. J. Agric. Food Chem.
62:413544
Girgih AT, He R, Malomo SA, Offengenden M, Wu J, Aluko RE. 2014c. Structural and functional characterization of hemp seed (Cannabis sativa L.) protein-derived antioxidant and antihypertensive peptides.
J. Funct. Foods 6:38494
Girgih AT, Udenigwe CC, Aluko RE. 2011a. In vitro antioxidant properties of hempseed (Cannabis sativa L.)
protein hydrolysate fractions. J. Am. Oil Chem. Soc. 88:38189
Girgih AT, Udenigwe CC, Li H, Adebiyi AP, Aluko RE. 2011b. Kinetics of enzyme inhibition and antihypertensive effects of hemp seed (Cannabis sativa L.) protein hydrolysates. J. Am. Oil Chem. Soc. 88:176774
Gu Y, Wu J. 2013. LC-MS/MS coupled with QSAR modeling in characterizing of angiotensin I-converting
enzyme inhibitory peptides from soybean proteins. Food Chem. 141:268290
Gunkel AR, Thurner KH, Kanonier G, Sprinzl GM, Thumfart WF. 1996. Angioneurotic edema as a reaction
to angiotensin-converting enzyme inhibitors. Am. J. Otolaryngol. 17:8791
Hayes M, Stanton C, Slattery H, OSullivan O, Hill C, et al. 2007. Casein fermentate of Lactobacillus animalis
DPC6134 contains a range of novel propeptide angiotensin-converting enzyme inhibitors. Appl. Environ.
Microbiol. 73:465867
He R, Alashi A, Malomo SA, Girgih AT, Chao D, et al. 2013a. Antihypertensive and free radical scavenging
properties of enzymatic rapeseed protein hydrolysates. Food Chem. 141:15359
He R, Aluko RE, Ju X. 2014. Evaluating molecular mechanism of hypotensive peptides interactions with renin
and angiotensin converting enzyme. PLoS ONE 9(3):e91051
He R, Malomo SA, Alashi A, Girgih AT, Ju X, Aluko RE. 2013b. Purication and hypotensive activity of
rapeseed protein-derived renin and angiotensin converting enzyme inhibitory peptides. J. Funct. Foods
5:78189
He R, Malomo SA, Girgih AT, Ju X, Aluko RE. 2013c. Glycinyl-histidinyl-serine (GHS), a novel rapeseed
protein-derived peptide has blood pressure-lowering effect in spontaneously hypertensive rats. J. Agric.
Food Chem. 61:8396402
Hernandez-Ledesma B, Miralles B, Amigo L, Ramos M, Recio I. 2005. Identication of antioxidant and
ACE-inhibitory peptides in fermented milk. J. Sci. Food Agric. 85:104148

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

256

Aluko

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

ARI

14 March 2015

12:22

Herregods G, van Camp J, Morel N, Ghesqui`ere B, Gevaert K, et al. 2011. Angiotensin I-converting enzyme
inhibitory activity of gelatin hydrolysates and identication of bioactive peptides. J. Agric. Food Chem.
59:55258
Hirota T, Nonaka A, Matsushita A, Uchida N, Ohki K, et al. 2011. Milk casein-derived tripeptides, VPP
and IPP induced NO production in cultured endothelial cells and endothelium-dependent relaxation of
isolated aortic rings. Heart Vessels 26:54956
Hiwatashi K, Shirakawa H, Hori K, Yoshiki Y, Suzuki N, et al. 2010. Reduction of blood pressure by soybean
saponins, renin inhibitors from soybean, in spontaneously hypertensive rats. Biosci. Biotechnol. Biochem.
74:231012
Holmquist B, Bunning P, Riordan JF. 1979. A continuous spectrophotometric assay for angiotensin converting
enzyme. Anal. Biochem. 95:54048
Huang W-H, Sun J, He H, Dong H-W, Li J-T. 2011. Antihypertensive effect of corn peptides, produced by
a continuous production in enzymatic membrane reactor, in spontaneously hypertensive rats. Food Chem.
128:96873
Imai T, Miyazaki H, Hirose S, Hori H, Hayashi T, et al. 1983. Cloning and sequence analysis of cDNA for
human renin precursor. Proc. Natl. Acad. Sci. USA 80:74059
Inoue K, Gotou T, Kitajima H, Mizuno S, Nakazawa T, Yamamoto N. 2009. Release of antihypertensive
peptides in miso paste during its fermentation, by the addition of casein. J. Biosci. Bioeng. 108:11115
Intarasirisawat R, Benjakul S, Wu J, Visessanguan W. 2013. Isolation of antioxidative and ACE inhibitory
peptides from protein hydrolysate of skipjack (Katsuwana pelamis) roe. J. Funct. Foods 5:185462
Ishida Y, Shibata Y, Fukuhara I, Yano Y, Takehara I, Kaneko K. 2011. Effect of an excess intake of casein
hydrolysate containing Val-Pro-Pro and Ile-Pro-Pro in subjects with normal blood pressure, high-normal
blood pressure, or mild hypertension. Biosci. Biotechnol. Biochem. 75:42733
Jakubczyk A, Karas M, Baraniak B, Pietrzak M. 2013. The impact of fermentation and in vitro digestion
on formation angiotensin converting enzyme (ACE) inhibitory peptides from pea proteins. Food Chem.
141:377480
Jang A, Lee M. 2005. Purication and identication of angiotensin converting enzyme inhibitory peptides
from beef hydrolysates. Meat Sci. 69:65361
Jang J-H, Jeong S-C, Kim J-H, Lee Y-H, Ju Y-C, Lee J-S. 2011. Characterisation of a new antihypertensive
angiotensin I-converting enzyme inhibitory peptide from Pleurotus cornucopiae. Food Chem. 127:41218
Jimsheena VK, Gowda LR. 2010. Arachin derived peptides as selective angiotensin I-converting enzyme
(ACE) inhibitors: Structure-activity relationship. Peptides 31:116576
Jung W-K, Mendis E, Je J-Y, Park P-J, Son BW, et al. 2006. Angiotensin I-converting enzyme inhibitory
peptide from yellown sole (Limanda aspera) frame protein and its antihypertensive effect in spontaneously
hypertensive hypertensive rats. Food Chem. 94:2632
Katayama K, Mori T, Kawahara S, Miake K, Kodama Y, et al. 2007. Angiotensin converting enzyme inhibitory
peptide derived from porcine skeletal muscle myosin and its antihypertensive activity in spontaneously
hypertensive rats. J. Food Sci. 72:S7026
Kazuki N, Yoshie-Stark Y, Ogushi M. 2009. Comparison of ACE-inhibitory and DPPH radical scavenging
activities of sh muscle hydrolysates. Food Chem. 114:84451
Ko S-C, Kang N, Kim E-A, Kang MC, Lee S-H, et al. 2012a. A novel angiotensin I-converting enzyme (ACE)
inhibitory peptide from a marine Chlorella ellipsoidea and its antihypertensive effect in spontaneously
hypertensive rats. Process Biochem. 47:200511
Ko S-C, Kim DG, Han C-H, Lee YJ, Lee J-K, et al. 2012b. Nitric oxide-mediated vasorelaxation effects of antiangiotensin I-converting enzyme (ACE) peptide from Styela clava esh tissue and its anti-hypertensive
effect in spontaneously hypertensive rats. Food Chem. 134:114145
Kodera T, Nio N. 2006. Identication of an angiotensin I-converting enzyme inhibitory peptides from protein
hydrolysates by a soybean protease and the antihypertensive effects of hydrolysates in spontaneously
hypertensive rats. J. Food Sci. 71:C16473
Kontani N, Omae R, Kagebayashi T, Kaneko K, Yamada Y, et al. 2014. Characterization of Ile-His-Arg-Phe,
a novel rice-derived vasorelaxing peptide with hypotensive and anorexigenic activities. Mol. Nutr. Food
Res. 58:359364
www.annualreviews.org Antihypertensive Peptides

257

ARI

14 March 2015

12:22

Koyama M, Naramoto K, Nakajima T, Aoyama T, Watanabe M, Nakamura K. 2013. Purication and identication of antihypertensive peptides from fermented buckwheat sprouts. J. Agric. Food Chem. 61:301321
Ktari N, Nasri R, Mnafgui K, Hamden K, Belguith O, et al. 2014. Antioxidative and ACE inhibitory activities of
protein hydrolysates from zebra blenny (Salaria basilisca) in alloxan-induced diabetic rats. Process Biochem.
49:89097
Kvam F, Ofstad J, Iversen BM. 1998. Effects of antihypertensive drugs on autoregulation of RBF and glomerular capillary pressure in SHR. Am. J. Physiol. 275:F57694
Lahogue V, Rehel K, Taupin L, Haras D, Allaume P. 2010. A HPLC-UV method for the determination of
angiotensin I-converting enzyme (ACE) inhibitory activity. Food Chem. 118:87075
Lee DH, Kim JH, Park JS, Choi YJ, Lee JS. 2004. Isolation and characterization of a novel angiotensin Iconverting enzyme inhibitory peptide derived from the edible mushroom Tricholoma giganteum. Peptides
25:62127
Lee HJ. 2002. Protein drug oral delivery: the recent progress. Arch. Pharm. Res. 25:57284
Lee J-E, Bae IY, Lee HG, Yang C-B. 2006. Tyr-Pro-Lys, an angiotensin I-converting enzyme inhibitory
peptide derived from broccoli (Brassica oleracea Italica). Food Chem. 99:14348
Lee S-H, Qian Z-J, Kim S-K. 2010. A novel angiotensin I converting enzyme inhibitory peptide from tuna
frame protein hydrolysate and its antihypertensive effect in spontaneously hypertensive rats. Food Chem.
118:96102
Lee S-J, Kim Y-S, Kim S-E, Kim E-K, Hwang J-W, et al. 2012. Purication and characterization of a novel
angiotensin I-converting enzyme inhibitory peptide derived from an enzymatic hydrolysate of duck skin
byproducts. J. Agric. Food Chem. 60:1003540
Li G-H, Shi Y-H, Liu H, Le G-W. 2006. Antihypertensive effect of alcalase generated mung bean protein
hydrolysates in spontaneously hypertensive rats. Eur. Food Res. Technol. 222:73336
Li H, Aluko RE. 2010. Identication and inhibitory properties of multifunctional peptides from pea protein
hydrolysate. J. Agric. Food Chem. 58:1147176
Li H, Prairie N, Udenigwe CC, Adebiyi AP, Tappia P, et al. 2011. Blood pressure lowering effect of a pea
protein hydrolysate in hypertensive rats and humans. J. Agric. Food Chem. 59:985460
Li P, Jia J, Fang M, Zhang L, Guo M, et al. 2014. In vitro and in vivo ACE inhibitory of pistachio hydrolysates
and in silico mechanism of identied peptide binding with ACE. Process Biochem. 49:898904
Lin L, Lv S, Li B. 2012. Angiotensin-I-converting enzyme (ACE)-inhibitory and antihypertensive properties
of squid skin gelatin hydrolysates. Food Chem. 131:22530
Liu L, Liu L, Lu B, Chen M, Zhang Y. 2013. Evaluation of bamboo shoot peptide preparation with angiotensin
converting enzyme inhibitory and antioxidant abilities from byproducts of canned bamboo shoots. J. Agric.
Food Chem. 61:552633
Liu X, Zhang M, Zhang C, Liu C. 2012. Angiotensin converting enzyme (ACE) inhibitory, antihypertensive and antihyperlipidaemic activities of protein hydrolysates from Rhopilema esculentum. Food Chem.
134:213440
Lu J, Sawano Y, Miyakawa T, Xue Y-L, Cai M-Y, et al. 2011. One-week antihypertensive effect of Ile-Gln-Pro
in spontaneously hypertensive rats. J. Agric. Food Chem. 59:55963
Majumder K, Chakrabarti S, Morton JS, Panahi S, Kaufman S, et al. 2013. Egg-derived tri-peptide IRW
exerts antihypertensive effects in spontaneously hypertensive rats. PLoS ONE 8:e82829
Majumder K, Wu J. 2009. Angiotensin I converting enzyme inhibitory peptides from simulated in vitro
gastrointestinal digestion of cooked eggs. J. Agric. Food Chem. 57:47177
Makinen S, Johannson T, Gerd EV, Pihlava JM, Pihlanto A. 2012. Angiotensin I-converting enzyme inhibitory
and antioxidant properties of rapeseed hydrolysates. J. Funct. Foods 4:57583
Marczak ED, Usui H, Fujita H, Yang Y, Yokoo M, et al. 2003. New antihypertensive peptides isolated from
rapeseed. Peptides 24:79198
Marrufo-Estrada DM, Segura-Campos MR, Chel-Guerrero LA, Betancur-Ancona DA. 2013. Defatted
Jatropha curcus our and protein isolate as materials for protein hydrolysates with biological activity.
Food Chem. 138:7783
Masuda O, Nakamura Y, Takano T. 1996. Antihypertensive peptides are present in aorta after oral administration of sour milk containing these peptides to spontaneously hypertensive rats. J. Nutr. 126:306368

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

258

Aluko

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

ARI

14 March 2015

12:22

Matsui T, Imamura M, Oka H, Osajima K, Kimoto K-I, et al. 2004. Tissue distribution of antihypertensive
dipeptide, Val-Tyr, after its single oral administration to spontaneously hypertensive rats. J. Pept. Sci.
10:53545
Memarpoor-Yazdi M, Asoodeh A, Chamani JK. 2012. Structure and ACE-inhibitory activity of peptides
derived from hen egg white lysozyme. Int. J. Pept. Res. Ther. 18:35360
Miguel M, Gomez-Ruiz JA, Recio I, Aleixandre A. 2010. Changes in arterial blood pressure after single oral
administration of milk-casein-derived peptides in spontaneously hypertensive rats. Mol. Nutr. Food Res.
54:142227
Morais HA, Silvestre MPC, Amorin LL, Silva VDM, Silva MR, et al. 2014. Use of different proteases to obtain
whey protein concentrate hydrolysates with inhibitory activity toward angiotensin-converting enzyme.
J. Food Biochem. 38:1029
Muguruma M, Ahhmed AM, Katayama K, Kawahara S, Maruyama M, Nakamura T. 2009. Identication of
pro-drug type ACE inhibitory peptide sourced from porcine myosin B: evaluation of its antihypertensive
effects in vivo. Food Chem. 114:51622
Mundi S, Aluko RE. 2014. Inhibitory properties of kidney bean protein hydrolysate and its membrane fractions
against renin, angiotensin converting enzyme, and free radicals. Austin J. Nutr. Food Sci. 2(1):11
Murakami M, Tonouchi H, Takahashi R, Kitazawa H, Kawai Y, et al. 2004. Structural analysis of a new antihypertensive peptide (-lactosin B) isolated from a commercial whey product. J. Dairy Sci. 87:196774
Nakamura Y, Yamamoto N, Sakai K, Takano T. 1995. Antihypertensive effect of sour milk and peptides
isolated from it that are inhibitors of angiotensin converting enzyme. J. Dairy Sci. 78:125357
Nakano D, Ogura K, Miyakoshi M, Ishii F, Kawanishi H, et al. 2006. Antihypertensive effect of angiotensin Iconverting enzyme inhibitory peptides from a sesame protein hydrolysate in spontaneously hypertensive
rats. Biosci. Biotechnol. Biochem. 70:111826
Nasri R, Younes I, Jridi M, Trigui M, Bougatef A, et al. 2013. ACE inhibitory and antioxidative activities of
Goby (Zosterissessor ophiocephalus) sh protein hydrolysates: effect on meat lipid oxidation. Food Res. Int.
54:55261
Noach A, Hurni MA, De Boer AG, Breimer DD. 1994. The paracellular approach: drug transport and its enhancement via the paracellular pathways. In Drug Absorption Enhancement. Concept, Possibilities, Limitations
and Trends, Vol. 3, ed. AG De Boer, pp. 291324. Chur, Switz: Harwood Acad. Publ.
Norris R, Poyarkov A, OKeeffe MB, FitzGerald RJ. 2014. Characterisation of the hydrolytic specicity of
Aspergillus niger derived prolyl endoprotease on bovine -casein and determination of ACE inhibitory
activity. Food Chem. 156:2936
Nwachukwu ID, Girgih AT, Malomo SA, Onuh J, Aluko RE. 2014. Thermoase-derived axseed protein
hydrolysates and membrane ultraltration peptide fractions have systolic blood pressure-lowering effects
in spontaneously hypertensive rats. Int. J. Mol. Sci. 15:1813147
Onuh JO, Girgih AT, Aluko RE, Aliani M. 2013. Inhibitions of renin and angiotensin converting enzyme
activities by enzymatic chicken skin protein hydrolysates. Food Res. Int. 53:26067
Pappenheimer JR, Michel CC. 2003. Role of villus microcirculation in intestinal absorption of glucose: coupling of epithelial with endothelial transport. J. Physiol. 553:56174
Perez-Vega JA, Olivera-Castillo L, Gomez-Ruiz JA, Hernandez-Ledesma B. 2013. Release of multifunctional
peptides by gastrointestinal digestion of sea cucumber (Isostichopus badionotus). J. Funct. Foods 5:86977
Picot L, Ravallec-Ple R, Fouchereau-Peron M, Vandanjon L, Jaouen P, et al. 2010. Impact of ultraltration
and nanoltration of an industrial sh protein hydrolysate on its bioactive properties. J. Sci. Food Agric.
90:181926
Pihlanto A, Akkanen S, Korhonen HJ. 2008. ACE-inhibitory and antioxidant properties of potato (Solanum
tuberosum). Food Chem. 109:10412
Puchalska P, Garcia MC, Marina ML. 2014. Identication of native angiotensin-I converting enzyme inhibitory peptides in commercial soybean based infant formulas using HPLC-Q-ToF-MS. Food Chem.
157:6269
Qian Z-J, Je J-Y, Kim S-K. 2007. Antihypertensive effect of angiotensin I converting enzyme-inhibitory
peptide from hydrolysates of bigeye tuna dark muscle, Thunnus obesus. J. Agric. Food Chem. 55:8398403
Rao S, Sun J, Liu Y, Zeng H, Su Y, Yang Y. 2012a. ACE inhibitory peptides and antioxidant peptides derived
from in vitro digestion hydrolysate of hen egg white lysozyme. Food Chem. 135:124552
www.annualreviews.org Antihypertensive Peptides

259

ARI

14 March 2015

12:22

Rao S-Q, Ju T, Sun J, Su Y-J, Xu R-R, Yang Y-J. 2012b. Purication and characterization of angiotensin
I-converting enzyme inhibitory peptides from enzymatic hydrolysate of hen egg white lysozyme. Food
Res. Int. 46:12734
Ruiz-Ruiz J, Davila-Ortiz G, Chel-Guerrero L, Betancur-Ancona D. 2013. Angiotensin I-converting enzyme
inhibitory and antioxidant peptide fractions from hard-to-cook bean enzymatic hydrolysates. J. Food
Biochem. 37:2635
Sagardia I, Roa-Ureta RH, Bald C. 2013. A new QSAR model, for angiotensin I-converting enzyme inhibitory
oligopeptides. Food Chem. 136:137076
Sanchez D, Quinones M, Moulay L, Muguerza B, Miguel M, Aleixandre A. 2010. Changes in arterial blood
pressure of a soluble cocoa ber product in spontaneously hypertensive rats. J. Agric. Food Chem. 58:1493
501
Satake M, Enjoh M, Nakamura Y, Takano T, Kawamura Y, et al. 2002. Transepithelial transport of the
bioactive peptides, Val-Pro-Pro, in human intestinal Caco-2 cell monolayers. Biosci. Biotechnol. Biochem.
66:37884
Sato M, Hosokawa T, Yamaguchi T, Nakano T, Muramoto K, et al. 2002. Angiotensin I-converting enzyme inhibitory peptides derived from wakame (Undaria pinnatida) and their antihypertensive effect in
spontaneously hypertensive rats. J. Agric. Food Chem. 50:624552
Seguro Campos MR, Chel Guerrero LA, Betancor Ancona DA. 2010. Angiotensin I-converting enzyme
inhibitory and antioxidant activities of peptide fractions extracted by ultraltration of cowpea Vigna
unguiculata hydrolysates. J. Sci. Food Agric. 90:251218
Semple PF. 1995. Putative mechanisms of cough after treatment with angiotensin converting enzyme inhibitors. J. Hypertens. 13(Suppl. 3):S1721
Shalaby SM, Zakora M, Otte J. 2006. Performance of two commonly used angiotensin-converting enzyme
inhibition assays using FA-PGG and HHL as substrates. J. Dairy Res. 73:17886
Sheih I-C, Fang TJ, Wu T-K. 2009. Isolation and characterisation of a novel angiotensin I-converting enzyme
(ACE) inhibitory peptide from the algae protein waste. Food Chem. 115:27984
Sipola M, Finckenberg P, Santisteban J, Korpela R, Vapataalo H, Nurminen M-L. 2001. Long-term intake
of milk peptides attenuates development of hypertension in spontaneously hypertensive rats. J. Physiol.
Pharmacol. 52:74554
Stevenson BR, Keon BH. 1998. The tight junction: morphology to molecules. Annu. Rev. Cell Dev. Biol.
14:89109
Suetsuna K, Maekawa K, Chen J-R. 2004. Antihypertensive effects of Undaria pinnatida (wakame) peptide
on blood pressure in spontaneously hypertensive rats. J. Nutr. Biochem. 15:26772
Takahashi S, Hori K, Shinbo M, Hiwatahi K, Gotoh T, Yamada S. 2008. Isolation of human renin inhibitor
from soybean: soyasaponin I is the novel human renin inhibitor in soybean. Biosci. Biotechnol. Biochem.
72:323236
Tanzadehpanah H, Asoodeh A, Saberi MR, Chamani J. 2013. Identication of a novel angiotensin-I converting
enzyme inhibitory peptide from ostrich egg white and studying its interaction with the enzyme. Innov.
Food Sci. Emerg. Technol. 18:21219
Tavares T, Sevilla M-A, Montero M-J, Carron R, Malcata FX. 2012. Acute effect of whey peptides upon
blood pressure of hypertensive rats, and relationship with their angiotensin-converting enzyme inhibitory
activity. Mol. Nutr. Food Res. 56:31624
Tenenbaum A, Grossman E, Shemesh J, Fisman EZ, Nosrati I, Motro M. 2000. Intermediate but not low
doses of aspirin can suppress angiotensin-converting enzyme inhibitor-induced cough. Am. J. Hypertens.
13:77682
Thayer AM. 2011. Improving peptides. Chem. Eng. News 89:1320
Tomatsu M, Shimakage A, Shinbo M, Yamada S, Takahashi S. 2013. Novel angiotensin I-converting enzyme
inhibitory peptides derived from soya milk. Food Chem. 136:61216
Tschudi MR, Criscione L, Novosel D, Pfeiffer K, Luscher TF. 1994. Antihypertensive therapy augments
endothelium-dependent relaxations in coronary arteries of spontaneously hypertensive rats. Circulation
89:221218

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

260

Aluko

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

FO06CH11-Aluko

ARI

14 March 2015

12:22

Udenigwe CC, Adebiyi AP, Doyen A, Li H, Bazinet L, Aluko RE. 2012a. Low molecular weight axseed
protein-derived arginine-containing peptides reduced blood pressure of spontaneously hypertensive rats
faster than amino acid form of arginine and native axseed protein. Food Chem. 132:46875
Udenigwe CC, Li H, Aluko RE. 2012b. Quantitative structureactivity relationship modeling of renininhibiting dipeptides. Amino Acids 42:137986
Udenigwe CC, Lin Y-S, Hou W-C, Aluko RE. 2009. Kinetics of the inhibition of renin and angiotensin
I-converting enzyme by axseed protein hydrolysate fractions. J. Funct. Foods 1:199207
van Elswijk DA, Diefenbach O, van der Berg S, Irth H, Tjaden UR, van der Greef J. 2003. Rapid detection and identication of angiotensin-converting enzyme inhibitors by on-line liquid chromatographybiochemical detection, coupled to electrospray mass spectrometry. J. Chromatogr. A 1020:4558
Vastag Z, Popovic L, Popovic S, Krimer V, Pericin D. 2011. Production of enzymatic hydrolysates with
antioxidant and angiotensin-I converting enzyme inhibitory activity from pumpkin oil cake protein isolate.
Food Chem. 124:131621
Vercruysse L, van Camp J, Morel N, Rouge P, Herregods G, Smagghe G. 2010. Ala-Val-Phe and Val-Phe:
ACE inhibitory peptides derived from insect protein with antihypertensive activity in spontaneously
hypertensive rats. Peptides 31:48288
Vermeirssen V, van Camp J, Verstraete W. 2004. Bioavailability of angiotensin converting enzyme inhibitory
properties. Br. J. Nutr. 92:35766
Wang C, Tian J, Wang Q. 2011. ACE inhibitory and antihypertensive properties of apricot almond meal
hydrolysate. Eur. Food Res. Technol. 232:54956
Wang GT, Chung CC, Holzman TF, Krafft GA. 1993. A continuous uorescence assay of renin activity.
Anal. Biochem. 210:35159
Wang J, Hu J, Cui J, Bai X, Du Y, et al. 2008. Purication and identication of a ACE inhibitory peptide from
oyster proteins hydrolysate and the antihypertensive effect of hydrolysate in spontaneously hypertensive
rats. Food Chem. 111:3028
Wang X, Wang L, Cheng X, Zhou J, Tang X, Mao X-Y. 2012. Hypertension-attenuating effect of whey
protein hydrolysate on spontaneously hypertensive rats. Food Chem. 134:12226
Wood JM, Schnell CR, Cumin F, Menard J, Webb RL. 2005. Aliskiren, a novel, orally effective renin inhibitor,
lowers blood pressure in marmosets and spontaneously hypertensive rats. J. Hypertens. 23:41726
Wu J, Aluko RE, Muir AD. 2002. Improved method for direct high-performance liquid chromatography assay
of angiotensin-converting enzyme-catalyzed reactions. J. Chromatogr. A 950:12530
Wu J, Aluko RE, Nakai S. 2006a. Structural requirements of angiotensin I-converting enzyme inhibitory
peptides: quantitative structure-activity relationship modeling of peptides containing 410 amino acids.
QSAR Comb. Sci. 25:87380
Wu J, Aluko RE, Nakai S. 2006b. Structural requirements of angiotensin I-converting enzyme inhibitory
peptides: quantitative structure-and-activity relationship study of di- and tri-peptides. J. Agric. Food Chem.
54:73238
Yamada A, Sakurai T, Ochi D, Mitsuyama E, Yamauchi K, Abe F. 2013. Novel angiotensin I-converting
enzyme inhibitory peptide from bovine casein. Food Chem. 141:378189
Yang CY, Dantzig AH, Pidgeon C. 1999. Intestinal peptide transport systems and oral drug availability. Pharm.
Res. 16:133143
Yang H-Y, Yang S-C, Chen J-R, Tzeng Y-H, Han B-C. 2004. Soyabean protein hydrolysate prevents the
development of hypertension in spontaneously hypertensive rats. Br. J. Nutr. 92:50712
Yang Y, Marczak ED, Yokoo M, Usui H, Yoshikawa M. 2003. Isolation and antihypertensive effect of angiotensin I-converting enzyme (ACE) inhibitory peptide from spinach Rubisco. J. Agric. Food Chem.
51:4897902
You S-J, Wu J. 2011. Angiotensin-I converting enzyme inhibitory and antioxidant activities of egg protein
hydrolysates produced with gastrointestinal and nongastrointestinal enzymes. J. Food Sci. 76:C8017
Yu Z, Yin Y, Zhao W, Chen F, Liu J. 2014. Antihypertensive effect of angiotensin-converting enzyme inhibitory peptide RVPSL on spontaneously hypertensive rats by regulating gene expression of the reninangiotensin system. J. Agric. Food Chem. 62:91217
Yuan W, Wang J, Zhou F. 2012. In vivo hypotensive and physiological effects of a silk broin hydrolysate on
spontaneously hypertensive rats. Biosci. Biotechnol. Biochem. 76:198789
www.annualreviews.org Antihypertensive Peptides

261

FO06CH11-Aluko

ARI

14 March 2015

12:22

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Zhang C, Cao W, Hong P, Ji H, Qin X, He J. 2009. Angiotensin I-converting enzyme inhibitory activity of
Acetes chinensis peptic hydrolysate and its antihypertensive effect in spontaneously hypertensive rats. Int.
J. Food Sci. Technol. 44:204248
Zhang J-H, Tatsumi E, Ding C-H, Li L-T. 2006. Angiotensin I-converting enzyme inhibitory peptides in
douche, a Chinese traditional fermented soybean product. Food Chem. 98:55157
Zhang Y, Olsen K, Grossi A, Otte J. 2013. Effect of pretreatment on enzymatic hydrolysis of bovine collagen
and formation of ACE-inhibitory peptides. Food Chem. 141:234354
Zheng T, Ichiyu S, Li NW, Mitsumine F, Yasuhiko T. 1997. Effects of antihypertensive drugs or glycemic
control on antioxidant enzyme activities in spontaneously hypertensive rats with diabetes. Nephron 76:323
30
Zhou F, Xue Z, Wang J. 2010. Antihypertensive effects of silk broin hydrolysate by alcalase and purication
of an ACE inhibitory dipeptide. J. Agric. Food Chem. 58:673540

262

Aluko

FO06-FrontMatter

ARI

25 March 2015

13:23

Annual Review
of Food Science
and Technology

Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org


Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

Contents

Volume 6, 2015

An Amazing Journey
Larry McKay p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Physical Modication of Food Starch Functionalities
James N. BeMiller and Kerry C. Huber p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p19
Nanostructured Fat Crystal Systems
Nuria C. Acevedo and Alejandro G. Marangoni p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p71
Antimicrobial Food Equipment Coatings: Applications and Challenges
Luis J. Bastarrachea, Anna Denis-Rohr, and Julie M. Goddard p p p p p p p p p p p p p p p p p p p p p p p p p p p97
Non-Nutritive Sweeteners and Obesity
John D. Fernstrom p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 119
Genetic Mechanisms of Prebiotic Oligosaccharide Metabolism in
Probiotic Microbes
Yong Jun Goh and Todd R. Klaenhammer p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 137
Electrostatic Coating Technologies for Food Processing
Sheryl A. Barringer and Nutsuda Sumonsiri p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 157
The Role of Oxygen in Lipid Oxidation Reactions: A Review
David R. Johnson and Eric A. Decker p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 171
Stress Adaptation in Foodborne Pathogens
Maire Begley and Colin Hill p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 191
Colloids in Food: Ingredients, Structure, and Stability
Eric Dickinson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 211
Antihypertensive Peptides from Food Proteins
Rotimi E. Aluko p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 235
Pickering Emulsions for Food Applications: Background, Trends, and Challenges
Claire C. Berton-Carabin and Karin Schroen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 263
The Nutraceutical Bioavailability Classication Scheme: Classifying Nutraceuticals
According to Factors Limiting their Oral Bioavailability
David Julian McClements, Fang Li, and Hang Xiao p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 299
v

FO06-FrontMatter

ARI

25 March 2015

13:23

Comparative Analysis of Intestinal Tract Models


C.F. Williams, G.E. Walton, L. Jiang, S. Plummer, I. Garaiova,
and G.R. Gibson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 329
Bacillus and Other Spore-Forming Genera: Variations in Responses
and Mechanisms for Survival
Aleksandra Checinska, Andrzej Paszczynski, and Malcolm Burbank p p p p p p p p p p p p p p p p p p p p 351
Protein-Polysaccharide Interactions to Alter Texture
Fred van de Velde, Els H.A. de Hoog, Alexander Oosterveld, and R. Hans Tromp p p p p p 371
Annu. Rev. Food Sci. Technol. 2015.6:235-262. Downloaded from www.annualreviews.org
Access provided by New York University - Bobst Library on 04/17/15. For personal use only.

High Hydrostatic Pressure Processing: A Promising Nonthermal


Technology to Inactivate Viruses in High-Risk Foods
Fangfei Lou, Hudaa Neetoo, Haiqiang Chen, and Jianrong Li p p p p p p p p p p p p p p p p p p p p p p p p p p 389
Human Norovirus as a Foodborne Pathogen: Challenges
and Developments
Matthew D. Moore, Rebecca M. Goulter, and Lee-Ann Jaykus p p p p p p p p p p p p p p p p p p p p p p p p p p 411
Principles and Application of High PressureBased Technologies in
the Food Industry
V.M. (Bala) Balasubramaniam, Sergio I. Martnez-Monteagudo,
and Rockendra Gupta p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 435
Challenges and Solutions to Incorporation of Nutraceuticals in Foods
Mary Ann Augustin and Luz Sanguansri p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 463
Statistical Aspects of Food Safety Sampling
I. Jongenburger, H.M.W. den Besten, and M.H. Zwietering p p p p p p p p p p p p p p p p p p p p p p p p p p p p 479
Diet-Based Strategies for Cancer Chemoprevention: The Role of
Combination Regimens Using Dietary Bioactive Components
Christina DiMarco-Crook and Hang Xiao p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 505
Collagen and Gelatin
Dasong Liu, Mehdi Nikoo, Gokhan Boran, Peng Zhou, and Joe M. Regenstein p p p p p p p p 527
Indexes
Cumulative Index of Contributing Authors, Volumes 16 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 559
Cumulative Index of Article Titles, Volumes 16 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 562
Errata
An online log of corrections to Annual Review of Food Science and Technology articles may
be found at http://www.annualreviews.org/errata/food

vi

Contents

Das könnte Ihnen auch gefallen