Maisuria 2012

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Biochemical Engineering Journal 63 (2012) 22–30

Contents lists available at SciVerse ScienceDirect

Biochemical Engineering Journal


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

Biochemical properties and thermal behavior of pectate lyase produced by


Pectobacterium carotovorum subsp. carotovorum BR1 with industrial potentials
V.B. Maisuria, A.S. Nerurkar ∗
Department of Microbiology and Biotechnology Centre, Faculty of Science, M. S. University of Baroda, Vadodara 390002, Gujarat, India

a r t i c l e i n f o a b s t r a c t

Article history: An extracellular pectate lyase (PL, EC 4.2.2.2) produced by Pectobacterium carotovorum subsp. carotovorum
Received 31 July 2011 BR1 was purified and characterized with respect to its biochemical properties including the thermo-
Received in revised form 15 January 2012 dynamic parameters of substrates hydrolysis and thermal behavior. The PL showed higher affinity for
Accepted 25 January 2012
polygalacturonic acid (Km , 0.4 g l−1 ) as compared to 70% methylated pectin (Km , 0.77 g l−1 ). Nonlinear
Available online 1 February 2012
thermal inactivation curves of PL at 50 and 60 ◦ C, fitted to a three-fraction first-order kinetic model. Fur-
thermore, the thermodynamic parameters of thermal deactivation of purified PL viz. H*, S*, Ed and
Keywords:
G* were determined over pH range 6–10 and in presence of different cations. Alkaline pH and Ca2+ ions
Pectate lyase
Enzyme biocatalysis
individually influenced the thermostability of PL at elevated temperatures 60 and 70 ◦ C, based on the
Kinetic parameters thermodynamic parameters. The studies suggested that this alkaline PL can be considered as potential
Enzyme deactivation candidate for various applications like pectic waste management, degumming of fibers, food, paper and
Thermodynamic parameters textile industries.
Enzyme technology © 2012 Elsevier B.V. All rights reserved.

1. Introduction carotovora subsp. carotovora (Ecc) reclassified or renamed as Pecto-


bacterium carotovorum subsp. carotovorum (Pcc). Pcc also produces
Microbial pectolytic enzymes specifically degrade the pectin, other pectolytic enzymes viz. polygalacturonase (PG), pectin lyase
which constitutes the middle lamella of plant cell-wall. Pectolytic (PNL), and pectin methyl esterase (PME) for penetration into the
enzymes differ in their substrate preference (polygalacturonic acid plant cell wall leading to maceration of plant tissues [2].
and low-methylated pectin or high-methylated pectin), in their Microbial alkaline pectinases have wide-range of biotech-
reaction mechanism (␣/␤-elimination or hydrolysis), and mode nological applications in food processing, treatments of pectic
of attack (terminal or in random) on the pectin polymer (exo wastewaters, textile processing, paper making, degumming of
or endo) [1]. Among pectolytic enzymes, pectate lyase (PL, EC plant bast fibers, coffee and tea fermentations [3] also for plant
4.2.2.2) catalyzes cleavage by ␤-eleminative mechanism of the protoplast isolation and enzymatic solubilization of potato pulp
␣-1,4-glycosidic bond of polygalacturonic acid, which produces in dietary fiber extraction processes and agro-industrial wastes
4,5-unsaturated galacturonosyl at the non-reducing end [2]. Cal- management [4–6]. There are several reports of fungal alkaline
cium was found to be essential for the activity of pectate lyases from PLs being extensively used in degumming of ramie fibers, jute
both bacterial and fungal origin. It plays an important role in the retting, textile and paper industries [3], but scarcity of informa-
decomposition of plant residues, during infection caused by Erwinia tion available on bacterial alkaline PL for industrial applications.
Biotechnological applications of these enzymes are highly depen-
dent on parameters such as pH, salt concentration and thermal
stability at high temperature, since their application in industrial
Abbreviations: PL, pectate lyase; PG, polygalacturonase; PNL, pectin lyase; Pcc, processes requires reactions to be conducted at high temperatures
Pectobacterium carotovorum subsp. carotovorum; Ecc, Erwinia carotovora subsp. caro- in order to improve productivity [7]. Microbial sources of alkaline
tovora; Eca, E. carotovora subsp. atroseptica; Ech, E. chrysanthemi; w/v, weight per
volume; H# , enthalpy of substrate hydrolysis; S# , entropy of substrate hydrol-
pectinases is restricted to very few bacterial genera i.e. Bacillus
ysis; G# , free energy changes during substrate hydrolysis; GE#–S , free energy sp., Pseudomonas sp., Xanthomonas sp., and also fungi and acti-
changes of enzyme–substrate complex; GE#–T , free energy changes of transition nomycetes [3], among them a small number of microbial strains
state formation; Ea , activation energy; Q10 , temperature quotient; Kd , deactiva- such as Debaryomyces nepalensis [8], Aspergillus niger [9], and P.
tion rate constant; t1/2 , half-life time; H*, enthalpy of thermal inactivation; S*,
carotovorum subsp. carotovorum (Pcc) [2], are known to produce
entropy of thermal inactivation; G*, free energy change during thermal inactiva-
tion; Ed , thermal inactivation energy. both PL and PNL in appreciable amounts. A Pcc strain was found
∗ Corresponding author. Tel.: +91 265 2794396; fax: +91 265 2792508. to produce PL that demonstrated properties which were found
E-mail address: anuner26@yahoo.com (A.S. Nerurkar). worth exploring for their biotechnological potential. Studies like

1369-703X/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.bej.2012.01.007
V.B. Maisuria, A.S. Nerurkar / Biochemical Engineering Journal 63 (2012) 22–30 23

purification and characterization of PL and heterologous expression various temperatures ranging from 20 to 70 ◦ C for 1 h, the sub-
of pel gene from Pcc strain has been reported [10–12]. However, strate being pre-incubated for 5 min at each temperature. The Ea
there is sparse information regarding enzyme kinetic properties, was determined using Arrhenius plot and temperature quotient
thermostability, inactivation kinetics and thermodynamic param- (Q10 ) was estimated as reported earlier [13]. The effect of the pH
eters studies of the PL produced by Pcc. This, to our knowledge is on the PL was determined by measuring the activity at 50 ◦ C for 1 h
first report of extensive characterization of purified PL from Pcc using different buffers (50 mM) containing 1 mM CaCl2 , viz. phtha-
BR1, with respect to biochemical properties, thermal inactivation late buffer (pH 2.6–3.0), Na-acetate buffer (pH 4.0–5.4), phosphate
kinetics, pH and cation dependent thermal behavior. Based on these buffer (pH 6.0–7.5), Tris–HCl buffer (pH 7.5–8.5) and glycine–NaOH
studies it is proposed that PL from Pcc BR1 possesses biotechnolog- buffer (pH 9.0–11.0).
ical potential.
2.5. Estimation of enzyme kinetics and thermodynamic
2. Materials and methods
parameters

2.1. Microorganism and culture conditions


The kinetic parameters, Km , Vmax and Kcat were determined by
measuring the PL activity reaction rates (as in Section 2.4 assay
The pectolytic bacterium used in this study was isolated from
conditions) at different substrate concentrations ranging from 0.01
macerated tissue of Brinjal fruit (Solanum melongena var. esculen-
to 0.5% (w/v) with a fixed amount of the enzyme in a 50 mM
tum), and was identified and renamed as P. carotovorum subsp.
Tris–HCl buffer (pH 8.5) with 1 mM CaCl2 at 50 ◦ C for 1 h. The Km
carotovorum (Pcc) BR1 (NCBI-GenBank accession no. FJ187821,
and Vmax values were obtained by analysis of data according to
earlier reported as E. carotovora subsp. carotovora BR1). The prop-
Lineweaver–Burk equation, allowing the catalytic efficiency i.e., the
agation was done on nutrient agar medium (Himedia Lab. Pvt. Ltd.,
ratio Vmax /Km , and kinetic constants (Kcat /Km ) to be determined
India) containing 0.5% (w/v) pectin and incubated overnight at 30 ◦ C
[13]. The substrate specificity in terms of the specific activity was
[13].
established using a 0.5% (w/v) solution of different pectic substrates
viz. polygalacturonic acid and pectin with a 28%, 35%, 65% and
2.2. Production and purification of PL
70% degree of esterification (DE). The thermodynamic parameters
(H# , S# , G# , GE#–S and GE#–T ) for the substrate hydrolysis
Pectolytic enzyme production was carried out in 500 ml of liq-
were calculated by rearranging Eyring’s absolute rate equation as
uid medium containing 0.5% (w/v) pectin, and at the end of the
described earlier [13].
growth phase culture supernatant was precipitated with chilled
acetone [13]. The purification of PL was followed with some mod-
ifications as per Basu et al. [7] and Kamiyama et al. [14]. The 2.6. Thermal stability assay
precipitates were centrifuged at 15,880 × g for 15 min, after stor-
ing at 4 ◦ C for 4 h, dissolved in a 20 mM Tris–HCl buffer (pH 8.0), The thermal deactivation was determined by incubating the
dialyzed against the same buffer, and then chromatographed on a enzyme at 20–70 ◦ C temperature range. Aliquots were withdrawn
CM-cellulose column (0.8 cm × 12 cm) at a flow rate of 0.5 ml min−1 at 10 min intervals up to 90 min, cooled on ice for 30 min, and PL
using a linear gradient of 0–1 M NaCl in 20 mM Tris–HCl buffer (pH activity was assayed. The residual activity was estimated as [18]:
8.0). Protein molecular weight markers (PMWH; Banglore GenieTM , E 
India): carbonic anhydrase, 29 kDa; ovalbumin, 43 kDa; bovine t
Residual PL activity (%) = 100 (1)
serum albumin, 66 kDa; myosin-rabbit muscle, 205 kDa were run E0
along the samples in the SDS–PAGE using a 12% polyacrylamide
where Et is the activity at time t (min), and E0 is the initial activity
gel [15], and the gel was stained using silver salts after the elec-
at time t = 0 min.
trophoresis [16]. The activity staining assay of the PL was modified
and performed using 12% polyacrylamide gel containing polygalac-
turonic acid as the substrate [17]. After gel-electrophoresis, the 2.7. Models for PL thermal deactivation kinetics
gel was incubated overnight at 50 ◦ C in a 50 mM Tris–HCl buffer
(pH 8.5) containing 1 mM CaCl2 and then stained with 0.01% (w/v) 2.7.1. First-order model
Toluidine blue-O instead of Ruthenium red. As described by Ortega et al. [18] thermal deactivation of
enzymes can often be explained by a first-order kinetic model.
2.3. Enzyme activity assay of PL According to this model an enzyme activity decreases log-linearly
as a function of time as described in the following equation:
The PL activity was determined by measuring oligosaccharides E 
released as a result of the cleavage of polygalacturonic acid using a t
ln = −Kd t (2)
thiobarbituric acid (TBA) reagent [7]. A reaction mixture containing E0
500 ␮l of 0.5% (w/v) polygalacturonic acid (PGA) in 50 mM Tris–HCl
where Et is the enzyme activity at time t (min), E0 is initial enzyme
buffer (pH 8.5) containing 1 mM CaCl2 and 100 ␮l of the appropri-
activity, t is time of treatments, and Kd is the first-order deactivation
ately diluted enzyme solution was incubated at 50 ◦ C for 1 h and the
rate constant.
final absorbance of the reaction products was measured at 550 nm.
One unit of PL activity was defined as the amount of enzyme that
caused a change in the absorbance of 0.01 per hour at 50 ◦ C and pH 2.7.2. Multi-fraction deactivation model
8.5. Multi-fraction first-order deactivation kinetic model according
to Ortega et al. [18] was used to analyze kinetic data. Following the
2.4. Effect of temperature and pH on PL activity first-order kinetics, the implicit assumptions were that n fractions
of PL existed and that each fraction was deactivated independently,
The temperature optima and activation energy (Ea ) were deter- i.e.,
mined by incubating an appropriate amount of the enzyme with
0.5% polygalacturonic acid in 50 mM Tris–HCl buffer (pH 8.5) at N1 → I1 , N1 → I1 , . . . , Ni → Ii (3)
24 V.B. Maisuria, A.S. Nerurkar / Biochemical Engineering Journal 63 (2012) 22–30

where N is the active enzyme and I is the inactive enzyme. The where Kb is the Boltzmann’s constant (R/N) = 1.38 × 10−23 J/K,
first-order inactivation rate process is: R the gas constant = 8.314 J/K/mol and h the Planck’s con-
stant = 6.626 × 10−34 J s.
dEi
= Ki Ei (4) Free energy change were calculated by using the following rela-
dt
tionship
After integration
G∗ = H ∗ − T S ∗ (12)
Ei = E0i e−Ki t (5)
where T is the absolute temperature (K).
Residual activity at time t is the sum of the activities of individual Energy of deactivation was estimated using the Arrhenius equa-
fractions, i.e., tion:

n
Kd = Ae(−Ed /RT ) (13)
E= E0i e−Ki t (6)
i=1
So that,
Ed
where Ki and E0i are the deactivation rate constant and enzyme ln[Kd ] = − + ln A (14)
activity fraction of the ith fraction. Eq. (6) was fitted to the thermal RT
deactivation kinetics data for i = 1, 2 and 3, corresponding to first Energy (Ed ) involved in this deactivation process was calculated
order, two-fraction first order and three-fraction first order models, from the slope of a linear plot of 1/T versus ln[Kd ], Ed = −slope × R,
using a non-linear regression routine in a SigmaStat3.5 and vali- where R (gas constant) = 8.314 J/K/mol. Thermalstability of enzyme
dated with Origin 6.0 software package to obtain deactivation rate in the presence of different pH and cations was determined by heat-
constants and activity fractions. ing the enzyme solution at 40–70 ◦ C in the presence of the salt or in
specified pH in sealed tubes. The results represented are the mean
2.8. Effect of pH or cations on thermal stability values of the experiments conducted in triplicate. Data were statis-
tically analyzed using SigmaStat 3.5, GraphPad Prism 4 and Origin
A pH range from 6.0 to 10.0 was selected for determining the 6.0 software packages.
thermal stability of purified PL. The pH of the enzyme solutions
was adjusted using 1 M HCl or 1 M NaOH, followed by a thermal 3. Results and discussion
stability assay according to Gohel and Naseby [19]. The different
salts NaCl (0.1 M), KCl (0.15 M), MgCl2 (5 mM), CaCl2 (1 mM), and 3.1. Purification and characterization of pectate lyase from Pcc
MnCl2 (1 mM) were selected on the basis of observations of activity BR1
enhancement reported by other authors [14,20]. The purified and
dialyzed enzyme solutions was initially incubated with each salt The pectate lyase purified from culture supernatant of Pcc BR1
separately at 4 ◦ C for 10 min, followed by a thermal stability assay using ion-exchange column chromatography was more than 95%
at individual temperature. pure as judged on SDS–PAGE with 34% yield giving 23 fold purifi-
cation. The size of the purified PL was found to be around 40 kDa
2.8.1. Estimation of deactivation rate constant and as validated by activity staining zymogram for PL (Fig. 1) which
thermodynamic parameters for thermal deactivation was similar to those reported for E. carotovora, PLI (44 kDa), PLII
Purified PL samples were subjected to temperature between 40 (41 kDa) by Lei et al. [11]. PL showed enhanced and optimum activ-
and 70 ◦ C for up to 120 min to observe the thermal stability in pres- ity at 50 ◦ C and pH 8.5 (data not shown). Sugiura et al. [12] have
ence of different salts and pH conditions. The residual PL activity also reported similar temperature optima for PL of E. carotovora Er.
was estimated before and after the incubation period at each tem- The activation energy (Ea ) as determined by Arrhenius model for
perature under the respective conditions. The deactivation rate of PL was 2.02 kJ mol−1 , lower than that reported for PL of B. pumilus
PL was calculated by first order expression [9]: DKS1 (13.37 kJ mol−1 ) [7]. The temperature quotient (Q10 ) for PL of
Pcc BR1 was obtained 1.0 at 20–70 ◦ C, similar to that of PG (1.03) of
dE same strain reported earlier [13]. It has broad substrate specificity
= −Kd E (7)
dt as demonstrated by the relative activity (RA) with polygalacturonic
So that, acid (100% RA), 28% DE pectin (92% RA), 35% DE pectin (82% RA),
E  and 70% DE pectin (80% RA). This unusual substrate preference of
t
ln = −Kd t (8) PL toward moderate and high methyl esterified pectin has been
E0
observed earlier for PelB of E. carotovora subsp. carotovora (Ecc)
The Kd (deactivation rate constant or first order rate constant) [10]; and PL3 of E. carotovora subsp. atroseptica (Eca) [21].
values were calculated from a plot of time (t) versus ln[Et /E0 ] at a
particular temperature. The half-life of an enzyme is defined as the 3.2. Kinetics and thermodynamic parameters of PL
time required by the enzyme to lose half of its initial activity, which
is calculated as: 3.2.1. Michaelis–Menten kinetic parameters
ln 2 PL activity showed a typical Michaelis–Menten profile with two
t1/2 = (9) different substrates viz. polygalacturonic acid and 70% DE pectin,
Kd
in both cases the curves were linear with correlation coefficient
Thermodynamic parameters were calculated by rearranging the (R2 ) of 0.9945 and 0.9845 respectively. The values of Km and Vmax
Eyring absolute rate equation [13]. H* and S* values were esti- were 0.4 g l−1 and 67.57 U ml−1 respectively for polygalacturonic
mated from the slope and intercept of a 1/T versus ln[Kd /T] plot, acid and 0.77 g l−1 and 57.35 U ml−1 respectively for 70% DE pectin.
respectively. So that, The values of Km indicate that the PL had a relatively low affinity
for 70% DE pectin than for the polygalacturonic acid. These appar-
H ∗ = −(slope)R (10)
  K  ent values of Km are higher than PLs from Eca (0.27–0.30 g l−1 )
S ∗ = R intercept − ln b
(11) [21] and E. chrysanthemi (Ech) (0.20–0.32 g l−1 ) [22], but within
h the range of that reported for other bacterial PLs from B. pumilus
V.B. Maisuria, A.S. Nerurkar / Biochemical Engineering Journal 63 (2012) 22–30 25

Fig. 1. (A) SDS–PAGE analysis and (B) activity staining zymogram of purified PL from Pcc BR1 at different purification steps. Lane M: protein MW markers, 1: purified PL, and
2: dialyzed acetone precipitates.

DKS1 (0.44 g l−1 ) [7] and B. pumilus BK2 (0.24–1.35 g l−1 ) [23]. The spontaneously forms the product  4:5 unsaturated oligogalac-
turn over number (Kcat ) and second order rate constant (Kcat /Km ) turonides.
were 8.7 min−1 and 21.75 respectively with polygalacturonic acid,
while 7 min−1 and 9.1 respectively with 70% DE pectin. Moreover,
the catalytic efficiency (Vmax /Km ) of PL was found to be higher for 3.3. Thermalstability kinetics
polygalacturonic acid (169.0) as compared to 70% DE pectin (70.6).
The catalytic efficiency value provides a practical model for select- Thermostable enzymes have substantial potential for many
ing the most efficient enzyme for a commercial application process industrial applications. Some specific applications as in food, paper
using a fixed initial substrate concentration [18]. and textile industrial processes are carried out at high temper-
ature in order to reduce microbial contamination, maximize the
reaction rate and improve the productivity. Therefore there is
emphasis on the elucidation of thermal deactivation mechanisms
3.2.2. Thermodynamic parameters for substrate hydrolysis and the development of strategies for enhancing the stability of
Thermodynamic parameters viz. enthalpy of activation (H# ), thermostable enzyme [18]. Thermalstability is the property of an
entropy of activation (S# ) and Gibbs free energy of activation enzyme to resist thermal unfolding in the absence of a substrate,
(G# ) measured during enzymatic substrate hydrolysis process while thermophilicity is the ability of an enzyme to catalyze the
provide a relative idea about the functionality of enzyme toward reaction at elevated temperatures in the presence of a substrate
specific substrate [13]. The H# , S# and G# of PL for polygalac- [24].
turonic acid hydrolysis were 0.098 kJ mol−1 , −193.73 J mol−1 K and The inactivation curves of PL of Pcc BR1 depicted in semi-
58.8 kJ mol−1 respectively, while for 70% DE pectin hydrolysis these logarithmic graphs were not linear in the temperature range
values were 0.098 kJ mol−1 , −191.97 J mol−1 K and 58.267 kJ mol−1 50–70 ◦ C (Fig. 2). Evidently, they have multiphase nature there-
respectively. Similar values were observed previously for PG from fore it is inferred that the thermal inactivation of PL was not
Ecc BR1 [13]. The low enthalpy value suggests that the PL is simple first-order process commonly observed for enzyme inac-
very efficient for formation of transition state complex between tivation but a multi-fraction process. Multi-fraction inactivation
enzyme–substrate (E–S). The feasibility and extent of the chem- kinetics has been also reported for other pectic enzymes [18]. Nath
ical reaction is evaluated by estimating the change in the Gibbs et al. [25] have marked that biphasic thermal inactivation is shown
free energy (G# ) for conversion of substrate to product [13]. G# by the enzyme when it consists of two groups of thermostable
value was lower with both substrates indicating that the conversion and thermolabile subunits or isoenzymes or due to formation of
of the transition state (E–S) into product is spontaneous, whereas enzyme aggregates during thermal-inactivation, as observed by
the low S# value with both substrates suggested that the transi- Ortega et al. [18] for Rapidase C80 (commercial preparation of
tion state complex of PL had less disorder. PG).
The free energy changes for the activation of the substrate Non-linear, multi phase behavior was shown by PL of Pcc BR1,
binding (GE#–S ) and free energy changes for the formation of evident from the inactivation curves at 50 and 60 ◦ C where they
the activation complex (GE#–T ) of the PL were determined as exhibit different slopes at short and long incubation times (Fig. 2).
−2.31 kJ mol−1 and −7.76 kJ mol−1 respectively for polygalactur- The multi-fraction first-order model was fitted to inactivation
onic acid; −0.66 kJ mol−1 and −5.563 kJ mol−1 respectively for 70% results obtained for PL from non linear curves at 50 and 60 ◦ C
DE pectin. This reconfirmed that the PL had a high affinity toward (Table 1). The high coefficient of determination (0.993 at 50 ◦ C and
polygalacturonic acid than 70% DE pectin for hydrolysis which 0.988 at 60 ◦ C) for the three-fraction first order model and highly
26 V.B. Maisuria, A.S. Nerurkar / Biochemical Engineering Journal 63 (2012) 22–30

o o o o o o
20 C 30 C 40 C 50 C 60 C 70 C

7.36 × 10−6 (2.98 × 10−6 )g


1.56 × 10−3 (4.09 × 10−4 )g
1.24 × 10−1 (1.53 × 10−2 )f
4.5

Three-fraction
4.0

2.74 × 10−2 f

ln [residual PL activity, %]
3.5
26.48

1.268
20.73

0.988
22.00
3.0

2.5

2.0
8.897 × 10−2 (8.86 × 10−3 )f
5.71 × 10−4 (9.48 × 10−5 )f

1.5

1.0
Two-fraction

4.74 × 10−2 f

0.5
48.48

1.267
20.73

0.977

0.0
0 20 40 60 80

Time (min)

Fig. 2. Pseudo-first order plot for irreversible thermal inactivation of purified PL


from Pcc BR1. Data presented are average values of n = 3 independent experiments
3.84 × 10−2 (5.58 × 10−3 )f

and error bars represent the standard deviation.


Single-fraction

significant mean square errors at 50 ◦ C (4.46 × 10−3 at P < 0.001)


0.257f

and 60 ◦ C (2.74 × 10−2 at P < 0.001) revealed that the inactivation


69.21

4.146
0.855
60 ◦ C

curve was better explained by the three-fraction first-order model.




In both the cases of inactivation curves at 50 and 60 ◦ C, the stan-


dard errors of estimated values of dissociation rate constants (Kd1
and Kd2 ) were higher than that for the two-fraction first-order
7.38 × 10−4 (1.65 × 10−4 )g
2.94 × 10−3 (6.17 × 10−3 )

5.5 × 10−6 (1.2 × 10−6 )g

model. In contrast, as shown in Table 1, in case of three-fraction


first-order model, the standard errors for estimated constants Kd1 ,
Kd2 , and Kd3 were considerably lower than the parameter values
Three-fraction

of inactivation curves at both temperatures 50 and 60 ◦ C. There-


4.46 × 10−3 f
Determination of thermal inactivation using multi fraction first order kinetic models for PL of Pcc BR1.

fore, the three-fraction first-order model was found most suitable


0.0570
66.69
21.61

0.993

than the two-fraction first-order model for explaining the inactiva-


11.7

tion kinetic data of PL. Alternatively, the three-fraction first-order


model was not adequate to explain the inactivation curve of PL
at 70 ◦ C, because PL shows very low i.e. only 1.3% residual activ-
ity after 30 min of treatment at 70 ◦ C (Fig. 2). Similar observations
2.24 × 10−2 (1.35 × 10−3 )g
4.52 × 10−6 (5.7 × 10−5 )

have been reported for PL of B. pumilus DKS1 at temperatures


beyond 70 ◦ C [7] and commercial preparation of PG (Rapidase
C80) at 60 ◦ C [18]. As observed by Ortega et al. [18], at elevated
Two-fraction

temperature, the labile fraction was inactivated rapidly therefore


1.72 × 10−2 f

only rate constants of thermostable fraction could be determined.


Unit for Ei is % PL activity and for Kdi is s−1 where i = 1; 2; 3.
0.971
0.384

Three phases of PL inactivation at higher rate were designated, as


88.3
11.7

phase-I, phase-II and phase-III (Fig. 3). The deactivation energy (Ed )

of PL calculated from the Arrhenius plot of thermal inactivation


(Fig. 3) at three phases-I (R2 , 0.97), II (R2 , 0.98) and III (R2 , 0.99)
were 95.94 kJ mol−1 , 97.05 kJ mol−1 and 143.81 kJ mol−1 respec-
2.2 × 10−2 (1.35 × 10−3 )f

Values in parentheses are standard errors.

tively. These values clearly emphasize robustness of inactivation


Predicted residual error sum of square.

model for PL. Furthermore, the effect of temperature on half-life


Statistically significant at P < 0.001
Single-fraction

Statistically significant at P < 0.05.

time of PL has been studied which is summarized in Table 2. The


Temperatures

1.5 × 10−2 f

half-life time (t1/2 ) value is generally assumed as additive [26] and


Coefficient of determination.

therefore values of phase-III are combined t1/2 values of other two


0.971
0.206
50 ◦ C

phases. The t1/2 value of PL in phase-III at 50 ◦ C was 119 folds higher,


97.9


whereas at 60 ◦ C the value was 25 fold higher than combined val-


Mean square error.

ues of phase-I and -II. This suggests that the enzyme was more
heat-stable in phase-III (Table 2). The denaturing reactions of con-
Parametersa

formationally intact proteins at high temperatures occur slowly,


suggesting that the conformational stability of the protein struc-
PRESSd
Table 1

MSEb
Kd1 e
Kd2 e
Kd3 e

ture may be important to explain the upper temperature limit for


R2 c
E1
E2
E3

enzyme activity [27].


V.B. Maisuria, A.S. Nerurkar / Biochemical Engineering Journal 63 (2012) 22–30 27

Table 2
Effect of temperature on half-life time with three-fraction first-order inactivation kinetic model of PL.

Temperature (◦ C) Half life (min)

Phase-I Phase-II Phase-III

20 1.73 × 103 6.93 × 104 2.31 × 107


30 2.24 × 102 1.73 × 104 3.47 × 106
40 1.61 × 102 1.16 × 104 1.39 × 106
50 1.10 × 102 8.66 × 102 1.16 × 105
60 10.44 6.93 × 102 1.73 × 104
70 3.68 2.39 × 102 –

3.4. Thermal inactivation studies of PL 8.0 > pH 7.0 > pH 9.0 ≥ pH 10.0 > pH 6.0 (Fig. 4A and C). The inacti-
vation of enzyme involves the unfolding of protein structure due
The thermal inactivation studies are crucial to understand- to changes in the balance of electrostatic and hydrogen bonds. This
ing the relationship between structure and function of enzyme. is governed by pH induced changes of ionization state of ionogenic
Enzymes inactivation is a major constraint in the biotechnologi- groups of proteins [28]. It is envisaged that the thermal unfolding
cal process development. Understanding the complex process of of the PL in relaxed state at 60 ◦ C and in alkaline pH conditions
enzyme inactivation is helpful in enhancing the feasibility of many due to changes in the balance of the electrostatic and hydrogen
biological processes [28]. There is paucity of information avail- bonds, possibly influence its thermalstability. Also the half-life time
able concerning transition state parameters of heat inactivation (t1/2 ) of the PL with different pH and temperature combinations,
of PL from soft rot Erwinia. The effect of various combinations increased significantly (P < 0.001) toward the alkaline pH (up to pH
of pH and temperature on inactivation of PL as well as the 8.0) at 40–70 ◦ C. As shown in Fig. 4C, at a elevated temperature
effect of different salts on thermal inactivation of PL was evalu- of 70 ◦ C, the t1/2 of the purified PL was significantly higher at pH
ated as described in Section 2.8.1. The extent of deactivation is 8.0 (7.5 min) and lower at pH 6.0 (5.4 min), suggesting that alka-
assessed by the deactivation rate, which is proportional to the line pH condition influences the thermal stability of PL to certain
active enzyme concentration and deactivation rate constant (Kd ) extent.
[9]. Fig. 3 shows linear inactivation curves in Arrhenius plots for
heat induced inactivation of PL at 20–70 ◦ C temperature range. The
first-order inactivation kinetic model fitted these curves based on
which the deactivation rate constant was calculated for each con- 3.4.2. Effect of cations on the thermal inactivation of PL
dition. Similarly, the deactivation rate constant (Kd ) and half-life time
(t1/2 ) of the PL were calculated at a temperature range of 40–70 ◦ C
in the presence of different cations. The presence of salts ions plays
3.4.1. Thermal inactivation of PL with different pHs
prominent role on enzyme activity and stability [28]. Influence of
The deactivation rate constants (Kd ) and half-life time (t1/2 )
different divalent cations (Ba2+ , Co2+ , Cu2+ , Mg2+ , Mn2+ , Sr2+ and
were estimated at 40–70 ◦ C temperature range with different pHs
Zn2+ ) and monovalent cations (Na+ and K+ ) for optimum PL activity
– 6.0, 7.0, 8.0, 9.0, and 10.0 (Fig. 4A and C). pH is one of the most
have been extensively investigated [1,7,14,20]. PLs are known to
important factors affecting the tertiary and quaternary structure
have an absolute requirement for Ca2+ ions for their production and
of proteins and therfore enzymes. Since the PL exhibits optimum
activity. Also the Mn2+ ion similar to Ca2+ was reported as a strong
activity at pH 8.5, the pH range 6–10 was selected for these studies.
cofactor, that could act as inducer of PL activity [7,29]. This led to
The thermalstability of the PL was significantly higher at pH 8.0 at
the choice of five different cations (Na+ , K+ , Ca2+ , Mg2+ and Mn2+ )
temperature 50 and 60 ◦ C (Fig. 4A and C). The order of stabiliza-
in chloride form for the study of their effect on thermal inactivation
tion of PL in the various pH conditions at 70 ◦ C was as follows: pH
of PL in the 40–70 ◦ C temperature range.
The PL exhibited reduced deactivation rate and increased half-
life time in presence of cations at each temperatures studied in
comparison to the control PL (without any cations) (Fig. 4B and D).
0 Overall, the cations used in this study were found to increase the
stability of the PL, especially at temperature 40 and 50 ◦ C. The effec-
-3 tiveness of the various cations during heat induced inactivation
was in the following order: Ca2+ > Mn2+ > Mg2+ > Na+ > K+ , indicat-
ing that the effect of divalent cations on the thermostability of
Ln Kd (min )

-6
-1

Phase I the enzyme was higher as compared to monovalent cations. The


stabilization of the PL by salts may have been due to a reduction
-9
in unfavorable electrostatic repulsion leading to reduced unfavor-
able electrostatic free energy. The stability of the PL was found to
-12
Phase II be significantly higher with Ca2+ , with the lowest Kd values being
obtained at and between 40 and 70 ◦ C (Fig. 4B). The t1/2 values of
-15 the enzyme with the different cations were found to be signifi-
Phase III cantly increased particularly with Ca2+ at a temperature range of
-18 40–70 ◦ C (Fig. 4D). These half-life time values of PL were found to
be lower than that for the PG from the same strain Pcc BR1, with
2.9 3.0 3.1 3.2 3.3 3.4 3.5
-1 these cations (except for Ba2+ ) [13].
1000/T (K ) The presence of Ca2+ significantly (P < 0.001) influences ther-
malstability of PL at temperatures 60 and 70 ◦ C. According to
Fig. 3. Arrhenius plot to calculate activation energy ‘Ed ’ for irreversible thermal
inactivation/denaturation of PL using relationship ln Kd = −Ed /RT, where R (gas con- thermalstability data obtained, the stabilizing effect of 1 mM Ca2+
stant) = 8.314 J/K/mol. was optimal where all the enzyme molecules would be present
28 V.B. Maisuria, A.S. Nerurkar / Biochemical Engineering Journal 63 (2012) 22–30

pH-6 pH-7 pH-8 pH-9 pH-10 Control PL NaCl KCl MgCl2 CaCl2 MnCl2

0.18
(A) 0.18
(B)

ab
ab
0.16 0.16

b
a
0.14 0.14

ab
a
a

c
a

a
0.12

b
b
0.12

bc

bc
a
b

b
ab
a
K d (min )

c
b
b
-1

0.10 0.10

d
a

c
a
a

c
0.08 0.08

b
b

0.06 0.06

a
c

d
b
b
0.04 0.04

f
e
cd
c
d
0.02 0.02

0.00 0.00
40 50 60 70 40 50 60 70

18 50

c
(C)
c

16 (D)
40
14

b
12

b
b

d
30
10
t1/2 (min)

b
a

b
a

e
a

8
ac
a

20
a

a
a

c
a

a
a
a

6
a

b
ab
4
10

bc
bc
c
ab
b
a

ac
b
a

a
2

a
a
a
0 0
40 50 60 70 40 50 60 70
o o
Temperature ( C) Temperature ( C)

Fig. 4. Deactivation rate constant (Kd ) and half-life (t1/2 ) of PL in response of different pHs (A and C) and cations (B and D). The dialyzed PL of Pcc BR1 (control PL, without
any salts) was pre-incubated with each salts [NaCl (0.1 M), KCl (0.15 M), MgCl2 (5 mM), CaCl2 (1 mM), and MnCl2 (1 mM)] individually at 4 ◦ C for 10 min, followed by thermal
stability assay. Bars represent average values of n = 3 independent experiments and error bars represent the standard deviation. One way ANOVA with Tukey’s test was used
to determine significant differences (P < 0.001) among enzyme systems at each temperature. Same alphabets above the bars indicate non-significant difference.

as enzyme–Ca2+ complexes which in agreement with inactivation of the number of noncovalent bonds broken and S* the net
model proposed by Violet and Meunier [30] which is as follows: enzyme/solvent disorder change associated with the N → Tn* tran-
K
sition. The number of non-covalent bonds broken during the Tn*
K 2
1
NCa⇔UCa−→ICa (15) state of enzyme is difficult to assess. In the N state, there is
some ambiguity connected with the bond energies and relative
where NCa is native enzyme–Ca2+
complex, UCa is heat unfolded
importance of various non-covalent bonds. Stabilizing factors of
enzyme–Ca2+ complex and ICa is the irreversible inactivated
the native state of enzyme are normally covalent disulfide bonds,
enzyme–Ca2+ complex.
hydrogen bonds, electrostatic forces of ion pairs, van der Waals
The rate-limiting step for the irreversible heat-inactivation of
and hydrophobic interactions [31]. It is assumed that all the other
enzymes is the formation of an unfolded enzyme (U) state at
hydrogen bonds in native confirmation of enzyme i.e. either the
moderate temperatures, describing irreversible inactivation as a
hydrogen bond donor or the hydrophobic bond acceptor, forms a
two-stage reaction [26]:
hydrogen bond with water molecules in addition to the intramolec-
N ⇔ U → I(andN → Ioverall) (16) ular hydrogen bonds. Since the same group should also hydrogen
bond to water molecules in the unfolded state of enzyme, they
where N is the native conformation for PL, U is the heat induced
should no longer contribute significantly to the conformational
unfolded enzyme and I is the irreversible inactivated PL. Results
stability through hydrogen bonding [26].
described in Table 3 and Fig. 5 show thermodynamic parameters
associated with the formation of a transition state (Tn*) according
to Eq. (17): 3.4.3. Enthalpy–entropy compensation during thermal
∗ deactivation
N ↔ Tn ↔ U (17)
In order to understand the behavior of PL in different physiolog-
Thus, H* and S* are the enthalpy and entropy change respec- ical conditions and the complex process of enzyme deactivation,
tively for the N → Tn* reaction. The H* provides a measure in addition to the deactivation kinetics, thermal inactivation
V.B. Maisuria, A.S. Nerurkar / Biochemical Engineering Journal 63 (2012) 22–30 29

Table 3
Thermal inactivation thermodynamic parameters of PL at temperature range of 40–70 ◦ C.

Enzyme systems H* (kJ mol−1 )† S* (J mol−1 K−1 )† Ed (kJ mol−1 )†

Pectate lyase (PL) 31.2 ± 1.59a


−169.13 ± 4.84 a
33.91 ± 1.59a
PL + 0.1 M NaCl 39.0 ± 2.32b −147.62 ± 6.97b 41.72 ± 2.32b
PL + 0.15 M KCl 35.14 ± 2.07ab −157.6 ± 4.98ab 38.14 ± 1.64ab
PL + 5 mM MgCl2 56.8 ± 1.74c −100.8 ± 2.56c 57.37 ± 3.58c
PL + 1 mM CaCl2 67.88 ± 1.31d −71.0 ± 3.51d 66.74 ± 2.87d
PL + 1 mM MnCl2 61.0 ± 2.01c −90.20 ± 2.79c 62.16 ± 2.01cd
PL + pH 6 12.0 ± 1.46e −226.88 ± 5.45e 15.03 ± 1.76e
PL + pH 7 13.2 ± 2.39ef −225.0 ± 7.55ef 15.91 ± 2.39e
PL + pH 8 16.22 ± 0.52g −218.0 ± 6.41g 19.0 ± 0.52f
PL + pH 9 9.01 ± 1.97e −237.12 ± 6.14e 11.73 ± 1.97e
PL + pH 10 8.6 ± 0.75eh −238.81 ± 2.29eh 11.32 ± 0.75e

Values with same alphabets (a, b, c, d) indicate non-significant difference (P < 0.001).

Values represent mean ± standard deviation of n = 3 independent experiments, analyzed by One way ANOVA with Tukey’s test to determine significant differences among
enzyme systems.

thermodynamic parameters (G*, H*, S*, and Ed ) were deter- 90 ◦ C and 4.7% at 70 ◦ C, compared to 40 ◦ C, which were considered
mined. The change in enthalpy (H*) and entropy (S*) were as unfavorable reaction condition at higher temperature from the
estimated based on the transition state theory. The S* values were thermodynamic viewpoint for PL of B. pumilus DKS1.
obtained negative for all systems used in this study (Table 3), which The kinetic analysis of enzyme inactivation mechanisms is of
is unique property of biocatalyst processes. The solvent and struc- prime importance since it allows better control over biocataly-
tural effects are two important factors which influence the numeric sis processes. In addition, the information about the resistance of
values of H* and S*. The negative values of S* may be due enzymes to thermal denaturation due to the intrinsic contribution
to agregation of the partially unfolded enzyme molecules [32]. As of the polypeptide chain (i.e. hydrophobic interaction, hydrogen
shown in Table 3, the H* and S* values decreased with increase bonding, and ionic stabilization) is also useful [33]. The broad
in alkaline pH conditions (above the pH 8.0) and were significantly nature of the optimum region for the thermostability can be deter-
higher at pH 8.0 compared to other pH conditions (Table 3). Prob- mined based on the different thermodynamic parameters of the
ably at higher pHs, the stable tertiary structure of enzyme gets enzyme with various extrinsic factors. To sum up, these studies
compressed, resulting in lower residual activity [9]. In the case contribute to facilitate better application of the biocatalysts.
of cations, the enthalpy and entropy values decreased in the fol- It has been mentioned by Solbak et al. [34] that under alka-
lowing order: Ca2+ > Mn2+ > Mg2+ > Na+ > K+ (Table 3), which can be line conditions and elevated temperatures (the optimal conditions
attributed to the binding of the Ca2+ ions to the enzyme molecule for scouring), noncellulosic part of cotton fibers mainly pectic sub-
providing compactness to the tertiary structure that results in stances in the form of pectate can rapidly be hydrolyzed by pectate
higher thermostability. lyase. The enzyme based bioscouring treatment for cotton fabrics
The thermal deactivation energy (Ed ) of PL estimated using and degumming of fibers, are generally carried out at temper-
Arrhenius model was significantly higher at pH 8.0 (near to opti- atures above 70 ◦ C and in an alkaline pH, preferably above pH
mum pH 8.5 for the PL activity) and starts decreasing above pH-8.0 8.0 [7]. The thermostable PL of Pcc BR1 is a catalytically efficient
(Table 3). Alternatively, the increase in the deactivation energy enzyme for polygalacturonic acid hydrolysis; also it can act on
in presence of Ca2+ , Mn2+ and Mg2+ suggested that the enzyme
requires more amount of energy to deactivate in presence of diva-
40 °C 50 °C 60 °C 70 °C
lent cations. However, the effect of monovalent cations (Na+ and 100
K+ ) on stability of PL was found to be negligible (Table 3).

95 cc cc
c cb cb
3.4.4. Free energy changes during PL deactivation b cb
b b
bb b a
The Gibbs free energy (G*) measures the combination of a
ba a a a a
90 a a a a b
changes in heat and entropy that occurs during the spontaneity of a a a
G * (kJ mol )

a a a a
-1

a reaction. The resistance of enzyme to thermal denaturation is due a d


a a d a d d
a a a d a a
to the ‘intrinsic’ contribution of polypeptide chain i.e. hydrophobic a a d a
85 a d a d d
a a a
interactions, hydrogen bonding and ionic stabilization. An increase
in S* implies an increase in the number of protein molecules in a
transition activated state, which in turn gives a lower value of G* 80
[19]. The G* values for the PL deactivation are shown in Fig. 5. It
was observed that with increasing the temperature, the free energy 75
change (G*) increased by 5.7% (control PL), 2.31% (PL with Ca2+ ),
2.94% (PL with Mn2+ ) and 3.31% (PL with Mg2+ ) at 70 ◦ C as compared
to 40 ◦ C. In case of pH range studied, the G* increased in the range 70
) l l
of 7.18–7.92% at 70 ◦ C as compared to 40 ◦ C. Therefore, it reiterated l2
(PL NaC M KC gC CaC
l2 Cl2 H6 H7 H8 H9 10
ase .1M 1 5 M M Mn L+ p L+ p L+ p L+ p + pH
that the reaction conditions were most favorable with Ca2+ , Mn2+ , te
ly . M
+ 0 L+ 0 + 5m + 1m + 1m
M P P P P PL
cta PL P PL PL PL
Mg2+ and alkaline condition (pH 8.0) at elevated temperatures from Pe
the thermodynamic point of view. Possibly the divalent cations Enzyme systems
binding have influence on structural stabilization of PL through
stabilization of native state (decrease in the free energy) and/or Fig. 5. The free energy changes during thermal inactivation of PL in presence of dif-
ferent cations and pH conditions. Bars represent mean ± standard deviation of n = 3
destabilization of transition state (increase in the free energy); as
independent experiments, analyzed by One way ANOVA with Tukey’s test to deter-
observed by other authors for peroxidase stability induced by heme mine significant differences (P < 0.001) among enzyme systems at each temperature.
binding [31]. Basu et al. [7] reported that G* increased by 7% at Same alphabets above bars indicate non-significant difference.
30 V.B. Maisuria, A.S. Nerurkar / Biochemical Engineering Journal 63 (2012) 22–30

methylated pectins under the alkaline conditions. Based on its [10] R. Heikinheimo, D. Flego, M. Pirhonen, M.B. Karlsson, A. Eriksson, A. Mae, V. Koiv,
kinetic and thermodynamic parameters, PL of Pcc BR1 can be E.T. Palva, Characterization of a novel pectate lyase from Erwinia carotovora
subsp. carotovora, Mol. Plant Microbe Interact. 8 (1995) 207–217.
expected to be versatile and efficient for industrial application. [11] S.P. Lei, H.C. Lin, L. Heffernan, G. Wilcox, Cloning of the pectate lyase genes
Therefore this enzyme could be utilized efficiently for agro- from Erwinia carotovora and their expression in Escherichia coli, Gene 35 (1985)
industrial pectic waste treatment, as well as industrial degumming 63–70.
[12] J. Sugiura, M. Yasuda, S. Kamimiya, K. Izaki, H. Takahashi, Purification and
treatment of plant fibers such as cotton, jute and ramie. Work on properties of two pectate lyases produced by Erwinia carotovora, J. Gen. Appl.
this aspect is under progress in our laboratory. The feasibility of Microbiol. 30 (1984) 167–175.
obtaining a PL with novel industrial potential is higher from a new [13] V.B. Maisuria, V.A. Patel, A.S. Nerurkar, Biochemical and thermal stabilization
parameters of polygalacturonase from Erwinia carotovora subsp. carotovora
microbial strain not previously reported for this purpose as in this
BR1, J. Microbiol. Biotechnol. 20 (2010) 1077–1085.
case should not be ignored. [14] S. Kamiyama, Y. Itoh, K. Izaki, H. Takahashi, Purification and properties of pec-
tate lyase in Erwinia aroideae, Agric. Biol. Chem. 41 (1977) 975–981.
[15] U.K. Laemmli, Cleavage of structural proteins during the assembly of the head
4. Conclusions
of bacteriophage T4, Nature 227 (1970) 680–685.
[16] J. Sambrook, D.W. Russel, Molecular Cloning—A Laboratory Manual, 3rd ed.,
The present studies explore the important biochemical proper- Cold Spring Harbour Laboratory Press, New York, 2001.
[17] G.P. McMillan, A.M. Barrett, M.C.M. Pérombelon, An isoelectric focusing study
ties of the PL produced by Pcc BR1 for the first time. It demonstrated
of the effect of methyl-esterified pectic substances on the production of extra-
activity on polygalacturonic acid as well as 70% methylated pectin. cellular pectin isoenzymes by soft rot Erwinia spp., J. Appl. Microbiol. 77 (1994)
Its nonlinear inactivation curves at 50 and 60 ◦ C fitted to the 175–184.
three-fraction first-order model. Based on the thermodynamic [18] N. Ortega, S. de Diego, M. Perez-Mateos, M.D. Busto, Kinetic properties and
thermal behaviour of polygalacturonase used in fruit juice clarification, Food
parameters of thermal inactivation of PL, it was found that alkaline Chem. 88 (2004) 209–217.
pH 8.0 and Ca2+ ions favor the thermal stability at 60 and 70 ◦ C. The [19] V. Gohel, D.C. Naseby, Thermalstabilization of chitinolytic enzymes of Pantoea
thermoactive PL of Pcc BR1 has an alkaline pH optimum (pH 8.5), dispersa, Biochem. Eng. J. 35 (2007) 150–157.
[20] H. Tanabe, Y. Kobayashi, Y. Matuo, N.W.F. Nishi, Isolation and fundamental
low activation energy and broad substrate specificity, which sug- properties of endo-pectate lyase isoenzymes from Erwinia carotovora, Agric.
gests that this strain could be utilized effectively as a novel source Biol. Chem. 48 (1984) 2113–2120.
of PL for the pectic waste management and degumming of plant [21] S. Bartling, C. Wegener, O. Olsen, Synergism between Erwinia pectate lyase
isoenzymes that depolymerize both pectate and pectin, Microbiology 141
fibers such as cotton, jute and ramie. (1995) 873–881.
[22] C. Schoedel, A. Collmer, Evidence of homology between the pectate lyase-
Acknowledgement encoding pelB and pelC genes in Erwinia chrysanthemi, J. Bacteriol. 167 (1986)
117–123.
[23] B.G. Klug-Santner, W. Schnitzhofer, M. Vrsanska, J. Weber, P.B. Agrawal,
Author MVB acknowledges the University Grants Commission V.A. Nierstrasz, G.M. Guebitz, Purification and characterization of a new
(New Delhi, India) for a meritorious student research fellowship bioscouring pectate lyase from Bacillus pumilus BK2, J. Biotechnol. 121 (2006)
(UGC sanction no. F.4-1/2006/XI plans/BSR). 390–401.
[24] J. Georis, F. de Lemos Esteves, J. Lamotte-Brasseur, V. Bougnet, B. Devreese,
F. Giannotta, B. Granier, J.M. Frere, An additional aromatic interaction
References improves the thermostability and thermophilicity of a mesophilic family
11 xylanase: structural basis and molecular study, Protein Sci. 9 (2000)
[1] F. Tardy, W. Nasser, J. Robert-Baudouy, N. Hugouvieux-Cotte-Pattat, Com- 466–475.
parative analysis of the five major Erwinia chrysanthemi pectate lyases: [25] S. Nath, A rapid method for determining kinetic parameters of enzymes exhibit-
enzyme characteristics and potential inhibitors, J. Bacteriol. 179 (1997) ing nonlinear thermal inactivation behavior, Biotechnol. Bioeng. 49 (1996)
2503–2511. 106–110.
[2] A. Collmer, N.T. Keen, The role of pectic enzymes in plant pathogenesis, Annu. [26] M.D. Busto, R.K. Owusu Apenten, D.S. Robinson, Z. Wu, R. Casey, R.K. Hughes,
Rev. Phytopathol. 24 (1986) 383–409. Kinetics of thermal inactivation of pea seed lipoxygenases and the effect of
[3] G.S. Hoondal, R.P. Tiwari, R. Tewari, N. Dahiya, Q.K. Beg, Microbial alkaline pecti- additives on their thermostability, Food Chem. 65 (1999) 323–329.
nases and their industrial applications: a review, Appl. Microbiol. Biotechnol. [27] R.M. Daniel, The upper limits of enzyme thermal stability, Enzyme Microb.
59 (2002) 409–418. Technol. 19 (1996) 74–79.
[4] H. Aoyagi, Application of plant protoplasts for the production of useful metabo- [28] R. Srinivas, T. Panda, Enhancing the feasibility of many biotechnological pro-
lites, Biochem. Eng. J. 56 (2011) 1–8. cesses through enzyme deactivation studies, Bioprocess Biosyst. Eng. 21 (1999)
[5] A.S. Meyer, B.P. Dam, H.N. Lærke, Enzymatic solubilization of a pectinaceous 363–369.
dietary fiber fraction from potato pulp: optimization of the fiber extraction [29] Y. Kobayashi, K. Komae, H. Tanabe, R. Matsuo, Approach to maceration mech-
process, Biochem. Eng. J. 43 (2009) 106–112. anism in enzymatic pulping of bast fibers by alkalophilic pectinolytic enzymes
[6] F. Gassara, S.K. Brar, R.D. Tyagi, M. Verma, R.Y. Surampalli, Screening of produced by Erwinia species, Biotechnol. Adv. 6 (1988) 29–37.
agro-industrial wastes to produce ligninolytic enzymes by Phanerochaete [30] M. Violet, J.C. Meunier, Kinetic study of the irreversible thermal denaturation
chrysosporium, Biochem. Eng. J. 49 (2010) 388–394. of Bacillus licheniformis alpha-amylase, Biochem. J. 263 (1989) 665–670.
[7] S. Basu, A. Ghosh, A. Bera, M.N. Saha, D. Chattopadhyay, K. Chakrabarti, Ther- [31] J.K.A. Kamal, D.V. Behere, Kinetic stabilities of soybean and horseradish perox-
modynamic characterization of a highly thermoactive extracellular pectate idases, Biochem. Eng. J. 38 (2008) 110–114.
lyase from a new isolate Bacillus pumilus DKS1, Bioresour. Technol. 99 (2008) [32] R.L. Foster, The Nature of Enzymology, Croom Helm, London, 1980.
8088–8094. [33] P.V. Iyer, L. Ananthanarayan, Enzyme stability and stabilization—aqueous and
[8] S.N. Gummadi, D.S. Kumar, Optimization of chemical and physical parameters non-aqueous environment, Process Biochem. 43 (2008) 1019–1032.
affecting the activity of pectin lyase and pectate lyase from Debary- [34] A.I. Solbak, T.H. Richardson, R.T. McCann, K.A. Kline, F. Bartnek, G. Tomlinson,
omyces nepalensis: a statistical approach, Biochem. Eng. J. 30 (2006) X. Tan, L. Parra-Gessert, G.J. Frey, M. Podar, P. Luginbuhl, K.A. Gray, E.J. Mathur,
130–137. D.E. Robertson, M.J. Burk, G.P. Hazlewood, J.M. Short, J. Kerovuo, Discovery of
[9] G.S.N. Naidu, T. Panda, Studies on pH and thermal deactivation of pectolytic pectin-degrading enzymes and directed evolution of a novel pectate lyase for
enzymes from Aspergillus niger, Biochem. Eng. J. 16 (2003) 57–67. processing cotton fabric, J. Biol. Chem. 280 (2005) 9431–9438.

You might also like