Pii: s0167-4838(00)00232-6
Biochimica et Biophysica Acta 1543 (2000) 275 293
Glucoamylase: structure/function relationships, and protein engineering
JÖrgen Sauer a, Bent W. Sigurskjold b, Ulla Christensen c, Torben P. Frandsen d,
Ekaterina Mirgorodskaya e, Matt Harrison e, Peter Roepstor¡ e, Birte Svensson a;*
a Department of Chemistry, Carlsberg Laboratory, Gamle Carlsberg Vej 10, DK-2500 Copenhagen, Valby, Denmark
b Department of Biochemistry, August Krogh Institute, University of Copenhagen, Universitetsparken 13, DK-2100 Copenhagen Ò, Denmark
c Chemical Laboratory 4, Department of Chemistry, University of Copenhagen, Universitetsparken 5, DK-2100 Copenhagen Ò, Denmark
d Novo Nordisk, Novo Alle¨, DK-2880 Bagsv×rd, Denmark
e Department of Molecular Biology, University of Southern Denmark, Odense University, Campusvej 55, DK-5230 Odense M, Denmark
Received 15 March 2000; received in revised form 31 August 2000; accepted 28 September 2000
Glucoamylases are inverting exo-acting starch hydrolases releasing L-glucose from the non-reducing ends of starch and
related substrates. The majority of glucoamylases are multidomain enzymes consisting of a catalytic domain connected to a
starch-binding domain by an O-glycosylated linker region. Three-dimensional structures have been determined of free and
inhibitor complexed glucoamylases from Aspergillus awamori var. X100, Aspergillus niger, and Saccharomycopsis fibuligera.
The catalytic domain folds as a twisted (K/K)6-barrel with a central funnel-shaped active site, while the starch-binding domain
folds as an antiparallel L-barrel and has two binding sites for starch or L-cyclodextrin. Certain glucoamylases are widely
applied industrially in the manufacture of glucose and fructose syrups. For more than a decade mutational investigations of
glucoamylase have addressed fundamental structure/function relationships in the binding and catalytic mechanisms. In
parallel, issues of relevance for application have been pursued using protein engineering to improve the industrial properties.
The present review focuses on recent findings on the catalytic site, mechanism of action, substrate recognition, the linker
region, the multidomain architecture, the engineering of specificity and stability, and roles of individual substrate binding
subsites. ß 2000 Elsevier Science B.V. All rights reserved.
Keywords: Catalytic base; Binding loop; O-Glycosylated linker; Site-directed mutagenesis; Sequence replacement variant;
Mass spectrometry; Bifunctional inhibitor; Isothermal titration calorimetry; Molecular recognition; Pre-steady-state kinetics;
lase, EC 3.2.1.3) catalyse hydrolysis of K-1,4 and
K-1,6 glucosidic linkages to release L-D-glucose
Glucoamylases (GAs) (1,4-K-D-glucan glucohydro-
from the non-reducing ends of starch and related
poly- and oligosaccharides ([1,2]; for reviews see
[3,4]). Fungal GA is widely used in the manufacture
of glucose and fructose syrups. Although activity
Abbreviations: CD, catalytic domain; GA, glucoamylase;
(kcat/Km) towards the K-1,6 linkage is only 0.2% of
ITC, isothermal titration calorimetry; SBD, starch-binding do-
that for the K-1,4 linkage [1,5 7] this su¤ces to ad-
* Corresponding author. Fax: +45-3327-4708;
versely a¡ect the yield in industrial sacchari¢cation.
E-mail:
[email protected]
This property together with a need for other develop-
0167-4838 / 00 / $ see front matter ß 2000 Elsevier Science B.V. All rights reserved.
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
ments motivated both fundamental and goal-ori-
e¢t from elevated thermostability and enhancement
ented protein engineering of GA.
of activity in the neutral pH range [3,4]. An increase
The GAs constitute glycoside hydrolase family 15
of the glucose yield in sacchari¢cation beyond the
[8,9] and at least 23 primary structures are known
current 96% level might be achieved by suppressing
from ¢lamentous fungi, yeast, eubacteria and archae
the activity of GA on K-1,6 linkages. A very di¡erent
[10,11]. GA has frequently served as a prototype in
way to exploit GA is to utilise the O-glycosylated
investigations on glycoside hydrolases as will be ap-
linker to connect di¡erent domains by generation
parent from the present review. The three-dimension-
of fusion proteins with new combinations of multiple
al structure of the catalytic domain (CD; aa 1 471)
functionalities. Variations on this approach include
of GA (aa 1 616) from Aspergillus awamori var.
fusions of SBD from GA to L-galactosidase for pu-
X100 has been described in detail for native and li-
ri¢cation purposes [37], or to the C-terminus of an
gand-complexed forms [12 17]. Furthermore, pre-
K-amylase to increase its capacity to bind onto and
liminary structure determination was made of wild-
degrade starch granules and other recalcitrant forms
type and mutants of A. niger GA [18] which has 94%
of starch (N. Juge, J.N. Larsen, C.S.M. Furniss, V.
sequence identity to GA from A. awamori var. X100.
Planchot, M.-F. Le Gal-Coe«¡et, D.B. Archer, B.
In addition, the crystal structure of the yeast GA
Svensson, G. Williamson, unpublished). Di¡erent
from Saccharomycopsis ¢buligera, which lacks a
binding polypeptide tails were also added to GA to
starch-binding domain (SBD), has been published
facilitate puri¢cation [38]. Recently GA has been
recently [19]. GA from A. awamori var. X100 folds
used in fusions as a vehicle for the production of
into an (K/K)6-barrel and the C-terminal part (aa
recombinant proteins in A. niger [39].
440 471) of the CD wraps around the (K/K)6-motif
and constitutes the N-terminal part of an O-glycosy-
lated linker (aa 440 508) that connects to a C-termi-
nal SBD (aa 509 616). The conformation of the most
highly O-glycosylated part of the linker (aa 472 508)
2.1. Catalytic domain
is unknown. The structure of SBD from A. niger GA
has been determined by NMR in free form [20] and
The catalytic domain (CD) of GA from A. awa-
bound to L-cyclodextrin, a well-known starch mimic
mori var. X100 contains 13 K-helixes of which 12
[21]. The structure of the entire GA is thus not avail-
form an (K/K)6-barrel. In this fold, six outer and
able. The domain-level organisation of GA has been
six inner K-helixes surround the funnel-shaped active
addressed, however, using di¡erent biophysical tech-
site, constituted by the six highly conserved KCK
niques [22 24].
segments [10,11] that connect the N-termini of the
GA catalyses hydrolysis of glucosidic linkages with
inner with the C-termini of the outer helixes [12
inversion of the anomeric con¢guration [25 29]. Sev-
15] (Fig. 1). The catalytic site includes the general
en subsites were identi¢ed kinetically to participate in
acid and base catalysts Glu179 and Glu400 situated
substrate recognition [30]. The general acid catalyst
at the bottom of a pocket [13,31,32,40]. CDs of GAs
and proton donor Glu179 and the catalytic base
from A. awamori var. X100, A. niger and S. ¢buligera
Glu400 in GA of A. niger are characterised by muta-
share a very similar fold. The S. ¢buligera GA con-
tional analysis [31,32]. Moreover, an array of amino
tains 14 K-helices, 12 of which makes up the (K/K)6-
acid residues that bind directly or via a network of
motif in an organisation identical to that of A. awa-
interactions with substrate at di¡erent subsites has
mori var. X100 and A. niger CD [19]. Two extra
been subjected to site-directed mutagenesis (for re-
short helices protrude from the K-helix connecting
views see [3,4]). Mutational analysis combined with
loops in the ¢rst and the last pair of antiparallel
biophysical techniques gave information on individ-
helices in the fold [19]. The most pronounced di¡er-
ual subsites and has described the impact of the pro-
ence between these GAs, however, is the lack of SBD
tein on transition state stabilisation and di¡erent
in the S. ¢buligera enzyme [10,41]. A single Ser re-
steps in the mechanism of action [33 36].
placement between two very closely related S. ¢buli-
Application of GA in starch industries would ben-
gera GAs is responsible for activity di¡erences [42].
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
Fig. 1. Stereoview of the catalytic domain (aa 1 471) of Aspergillus awamori var. X100 GA complexed with the pseudotetrasaccharide
acarbose (rings A, B, C, D are indicated). The C- and N-termini are indicated together with the side chains of the two catalytic resi-
dues E179 and E400 (from [15]).
2.2. Starch-binding domain
actions with L-cyclodextrin. Mutagenesis studies re-
vealed, however, that they vary only little in a¤nity,
The C-terminal SBD of A. niger was prepared
Ka being 3.6U104 and 1.6U105 M31, respectively
both by proteolysis and in recombinant form, and
[47]. Also the enthalpies and entropies of binding
solution structures of the free and the L-cyclodex-
are similar as analysis of the binding thermodynam-
trin-complexed SBD were determined by NMR spec-
ics of L-cyclodextrin and SBD by isothermal titration
troscopy [20,21]. SBD consists of eight L-strands or-
calorimetry (ITC) did not resolve the two sites
ganised in two L-sheets forming a twisted L-barrel
structure [20,43]. Two starch-binding sites, seen to
accommodate the starch mimic L-cyclodextrin [44
2.3. Linker region
46], are located on opposite sides of the top' of
the domain, i.e., away from the linker attachment
The serine- and threonine-rich O-glycosylated re-
point as seen in Fig. 2 [20,21]. These sites display
gion of A. niger GA (aa 440 508) contains a very
distinctly di¡erent structure and non-covalent inter-
highly O-glycosylated C-terminal segment of about
Fig. 2. Stereoview of the starch binding domain (SBD) from A. niger GA complexed with the starch mimic L-cyclodextrin at the two
binding sites. The C- and N-termini are indicated (from [21]).
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
Fig. 3. Transformed electrospray ionization mass spectrum of the heterogeneously-glycosylated glycopeptide linker (from Asn430 Phe519) prepared from A. niger GA [48] showing at least 17 di¡erent glycoforms. The average mass di¡erence between peaks in these
spectra is 162.1 Da, corresponding to a single hexose residue. The raw mass/charge spectrum prior to transformation to true mass is
shown as an inset, and the charge state of multiply-charged series of peaks is shown. Spectra are normalised to the most intense peak
in each spectrum, and the transformed spectrum has been background-subtracted and smoothed.
30 aa that connects with SBD [10,48]. This particular
in the structural model, but it has been speculated
part of the linker has been attributed roles in stabil-
that this part surrounds CD in a continuation from
ity, secretion, and digestion of raw starch [49 52].
residue 471, to place SBD with one of the two bind-
Mass spectrometric analysis of the peptide Asn430
ing sites near the active site. This resembles the ar-
Phe519 shows a high degree of heterogeneity in the
chitecture of cyclodextrin glucanotransferase in
amount of attached sugars. At least 17 di¡erent gly-
which a homologous C-terminal SBD is situated rel-
coforms can be identi¢ed from the transformed elec-
ative to CD to direct the substrate chain into the
trospray ionisation mass spectra (Fig. 3; M. Harri-
active site via one binding site of SBD and to be
son, P. Roepstor¡, and B. Svensson; unpublished
bound onto soluble or insoluble polymeric substrate
results). Based on a calculated molecular mass of
at the other [57]. The full-length linker is anticipated
the peptide of 8562.28 Da and the experimentally
to be conformationally £exible in accordance with
determined mass of 18 991.3 Da for the glycoform
the formation of 1:1 complexes between a bifunc-
of the lowest molecular mass (Fig. 3), approximately
tional inhibitor and GA [23,24].
63 moles of hexose are attached to the peptide. The
O-glycosidically linked units range from single man-
2.4. Overall structure
nosyl to branched mannotriosyl in wild-type A. niger
GA [53 55]. Heterologous expression of A. niger GA
SBD was earlier shown to be required for degra-
results in large host-dependent variation in the con-
dation of raw starch by GA, the natural G2 form (aa
tent of sugars ranging from hypermannosylation by
1 512) without SBD having very low activity on raw
Saccharomyces cerevisiae to modest over-glycosyla-
starch [58]. Recently, isolated SBD acting on starch
tion by both Pichia pastoris and a laboratory strain
granules together with G2 showed a synergistic e¡ect
of A. niger [50,56].
on the degradation of the insoluble substrate, sug-
The ¢rst part (aa 440 471) of the O-glycosylated
gesting that SBD binds onto starch as an individual
region carries, as seen in the structural model of GA,
entity and disrupts the compact structure of the
about 10 exposed single mannosyl residues [12] which
starch granule facilitating the hydrolysis by CD [59].
together with the two N-glycosidically linked units at
The complete three-dimensional structure of intact
Asp171 and Asp395 form a belt of carbohydrate
GA comprising CD, the linker region, and SBD, is
around the globular CD [12]. The highly O-glycosy-
not known. In an attempt to delineate the relative
lated part of the linker (aa 472 508) is not included
position of CD and SBD, scanning tunnelling
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
Fig. 4. Structures of the bifunctional ligands formed by acarbose and L-cyclodextrin joined by poly(ethyleneglycol) spacers of varying
length. The ligands have either no spacer (L0) or spacers of 14, 36, and 73 Aî (L14, L36, L73), respectively.
microscopy indicated that the two domains are
3. Mechanism of action
around 90 Aî apart [22]. Recently, a series of bifunc-
tional inhibitors, in which the CD speci¢c pseudo-
3.1. Catalytic site
tetrasaccharide inhibitor acarbose and the SBD-
speci¢c ligand L-cyclodextrin were coupled via
The widely accepted mechanism of hydrolysis in-
thioglycoside linkages, was used to further analyse
volves proton transfer to the glycosidic oxygen of the
binding to the di¡erent domains. The bifunctional
scissile bond from a general acid catalyst; formation
molecules were synthesised without and with varying
of an oxocarbenium ion; and a nucleophilic attack of
lengths of poly(ethyleneglycol) spacers connected to
water assisted by a general base catalyst [28,60 62].
the reducing end of acarbose and C6 of a glucose
Glu179 and Glu400 in GA from A. niger have been
ring in L-cyclodextrin shown in Fig. 4 [24]. Four
identi¢ed as the general acid and the general base
di¡erent heterobifunctional inhibitors were demon-
catalyst, respectively, and pH-dependencies of
strated by ITC to bind simultaneously at the active
steady-state kinetic parameters are in accordance
site of CD and one of the binding sites on SBD [23].
with a rate determining hydrolysis step involving
It was thus concluded that in solution the two do-
these two catalytic residues [13,31,32]. Also in accor-
mains of the GA molecule either are in, or can be
dance with this is the observation that mutation of
brought into, close proximity. The sum of enthalpies
Glu400 to Gln results in a reduction of kcat to 3% of
in binding of acarbose and L-cyclodextrin gave essen-
wild-type, showing the marked in£uence of this res-
tially the same value as found for the enthalpy of the
idue on the rate determining step [32].
bifunctional ligands. The binding a¤nities, however,
The GA catalysis occurs with inversion of the
were reduced approximately 105 times compared to
anomeric con¢guration (Fig. 5) in a single displace-
that of acarbose due to strong entropy penalties in
ment mechanism and the gap between the catalytic
binding of double-headed inhibitors [23]. Dynamic
acids is 9.2 Aî as is typical for inverting glycoside
light-scattering measurements on the binding of
hydrolases [28,64 66]. In contrast the distance be-
the bifunctional inhibitors suggested co-operation
tween the catalytic acids in retaining glycoside hydro-
between the domains [24]. The inhibitors were
lases is only 4.8 5.5 Aî and hydrolysis occurs in a
concluded to bind in a bimolecular complex with
double displacement mechanism that includes a co-
occupation of one site at each of the CD and
valent intermediate [28,64]. In this mechanism the
proposed covalent bond between substrate and pro-
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
However, as an important feature of the model, in
agreement with the suggestions made for Hormoconis
resinae GA [78] and in contrast to other Model I-
type suggestions [30,73,76,77], it was found [33,35,36]
that subsite +1 participates in the formation of the
ES-complex and that the transformation of ES to
E*S involves conformational changes, but not the
¢lling of a previously empty subsite 31 by relocation
Fig. 5. The generally accepted catalytic mechanism of GA illus-
of the substrate.
trating the action of the catalytic base E400 (top) and acid
Tables 1 and 2 summarise the results. Further-
E179 (bottom) in the water-assisted hydrolysis of substrate in-
more, it has been shown that the pre-steady-state
volving inversion of the con¢guration of the anomeric carbon.
kinetic results [35] are not in accordance with the
classical model [30,79,80] of GA catalysed reactions,
tein has the consequence that high precision of the
which involves strong non-productive binding of
spatial positioning of the two catalytic groups is nec-
substrates and substrate length independent values
essary for the nucleophilic attack on the glycosidic
of the intrinsic catalytic constant. Since this model
bond [65]. Such a strict geometrical requirement for
is usually used in subsite energy calculations, it has
the catalytic site seems not to apply for the inverting
led to the false general acceptance of subsite +1 of
GA, as it was best illustrated by the elevated activity
GA as the one providing most of the substrate bind-
of GA from A. niger in which the catalytic base was
ing energies [6,30,32,69,81 83].
replaced by cysteine which was subsequently oxidised
As seen from the three-dimensional structure of
to cysteinesulphinic acid [29,66].
GA-inhibitor complexes shown in close-up in Fig. 6
[13,15 18] Trp52 and Trp120 are hydrogen bonded
3.2. Binding mechanism
to the general acid catalyst, Glu179. The Trp52-bond
is 3.04 Aî [16] and is therefore not shown in Fig. 6.
Conserved tryptophan residues are involved in in-
Trp317 and Glu180 are situated on the opposite
teractions of the GAs with substrates and inhibitors
£ank of the active site (Fig. 6), but are not in close
[6,30,67 71] and changes in intrinsic enzyme £uores-
contact. Kinetic results (Tables 1 and 2) show that
cence result as binding occurs. These changes were
the Trp317- and Glu180-mutants react almost iden-
earlier assumed to involve only a tryptophan in sub-
tically, and structure energy minimisation calcula-
site +1 [30], but recent structure and function studies
tions further show that the same loss of Arg305
have shown that in addition to this tryptophan a
and Glu180 hydrogen bonds to the substrate occurs
number of other tryptophans are involved [15
in each of these mutants [36]. The Trp52 Trp120
£ank of the active site has been designated the
Pre-steady-state kinetics analysis of the binding
K-£ank, and the £ank with Trp317 and Glu180 has
mechanism of wild-type and mutant A. niger GA
been designated the L-£ank [36].
has been based on the intrinsic protein £uorescence
Mutations on the L-£ank primarily a¡ect the sec-
changes that occur when substrates and inhibitors
ond reaction step, assumed to be a conformational
bind [33,35,36,72 75]. Formation of complexes with
change, where the substrate obtains the correct posi-
single exponential kinetics was seen in all cases and
tion for catalysis after the initial association. Appar-
analysis of their concentration dependencies all
ently, this step involves the formation of hydrogen
showed results in accordance with a three-step reac-
bonds to Arg305 and to Glu180. This is in excellent
tion mechanism (Model I) of catalysis involving two
agreement with the critical role of Glu180 for the
intermediates: ES, the initial association complex,
induction of a productive conformation of isomal-
and E*S, the Michaelis complex (i.e., the most stable
tose [63]. In spite of the interaction of Trp317 with
enzyme substrate intermediate) [30,33,35,36,72 77].
Glu400, apparently this tryptophan plays no role in
the catalytic step, but exerts its e¡ect in the binding
E S01ES02ESÿ!
of the substrate, particularly assisting in the confor-
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
Results for the reaction of maltose with wild-type forms and mutant glucoamylases (pH 4.5, 8³C)Enzyme
G1, amino acid residues 1 616; G2, amino acid residues 1 512 [58].
mational change bringing the substrate in place for
The Trp52- and the Trp120-mutants at the K-£ank
catalysis. Interestingly, the resulting kinetic e¡ects
show similar changes of the kinetics, since both mu-
obtained when Trp317 is changed to Phe closely par-
tations lead to almost total loss of catalytic turnover.
allels that obtained when Glu180 was changed to
Apparently the correct position of Glu179 for catal-
Gln [35,36]. Furthermore, the changed pattern of
ysis is not obtained in the second reaction step here.
hydrogen bonds between Arg305, Asp309, Tyr306
The mutants, further show stronger substrate bind-
and Glu180, obtained when Trp317 is mutated to
ing than wild-type GA [33,36]. The kinetic parame-
Phe is the same as when Glu180 is mutated to Gln
ters (Tables 1 and 2) are very similar for the wild-
[36]. The mutation of Trp317 clearly a¡ects the posi-
type G1 and G2 forms of GA. Trp52 is situated at
tion of Glu400, but kc is not changed. This supports
the bottom of the active site, it stacks with the
the generally accepted view that the rate determining
C5 C6 part of substrate glucose moiety in subsite
step is in the actual hydrolysis. In the wild-type en-
31 (ring a in Fig. 6) and makes a hydrogen bond
zyme this is the formation of an oxocarbenium ion,
[16] to the catalytic acid, Glu179 (Fig. 6). Therefore
whereas when the assistance of the general base cat-
this tryptophan is in a position where it is expected
alyst is lacking at low pH and in Glu400-mutants it
to be of great importance in the catalytic mechanism.
may change to the nucleophilic attack of water. It is
The k2 values decrease generally more than in the
thus indicated that the catalytic base, Glu400, is not
Trp317- and Glu180-mutants. But this does not re-
involved in that elementary step of the catalysis,
sult in weaker binding, since the k32 values also
which is rate determining.
greatly decrease. The Michaelis complexes with the
Results for the reaction of maltotetraose with wild-type forms and mutant glucoamylases (pH 4.5, 8³C)Enzyme
vFmax([S]) (%) Kd (mM)
G1, amino acid residues 1 616; G2, amino acid residues 1 512 [58].
d From [35].
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
Fig. 6. Stereoview of the active site of glucoamylase from A. awamori var. X100 with bound D-gluco-dihydroacarbose [16]. The four
rings are marked a, b, c and d. Hydrogen bond interactions less than 3.0 Aî are represented by dashed lines, that of Trp52 to Glu179
is 3.04 Aî and therefore does not appear. We de¢ne the K-£ank of the active site as that on the right side of the ligand and the L-£ankas that on the left side in this representation.
Trp52CPhe mutant show Km values (within exper-
The results show that Trp52 plays a role in destabi-
imental error, Km = Kd) approximately one order of
lisation of the Michaelis complex as well as of the
magnitude less that those of the wild-type. This is
association complex. The catalytic rates, as is seen
similar to results obtained on the binding of the in-
from the kc values, are extremely low, and the power
hibitor 1-deoxynojirimycin [34].
of attraction lost here is greater than that gained
An important feature of the Michaelis complex
from the lowering of the Km values. The result is
most probably is the presence of a hydrogen bond
an absolute increase of the transition state energy
between the substrate oxygen of the scissile bond and
barrier. This increase is substantially larger than it
Glu179. The Trp52CPhe and Trp120CPhe muta-
would be if only a compensation of the stronger
tions perturb Glu179 and the observed e¡ects of the
binding in the Michaelis complex was involved. It
mutations on the rearrangement step in which the
is clearly indicated that Trp52 plays important roles
ES-complex transforms into the Michaelis complex,
in binding as well as in catalysis. The di¡erences
E*S, indicate that it is the formation of this bond in
between maltose and longer substrates (Tables 1
the second reaction step, which is impeded. Since the
and 2) appear to be a general phenomenon
complexes formed are only one order of magnitude
weaker than wild-type Michaelis complexes, and the
All in all Model I with a rate determining hydro-
catalytic activities are almost totally lost, it seems as
lysis step is supported by these ¢ndings. It has re-
if in these mutant GAs Glu179 does not make the
cently been suggested [75,85] that product dissocia-
right interactions with the substrate for catalysis. As
tion and not hydrolysis should be rate determining.
seen from Tables 1 and 2, the G2 form of the
This seems highly unlikely, however, in the light of
Trp52CPhe mutant shows K1 values less than or
the known pH-dependency of kc [2,5,32], the weak
equal to those of the wild-type, maltotriose and mal-
binding of glucose [30], and the results showing more
totetraose binding stronger in the ¢rst association
slow, but nevertheless fast second reaction step of all
complex, whereas maltose shows no signi¢cant
of the mutants. A study of the pH-dependence of the
change. The Km values decrease, which means stron-
pre-steady-state kinetic parameters of the interaction
ger binding. This is in accordance with the classical
of wild-type GA and maltose further has shown that
theory of enzyme function [84], which points out the
the k2 value is slightly increasing in the range pH 5
bene¢t of a maximal power of attraction of the tran-
7, while it is not decreasing with the pK of the cata-
sition state, a transition state stabilisation', at the
lytic acid as would be the case, if this step was the
expense of the attraction of the Michaelis complex.
actual hydrolysis step (U. Christensen, unpublished).
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
modi¢cation of the pseudotetrasaccharide exerted
considerable impact on the a¤nity which was re-
The pseudotetrasaccharide acarbose (Fig. 7) binds
duced to Ka = 3.2U107 M31. The determination of
with high a¤nity (Ka = 1012 M31) to GA [71,86] at
the bound conformation of D-gluco-dihydroacarbose
subsites 31 through +3 [15,17]. This is a substan-
by transferred NOE NMR experiments indicates that
tially stronger interaction than what is usually ob-
the inhibitor is bound in a conformation that is sim-
served between carbohydrates and proteins [87].
ilar to the conformation in the crystal structure of
The free energy of binding is composed of approx-
the complex, but di¡erent from the predominant so-
imately two-thirds enthalpy and one-third entropy.
lution conformation of the free inhibitor [88]. L-ido-
The pseudodisaccharide acarviosine (Fig. 7), which
Dihydroacarbose (Fig. 7) was also obtained in the
comprises the ¢rst two units of acarbose at the
preparation of the D-gluco isomer of reduced acar-
non-reducing end, binds with a much lower a¤nity
bose and this inhibitor with an inverted chair con-
of Ka = 7.8U106 M31 [86]. This means that the two
formation of the hydrogenated valeinamine ring
glucose units of acarbose are responsible for a con-
showed an even weaker binding of 2.2U105 M31
siderable amount of the total binding free energy.
D-gluco-Dihydroacarbose (Fig. 7) prepared by hydro-
A study of acarbose and 1-deoxynojirimycin (Fig.
genation of the valeinamine ring in acarbose showed
7) binding to a number of GA mutants with single
a similar binding as acarbose except for a subsite 31
amino acid substitutions in CD has been reported
distorted chair conformation [16,17]. The structural
[34]. There are vast changes in the a¤nity for acar-
bose ranging from a slight increase (Ka = 1013 M31)
down to a¤nities of KaW103 M31. The large reduc-
tions in a¤nity occurred for mutations in residues
directly involved in hydrogen bonds with the sub-
strate or in stacking interactions and also for muta-
tions in groups involved in stabilisation of either
substrate binding residues or catalytic residues. Other
alterations had little or no e¡ect on acarbose bind-
ing. Most of the mutants had almost wild-type a¤n-
ity for 1-deoxynojirimycin, while only two seem to
have abolished binding of this inhibitor completely
Novel thioglucoside disaccharide analogue inhibi-
tors of GA were synthesised and characterised in-
volving kinetic measurements, molecular modelling,
and detailed NMR conformational analysis which
revealed important structural details underlying e¤-
cient GA-oligosaccharide complexation [89,90].
Transferred NOE NMR measurements of methyl
5P-thio-4-N-K-maltoside, in complex with GA,
showed that GA bound this analogue in a conforma-
tion in the area close to the global energy minimum
[91]. Kinetic evaluation showed e¤cient competitive
inhibition of GA with a Ki value of only 4 WM by
this analogue [89]. The importance of GA-oligosac-
charide interactions at subsite +1 was emphasised by
the higher Ki values of a series of 5-thio-D-glucopyr-
Fig. 7. Competitive inhibitors of glucoamylase: acarbose (A),D-gluco-dihydroacarbose (B), L-ido-dihydroacarbose (C), methyl
anosylarylamines compared to methyl 5P-thio-4-N-K-
acarviosinide (D), and 1-deoxynojirimycin (E).
maltoside [89,90]. These compounds are, however,
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
more strongly bound by GA than the substrate
4.2. Replacement of the catalytic base by Cys-SO2H
p-nitrophenyl-K-D-glucopyranoside [6,29], showing
the importance of a ring sulphur and nitrogen in
In order to enable engineering of the distance be-
the interglycosidic linkage for e¤cient GA inhibition
tween the catalytic general acid, Glu179, and base,
Glu400, in GA, the catalytic base, which was previ-
ously found to best tolerate substitution [32], was
mutagenised to Cys. The side-chain was further at-
4. Protein engineering production
tempted to be carboxyalkyl extended by modi¢cation
of the SH group by reaction with various haloalkyl
4.1. Recombinant GA production
carboxyl acids. This, however, failed to produce al-
kylated Cys at position 400, but fortuitously, a GA
E¤cient heterologous expression of GA encoding
derivative was obtained with activity superior to that
cDNA from A. awamori (identical to GA from
of wild-type GA [29,66]. Subsequent chemical analy-
A. niger) was established in the methylotrophic yeast
sis involving HPLC-separation of peptide fragments
P. pastoris [56]. To describe the in£uence of host
prepared by treatment with the Endo-LysC protease
dependent posttranslational modi¢cation on enzy-
followed by matrix-assisted laser desorption/ionisa-
matic and structural features, the recombinant pro-
tion mass spectrometry combined with post-source
tein produced in P. pastoris was compared to GA
decay analysis enabled the unequivocal identi¢cation
produced in the related hosts, S. cerevisiae and
of the product as Cys400-SO2H [92]. Thus the thiol
A. niger [56]. Recombinant GA produced in all three
group had undergone spontaneous oxidation to the
hosts showed essentially identical catalytic proper-
sulphinic acid in the presence of the alkylating re-
ties, but di¡ers in thermostability [56]. Molecular
agent. Attempts to repeat the oxidation by a mixture
mass determination using matrix-assisted laser de-
of iodine and bromine successfully resulted in a GA
sorption/ionisation mass spectrometry and neutral
derivative with elevated activity [66]. Remarkably,
sugar analysis revealed small, but signi¢cant varia-
depending on the substrate, kcat increased up to
tions in the glycosylation of the three recombinant
300% of the value for wild-type, an e¡ect most pro-
GA forms. GA produced in S. cerevisiae thus has the
nounced for longer K-1,4 maltooligosaccharides and
highest content of carbohydrate with a measured
K-1,6 isomaltooligosaccharides (Table 3). In contrast,
molecular mass of 83.869 Da. GA produced in
the a¤nity decreased (i.e., Km increased) with oligo-
A. niger and P. pastoris had average molecular
saccharide length. Similarly the a¤nity (Ka) for acar-
mass of 82.839 Da and 82.327 Da, respectively [56].
bose decreased by a factor of approximately 103,
Kinetic parameters for hydrolysis of malto- and isomaltooligosaccharides by wild-type and the Cys400-SO2H derivativeaSubstrate
kcat/Km (s31 mM31)
kcat/Km (s31 mM31)
aData from [29].
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
while that of acarviosine decreased only by a factor
loops simultaneously. If substitution was made at
of 15, and the a¤nity for 1-deoxynojirimycin was
either loop 3 or loop 5, the mutant GAs had very
slightly raised by a factor of 3 (Table 4). Disaccha-
low activity compared to wild-type and the double
rides of varying bond type were also hydrolysed with
loop mutant [7].
di¡erent retained activities, some having superior kcat
(nigerose (K-1,3) and maltose), retained (kojibiose
4.4. Combination of protein engineering and substrate
(K-1,2)) or inferior kcat (isomaltose). Only for niger-
ose did the Km increase signi¢cantly [66]. Thus while
the mutation of Glu400 to Cys dramatically de-
The energetics of protein carbohydrate complexa-
creased the activity this was with most substrates
tion can, in principle, be described using two exper-
more than restored by oxidation of the Cys to
imental approaches: site-directed mutagenesis of se-
Cys-SO2H. It is currently not known whether this
lected active-site substrate binding residues and,
behaviour is unique to GA, applies to inverting gly-
molecular recognition of deoxygenated or otherwise
coside hydrolases, or to other retaining and inverting
chemically derivatised substrate analogues by the
glycosidases in general.
wild-type enzyme. Both strategies have been exten-
sively used to investigate protein substrate complex-
4.3. Replacement by homologue sequences at
ation in GA from A. niger [3,63,93,94] and by com-
bining the two approaches, direct identi¢cation of
interacting pairs of atoms or groups of atoms and
The architecture of the CD in GA has six loops
support for enzyme induced substrate conformation-
that connect the inner with the outer cylinder' of
al changes have been achieved.
K-helices that create the substrate binding site and
Mapping of the substrate key polar groups was
carry the catalytic residues [12]. The clearly best con-
done through wild-type GA recognition of deoxygen-
served parts of the GA sequences are in these loop
ated K-1,4- [93 95] and K-1,6-linked [63,96] substrate
analogues. From these studies it became evident that
GAs di¡er in stability and also show di¡erences in
GA catalysed hydrolysis is strongly dependent on
substrate speci¢city. The most extreme case of spe-
charged protein-substrate hydrogen bonds from GA
ci¢city variation is GA from Hormoconis resinae that
to OH-3, OH-4P, and OH-6P as summarised in Fig. 8
has only 50-fold higher activity for K-1,4 compared
for K-1,4-linked substrates [93 95] and to OH-4,
to K-1,6 linkages as opposed to most GAs showing
OH-4P, and OH-6P in K-1,6-linked substrates [63,
500 103 fold higher activity for the K-1,4-linked sub-
96]. Substitution of these particular OH groups by
strates [78]. The H. resinae GA contains unusual se-
hydrogen or methoxy groups are accompanied by
quences in certain loops. Engineering mutation of
loss in transition-state stabilisation of 11 19 kJ/mol
loops 3 and 5 in A. niger GA to mimic the character-
which is typical for groups involved in charged hy-
istics of GA from H. resinae was possible without
drogen bond interactions. Elimination of the corre-
signi¢cant loss of activity and accompanied by de-
sponding protein hydrogen-bond partners, Arg54,
crease of the relative speci¢city for the K-1,4 over the
Asp55, and Arg305, interacting with OH-4P, OH-6P,
K-1,6 bond. However, mutation was required at both
and OH-3/4, respectively, similarly resulted in dra-
matic losses in transition-state stabilisation of up to
22 kJ/mol [6,82].
In addition, by coupling site-directed mutagenesis
Ki for inhibitors of wild-type and Cys400-SO2H GAsa
of GA and substrate molecular recognition structural
details of transition-state interactions between GA
and substrate can be detected [63,94]. Analysis of
Glu180CGln GA using a series of deoxygenated
maltose and isomaltose analogues thus demonstrated
transition-state stabilising hydrogen bonds between
aData from [29].
Glu180 and OH-2 in maltose [94] and OH-4 and
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
Fig. 8. Representation of the energy contributions of the interactions between carbohydrate and protein as determined by mutagenesis
and substrate analogue studies; the corresponding vvGV values are shown in italics and bold, respectively.
OH-3 in isomaltose [63]. The identi¢cation of these
site +1 interactions which optimises transition-state
interactions was shown prior to the acquisition of the
stabilisation through charged hydrogen bonds to
high resolution three-dimensional complexes demon-
substrate OH-4P and 36P in subsite 31 [63]. GA
strating the powerful combination of protein and
substrate interactions at one subsite thus have critical
substrate engineering. By using conformationally
impact on crucial hydrogen bond formation at adja-
biased substrate analogues such investigations have
cent subsites.
been extended even further to obtain details on GA
catalysed hydrolysis of an K-1,6-linked substrate [63].
4.5. Linker region variants
The analysis of wild-type and variant GA catalysed
hydrolysis of conformationally biased isomaltoside
The question of the distance between the two do-
analogues demonstrated that Glu180 induces a con-
mains in GA has been addressed through a series of
formational change of the bound substrate via sub-
linker mutants in which the highly O-glycosylated
Fig. 9. Schematic representation of the constructed linker variants. The catalytic (CD; aa 1 466) and starch binding (SBD; aa 509
616) domains connected through the O-glycosylated (aa 467 508) linker region and sequence alignment of the linker regions of other
GAs used to replace the A. niger linker. Asterisks above the alignment indicate known O-glycosylation sites in A. niger GA. Numbers
in brackets refer to amino acid position in wild-type sequences.
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
sequence (aa 468 508) was replaced by shorter link-
heterogeneously glycosylated as in the wild-type
ers from related GAs [97] or by a proline rich un-
GA when produced by various homologous and het-
natural sequence (Fig. 9). Mass spectrometric analy-
erologous hosts [56]. The rather high mass increase
sis of the variants (Fig. 10) determined apparent
for the variant with the Rhizopus oryzae linker sug-
masses of 73 413 and 90 793 Da for the Humicola
gests that N-linked glycosylation did occur on the
grisea and Rhizopus oryzae GA linker variants, re-
sequon ThrGlyAsn introduced in this variant at the
spectively, and showed that the linker region was
C-terminal end of the linker just prior to SBD and
Fig. 10. MALDI-TOF spectra of wild-type and linker variant GAs. The mass of the singly charged species are indicated. Calculated
values for the molecular mass of the polypeptide chain are shown in brackets.
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
that additional glycosylation may also have occurred
cloning and expression of the target gene could alter-
elsewhere (Figs. 9 and 10).
natively ¢nd a thermostable GA. Only a few thermo-
Compared to wild-type GA these linker variant
stable wild-type GAs have, however, been reported
GAs have lower conformational stability in solutions
including the enzymes of H. grisea var. thermoidea
containing denaturants and also lower heat stability,
[105], A. fumigatus [106], and Thermomyces lanugino-
but normal catalytic activity towards raw starch and
sus [107]. Furthermore, gene sequences are known of
soluble oligo- and polysaccharide substrates (J. Sau-
thermostable GAs from Clostridium sp. G0005 [108],
er, T. Christensen, B.W. Sigurskjold, B. Svensson,
Methanococcus jannaschii [109], and Thermoanaero-
unpublished). A minimum length of 17 aa of the
bacterium thermosaccharolyticum [110]. These en-
replacing sequence seemed to be required for produc-
zymes show approximately 40% sequence identity
tion of functional GA variant protein in P. pastoris
to GAs from Aspergilli. A common molecular fea-
(J. Sauer, B. Svensson, unpublished). The essentially
ture for the thermophilic enzymes seems to be the
normal function of the variant with these short link-
lack of helices 9, 10, and 11 of the (K/K)6-barrel
ers as compared to the longer linker of 38 aa in wild-
CD [110]. Information from these and forthcoming
type GA, suggests that the natural linker £exibly
sequences of thermostable GAs may guide future ra-
positions CD relative to SBD. It would be desirable
tional protein engineering towards a GA that exhib-
to identify the structural elements that control the
its activity and stability at elevated temperatures.
correct positioning, secondly, to ¢nd conditions
under which a single structural conformer is predom-
inant, and thirdly, to identify which factors can force
5. Conclusion and perspectives
the association between the domains apart.
GA has been very thoroughly described using pro-
4.6. Engineering of industrial properties
tein engineering techniques to study fundamental
questions in the mechanism and speci¢city and for
GA is an industrially extremely important enzyme,
the further development of GA for industrial appli-
used in the enzymatic conversion of starch into high
cations. Whereas GA research in many respects have
glucose and fructose syrups [98]. Although GAs from
resulted in forefront discoveries and improved under-
most sources are unstable at higher temperatures in-
standing in the broad ¢eld of structure and function
dustrial sacchari¢cation is currently performed at
of glycoside hydrolases, the development of econom-
60³C. Development of a thermostable GA, capable
ical industrial enzymes have bene¢ted less from the
of performing industrial sacchari¢cation at elevated
vast amount of protein engineering data. Desirable
temperatures, would thus be of signi¢cant impor-
improvements from an industrial point of view have
tance to the starch processing industry. Small
been mentioned above. The major questions to ad-
achievements towards a thermostable GA were ful-
dress on the basic knowledge include (i) the three-
¢lled through protein engineering of the enzymes
dimensional structure of an intact GA with both
from A. niger and A. awamori (for a review see
SBD and CD, (ii) insight into the co-operation be-
[99]). Several approaches, including replacement of
tween CD and SBD and the possible relevance of
glycines in K-helices [100], elimination of fragile
variation in domain level organisation for degrada-
Asp-X bonds [101] and substitution of asparagine
tion of solid starches, (iii) further investigation of the
in Asn Gly sequences [102] have been pursued using
reaction mechanism and its dependence on the sub-
site-directed mutagenesis. The most successful strat-
strate to settle the discussion on under which condi-
egy applied, however, seems to be engineering of
tions hydrolysis or product release would be rate
additional disulphide bonds into the molecule, in-
limiting and to further understand the interplay be-
creasing the Tm value of GA by up to 4³C
tween enzyme and substrate/inhibitors, and (iv) time-
resolved structural analysis of GA and substrates
Screening of thermophilic bacteria and subsequent
during the catalysis.
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
Crystal structure of glucoamylase from Aspergillus awamori
var. X100 to 2.2-Aî resolution, J. Biol. Chem. 267 (1992)
The authors are grateful for support from the
19291 19298.
[13] E.M. Harris, A.E. Aleshin, L.M. Firsov, R.B. Honzatko,
Danish Research Councils' Committee on Biotech-
Re¢ned structure for the complex of 1-deoxynojirimycin
nology (Grant no. 9502014).
with glucoamylase from Aspergillus awamori var. X100 to
2.4-Aî resolution, Biochemistry 32 (1993) 1618 1626.
[14] A.E. Aleshin, C. Ho¡man, L.M. Firsov, R.B. Honzatko,
Re¢ned crystal structures of glucoamylase from Aspergillus
awamori var. X100, J. Mol. Biol. 238 (1994) 575 591.
[15] A.E. Aleshin, L.M. Firsov, R.B. Honzatko, Re¢ned struc-
[1] K. Hiromi, Z.I. Hamauzu, K. Takahashi, S. Ono, Kinetic
ture for the complex of acarbose with glucoamylase from
studies on gluc-amylase. II. Competition between two types
Aspergillus awamori var. X100 to 2.4-Aî resolution, J. Biol.
of substrate having K-1,4 and K-1,6 glucosidic linkage,
Chem. 269 (1994) 15631 15639.
J. Biochem. (Tokyo) 59 (1966) 411 418.
[16] B. Sto¡er, A.E. Aleshin, L.M. Firsov, B. Svensson, R.B.
[2] K. Hiromi, K. Takahashi, Z.I. Hamauzu, S. Ono, Kinetic
Honzatko, Re¢ned structure for the complex of D-gluco-di-
studies on gluc-amylase. 3. The in£uence of pH on the rates
hydroacarbose with glucoamylase from Aspergillus awamori
of hydrolysis of maltose and panose, J. Biochem. (Tokyo) 59
var. X100 to 2.2 Aî resolution: dual conformations for ex-
(1966) 469 475.
tended inhibitors bound to the active site of glucoamylase,
[3] T.P. Frandsen, H.-P. Fierobe, B. Svensson, Engineering spe-
FEBS Lett. 358 (1995) 57 61.
ci¢city and stability in glucoamylase from Aspergillus niger,
[17] A.E. Aleshin, B. Sto¡er, L.M. Firsov, B. Svensson, R.B.
in: L. Alberghina (Ed.), Protein Engineering in Industrial
Honzatko, Crystallographic complexes of glucoamylase
Biotechnology, Harwood Academic, Amsterdam, 1999, pp.
with maltooligosaccharide analogs: relationship of stereo-
chemical distortions at the nonreducing end to the catalytic
[4] P.J. Reilly, Protein engineering of glucoamylase to improve
mechanism, Biochemistry 35 (1996) 8319 8328.
industrial properties. A Review, Starch/Sta«rke 51 (1999)
[18] B. Sto¡er, T.P. Frandsen, B. Svensson, M. Gajhede, X-ray
structure of an active site mutant glucoamylase,
[5] M.R. Sierks, B. Svensson, Protein engineering of the relative
Tyr48CTrp, from Aspergillus niger, in: The Carbohydrate
speci¢city of glucoamylase from Aspergillus awamori based
Bioengineering Meeting,, Elsinore, Denmark, 1995, Abstract
on sequence similarities between starch-degrading enzymes,
Protein Eng. 7 (1994) 1479 1484.
[19] J. Sevcik, A. Solovicova, E. Hostinova, J. Gasperik, K.S.
[6] T.P. Frandsen, T. Christensen, B. Sto¡er, J. Lehmbeck, C.
Wilson, Z. Dauter, Structure of glucoamylase from Saccha-
Dupont, R.B. Honzatko, B. Svensson, Mutational analysis
romycopsis ¢buligera at 1.7 Aî resolution, Acta Crystallogr. D
of the roles in catalysis and substrate recognition of argi-
Biol. Crystallogr. 54 (1998) 854 866.
nines 54 and 305, aspartic acid 309, and tryptophan 317
[20] K. Sorimachi, A.J. Jacks, M.-F. Le Gal-Coe«¡et, G. William-
located at subsites 1 and 2 in glucoamylase from Aspergillus
son, D.B. Archer, M.P. Williamson, Solution structure of
niger, Biochemistry 34 (1995) 10162 10169.
the granular starch binding domain of glucoamylase from
[7] H.-P. Fierobe, B.B. Sto¡er, T.P. Frandsen, B. Svensson,
Aspergillus niger by nuclear magnetic resonance spectrosco-
Mutational modulation of substrate bond-type speci¢city
py, J. Mol. Biol. 259 (1996) 970 987.
and thermostability of glucoamylase from Aspergillus awa-
[21] K. Sorimachi, M.-F. Le Gal-Coe«¡et, G. Williamson, D.B.
mori by replacement with short homologue active site se-
Archer, M.P. Williamson, Solution structure of the granular
quences and thiol/disul¢de engineering, Biochemistry 35
starch binding domain of Aspergillus niger glucoamylase
(1996) 8696 8704.
bound to L-cyclodextrin, Structure 5 (1997) 647 661.
[8] B. Henrissat, A classi¢cation of glycosyl hydrolases based on
[22] G.F.H. Kramer, A.P. Gunning, V.J. Morris, N.J. Belshaw,
amino acid sequence similarities, Biochem. J. 280 (1991)
G. Williamson, Scanning tunnelling microscopy of Aspergil-
lus niger glucoamylase, J. Chem. Soc., Faraday Trans. 89
[9] B. Henrissat, A. Bairoch, New families in the classi¢cation
(1993) 2595 2602.
of glycosyl hydrolases based on amino acid sequence simi-
[23] B.W. Sigurskjold, T. Christensen, N. Payre, S. Cottaz, H.
larities, Biochem. J. 293 (1993) 781 788.
Driguez, B. Svensson, Thermodynamics of binding of heter-
[10] P.M. Coutinho, P.J. Reilly, Structural similarities in gluco-
obidentate ligands consisting of spacer-connected acarbose
amylase by hydrophobic cluster analysis, Protein Eng. 7
and L-cyclodextrin to the catalytic and starch-binding do-
(1994) 749 760.
mains of glucoamylase from Aspergillus niger shows that
[11] P.M. Coutinho, P.J. Reilly, Glucoamylase structural, func-
the catalytic and starch-binding sites are in close proximity
tional and evolutionary relationships, Proteins 29 (1997)
in space, Biochemistry 37 (1998) 10446 10452.
[24] N. Payre, S. Cottaz, C. Boisset, R. Borsali, B. Svensson, B.
[12] A.E. Aleshin, A. Golubev, L.M. Firsov, R.B. Honzatko,
Henrissat, H. Driguez, Dynamic light scattering evidence for
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
a ligand-induced motion between the two domains of glu-
revisiae by the addition of poly(aspartic acid) tails, Enzyme
coamylase G1 of Aspergillus niger using heterobivalent sub-
Microb. Technol. 15 (1993) 593 600.
strate analogues, Angew. Chem. Int. Ed. 38 (1999) 974 977.
[39] D.B. Archer, D.J. Jeenes, D.A. MacKenzie, Strategies for
[25] C.E. Weill, R.J. Burch, J.W. Van Dyk, An K-amyloglucosi-
improving heterologous protein production from ¢lamentous
dase that produces L-glucose, Cereal Chem. 31 (1954) 150
fungi, Antonie Van Leeuwenhoek 65 (1994) 245 250.
[40] B. Svensson, A.J. Clarke, I. Svendsen, H. MÖller, Identi¢ca-
[26] J.H. Pazur, T. Ando, The hydrolysis of glucosyl oligosaccha-
tion of carboxylic acid residues in glucoamylase that partic-
rides with K-D-(1,4) and K-D-(1,6) bonds by fungal amyloglu-
ipate in catalysis and substrate binding, Eur. J. Biochem. 188
cosidases, J. Biol. Chem. 235 (1960) 297 302.
(1990) 29 38.
[27] D.E. Koshland Jr., Stereochemistry and the mechanism of
[41] E. Hostinova, J. Balanova, J. Gasperik, The nucleotide se-
enzymatic reactions, Biol. Rev. 28 (1953) 416 436.
quence of the glucoamylase gene GLA1 from Saccharomy-
[28] J.D. McCarter, S.G. Withers, Mechanisms of enzymatic gly-
copsis ¢buligera KZ, FEMS Microbiol. Lett. 67 (1991) 103
coside hydrolysis, Curr. Opin. Struct. Biol. 4 (1994) 885
[42] A. Solovicova, T. Christensen, E. Hostinova, J. Gasperik, J.
[29] H.-P. Fierobe, A.J. Clarke, D. Tull, B. Svensson, Enzymatic
Sevcik, B. Svensson, Structure function relationships in glu-
properties of the cysteinesul¢nic acid derivative of the cata-
coamylases encoded by variant Saccharomycopsis ¢buligera
lytic-base mutant Glu400CCys of Glucoamylase from As-
genes, Eur. J. Biochem. 264 (1999) 756 764.
pergillus awamori, Biochemistry 37 (1998) 3753 3759.
[43] A.J. Jacks, K. Sorimachi, M.F. Le Gal-Coe¡et, G. William-
[30] K. Hiromi, M. Ohnishi, A. Tanaka, Subsite structure and
son, D.B. Archer, M.P. Williamson, 1H and 15N assignments
ligand binding mechanism of glucoamylase, Mol. Cell. Bio-
and secondary structure of the starch-binding domain of
chem. 51 (1983) 79 95.
glucoamylase from Aspergillus niger, Eur. J. Biochem. 233
[31] M.R. Sierks, C. Ford, P.J. Reilly, B. Svensson, Catalytic
(1995) 568 578.
mechanism of fungal glucoamylase as de¢ned by mutagene-
[44] N.J. Belshaw, G. Williamson, Interaction of L-cyclodextrin
sis of Asp176, Glu179 and Glu180 in the enzyme from As-
with the granular starch binding domain of glucoamylase,
pergillus awamori, Protein Eng. 3 (1990) 193 198.
Biochim. Biophys. Acta 1078 (1991) 117 120.
[32] T.P. Frandsen, C. Dupont, J. Lehmbeck, B. Sto¡er, M.R.
[45] N.J. Belshaw, G. Williamson, Speci¢city of the binding do-
Sierks, R.B. Honzatko, B. Svensson, Site-directed mutagen-
main of glucoamylase 1, Eur. J. Biochem. 211 (1993) 717
esis of the catalytic base glutamic acid 400 in glucoamylase
from Aspergillus niger and of tyrosine 48 and glutamine 401,
[46] B.W. Sigurskjold, B. Svensson, G. Williamson, H. Driguez,
both hydrogen-bonded to the Q-carboxylate group of gluta-
Thermodynamics of ligand binding to the starch-binding
mic acid 400, Biochemistry 33 (1994) 13808 13816.
domain of glucoamylase from Aspergillus niger, Eur. J. Bio-
[33] K. Olsen, U. Christensen, M.R. Sierks, B. Svensson, Reac-
chem. 225 (1994) 133 141.
tion mechanisms of Trp120CPhe and wild-type glucoamy-
[47] M.P. Williamson, M.-F. Le Gal-Coe«¡et, K. Sorimachi,
lases from Aspergillus niger. Interactions with maltooligodex-
C.S.M. Furniss, D.B. Archer, G. Williamson, Function of
trins and acarbose, Biochemistry 32 (1993) 9686 9693.
conserved tryptophans in Aspergillus niger glucoamylase 1
[34] C.R. Berland, B.W. Sigurskjold, B. Sto¡er, T.P. Frandsen,
starch binding domain, Biochemistry 36 (1997) 7535 7539.
B. Svensson, Thermodynamics of inhibitor binding to mu-
[48] B. Svensson, K. Larsen, I. Svendsen, E. Boel, The complete
tant forms of glucoamylase from Aspergillus niger deter-
amino acid sequence of the glycoprotein, glucoamylase G1,
mined by isothermal titration calorimetry, Biochemistry 34
from Aspergillus niger, Carlsberg Res. Commun. 48 (1983)
(1995) 10153 10161.
[35] U. Christensen, K. Olsen, B.B. Sto¡er, B. Svensson, Sub-
[49] C.B. Libby, C.A. Cornett, P.J. Reilly, C. Ford, E¡ect of
strate binding mechanism of Glu180CGln, Asp176CAsn,
amino acid deletions in the O-glycosylated region of Asper-
and wild-type glucoamylases from Aspergillus niger, Bio-
gillus awamori glucoamylase, Protein Eng. 7 (1994) 1109
chemistry 35 (1996) 15009 15018.
[36] T. Christensen, B.B. Sto¡er, B. Svensson, U. Christensen,
[50] T. Semimaru, M. Goto, K. Furukawa, S. Hayashida, Func-
Some details of the reaction mechanism of glucoamylase
tional analysis of the threonine- and serine-rich Gp-I domain
from Aspergillus niger; kinetic and structural studies on
of glucoamylase I from Aspergillus var. kawachi, Appl. En-
Trp52CPhe and Trp317CPhe mutants, Eur. J. Biochem.
viron. Microbiol. 61 (1995) 2885 2890.
250 (1997) 638 645.
[51] T. Christensen, B. Svensson, B.W. Sigurskjold, Thermody-
[37] L.J. Chen, C. Ford, A. Kusnadi, Z.L. Nikolov, Improved
namics of reversible and irreversible unfolding and domain
adsorption to starch of a L-galactosidase fusion protein con-
interactions of glucoamylase from Aspergillus niger studied
taining the starch-binding domain from Aspergillus gluco-
by di¡erential scanning and isothermal titration calorimetry,
amylase, Biotechnol. Prog. 7 (1991) 225 229.
Biochemistry 38 (1999) 6300 6310.
[38] I. Suominen, C. Ford, D. Stachon, H. Heimo, M. Nieder-
[52] M. Goto, M. Tsukamoto, I. Kwon, K. Ekino, K. Furukawa,
auer, H. Nurmela, C. Glatz, Enhanced recovery and puri¢-
Functional analysis of O-linked oligosaccharides in threo-
cation of Aspergillus glucoamylase from Saccharomyces ce-
nine/serine-rich region of Aspergillus glucoamylase by ex-
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
pression in mannosyltransferase-disruptants of yeast, Eur. J.
lus awamori by oxidation of the Glu400CCys catalytic-base
Biochem. 260 (1999) 596 602.
mutant to cysteinesul¢nic acid, Biochemistry 37 (1998) 3743
[53] A. Gunnarsson, B. Svensson, B. Nilsson, S. Svensson, Struc-
tural studies on the O-glycosidically linked carbohydrate
[67] A.J. Clarke, B. Svensson, Identi¢cation of an essential tryp-
chains of glucoamylase G1 from Aspergillus niger, Eur. J.
tophanyl residue in the primary structure of glucoamylase
Biochem. 145 (1984) 463 467.
G2 from Aspergillus niger, Carlsberg Res. Commun. 49
[54] B. Svensson, K. Larsen, A. Gunnarsson, Characterization of
(1984) 559 566.
glucoamylase G2 from Aspergillus niger, Eur. J. Biochem.
[68] A.J. Clarke, B. Svensson, The Role of tryptophanyl residues
154 (1986) 497 502.
in the function of Aspergillus niger glucoamylase G1 and G2,
[55] G. Williamson, N.J. Belshaw, M.P. Williamson, O-Glycosyl-
Carlsberg Res. Commun. 49 (1984) 111 122.
ation in Aspergillus glucoamylase; conformation and role in
[69] M.R. Sierks, C. Ford, P.J. Reilly, B. Svensson, Site-directed
binding, Biochem. J. 282 (1992) 423 428.
mutagenesis at the active site Trp120 of Aspergillus awamori
[56] H.-P. Fierobe, E. Mirgorodskaya, T.P. Frandsen, P. Roep-
glucoamylase, Protein Eng. 2 (1989) 621 625.
stor¡, B. Svensson, Overexpression and characterization of
[70] L.M. Firsov, K.N. Neustroev, A.E. Aleshin, C.M. Metzler,
Aspergillus awamori wild-type and mutant glucoamylase se-
D.E. Metzler, R.D. Scott, B. Sto¡er, T. Christensen, B.
creted by the methylotrophic yeast Pichia pastoris: compar-
Svensson, NMR spectroscopy of exchangeable protons of
ison with wild-type recombinant glucoamylase produced us-
glucoamylase and of complexes with inhibitors in the 9 15-
ing Saccharomyces cerevisiae and Aspergillus niger as hosts,
ppm range, Eur. J. Biochem. 223 (1994) 293 302.
Protein Expr. Purif. 9 (1997) 159 170.
[71] B. Svensson, M.R. Sierks, Roles of the aromatic side chains
[57] D. Penninga, B.A. van der Veen, R.M. Knegtel, S.A. van
in the binding of substrates, inhibitors, and cyclomalto-oli-
Hijum, H.J. Rozeboom, K.H. Kalk, B.W. Dijkstra, L.
gosaccharides to the glucoamylase from Aspergillus niger
Dijkhuizen, The raw starch binding domain of cyclodextrin
probed by perturbation di¡erence spectroscopy, chemical
glycosyltransferase from Bacillus circulans strain 251, J. Biol.
modi¢cation, and mutagenesis, Carbohydr. Res. 227 (1992)
Chem. 271 (1996) 32777 32784.
[58] B. Svensson, T.G. Pedersen, I. Svendsen, T. Sakai, M. Otte-
[72] M.R. Sierks, C. Ford, P.J. Reilly, B. Svensson, Functional
sen, Characterization of two forms of glucoamylase from
roles and subsite locations of Leu177, Trp178 and Asn182 of
Aspergillus niger, Carlsberg Res. Commun. 47 (1982) 55
Aspergillus awamori glucoamylase determined by site-di-
rected mutagenesis, Protein Eng. 6 (1993) 75 79.
[59] S.M. Southall, P.J. Simpson, H.J. Gilbert, G. Williamson,
[73] A. Tanaka, M. Ohnishi, K. Hiromi, Stopped-£ow kinetic
M.P. Williamson, The starch-binding domain from gluco-
studies on the binding of gluconolactone and maltose to
amylase disrupts the structure of starch, FEBS Lett. 447
glucoamylase, Biochemistry 21 (1982) 107 113.
(1999) 58 60.
[74] K. Olsen, B. Svensson, U. Christensen, Stopped-£ow £uo-
[60] M.L. Sinnott, Catalytic mechanisms of enzymatic glycosyl
rescence and steady-state kinetic studies of ligand-binding
transfer, Chem. Rev. 90 (1990) 1171 1202.
reactions of glucoamylase from Aspergillus niger, Eur. J.
[61] A. Konstantinidis, M.L. Sinnott, The interaction of 1-£uoro-
Biochem. 209 (1992) 777 784.
D-glucopyranosyl £uoride with glucosidases, Biochem. J. 279
[75] S.K. Natarajan, M.R. Sierks, Identi¢cation of enzyme-sub-
(1991) 587 593.
strate and enzyme-product complexes in the catalytic mech-
[62] Y. Tanaka, W. Tao, J.S. Blanchard, E.J. Hehre, Transition
anism of glucoamylase from Aspergillus awamori, Biochem-
state structures for the hydrolysis of K-D-glucopyranosyl £u-
istry 35 (1996) 15269 15279.
oride by retaining and inverting reactions of glycosylases,
[76] M. Ohnishi, K. Hiromi, Kinetic studies on the interaction of
J. Biol. Chem. 269 (1994) 32306 32312.
Rhizopus glucoamylase with maltodextrin and maltotriose,
[63] T.P. Frandsen, B.B. Sto¡er, M.M. Palcic, S. Hof, B. Svens-
utilizing the absorbance change near 300 nm, Carbohydr.
son, Structure and energetics of the glucoamylase isomaltose
Res. 61 (1978) 335 341.
transition-state complex probed by using modeling and de-
[77] M. Ohnishi, T. Matsumoto, T. Yamanaka, K. Hiromi, Bind-
oxygenated substrates coupled with site-directed mutagene-
ing of isomaltose and maltose to the glucoamylase from
sis, J. Mol. Biol. 263 (1996) 79 89.
Aspergillus niger, as studied by £uorescence spectrophotom-
[64] G. Davies, B. Henrissat, Structures and mechanisms of gly-
etry and steady-state kinetics, Carbohydr. Res. 204 (1990)
cosyl hydrolases, Structure 3 (1995) 853 859.
[65] S.L. Lawson, W.W. Wakarchuk, S.G. Withers, E¡ects of
[78] R. Fagerstro«m, Subsite mapping of Hormoconis resinae glu-
both shortening and lengthening the active site nucleophile
coamylases and their inhibition by gluconolactone, J. Gen.
of Bacillus circulans xylanase on catalytic activity, Biochem-
Microbiol. 137 (1991) 1001 1008.
istry 35 (1996) 10110 10118.
[79] K. Hiromi, Interpretation of dependency of rate parameters
[66] H.-P. Fierobe, E. Mirgorodskaya, K.A. McGuire, P. Roep-
on the degree of polymerization of substrate in enzyme-cat-
stor¡, B. Svensson, A.J. Clarke, Restoration of catalytic ac-
alyzed reactions. Evaluation of subsite a¤nities of exo-en-
tivity beyond wild-type level in glucoamylase from Aspergil-
zyme, Biochem. Biophys. Res. Commun. 40 (1970) 1 6.
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
[80] K. Hiromi, Y. Nitta, C. Numata, S. Ono, Subsite a¤nities of
glucoamylase probed by deoxygenated substrates, Biochem-
glucoamylase: examination of the validity of the subsite
istry 31 (1992) 8972 8977.
theory, Biochim. Biophys. Acta 302 (1973) 362 375.
[94] M.R. Sierks, B. Svensson, Kinetic identi¢cation of a hydro-
[81] M.M. Meagher, Z.L. Nikolov, P.J. Reilly, Subsite mapping
gen bonding pair in the glucoamylase-maltose transition
of Aspergillus niger glucoamylase I and II with malto- and
state complex, Protein Eng. 5 (1992) 185 188.
isomaltooligosaccharides, Biotechnol. Bioeng. 34 (1989) 681
[95] K. Bock, H. Pedersen, The substrate speci¢city of the en-
zyme amyloglucosidase (AMG). Part I. Deoxy derivatives,
[82] M.R. Sierks, B. Svensson, Functional roles of the invariant
Acta Chem. Scand. 41 (1987) 617 628.
aspartic acid 55, tyrosine 306, and aspartic acid 309 in glu-
[96] R.U. Lemieux, U. Spohr, M. Bach, D.R. Cameron, T.P.
coamylase from Aspergillus awamori studied by mutagenesis,
Frandsen, B.B. Sto¡er, B. Svensson, M.M. Palcic, Chem-
Biochemistry 32 (1993) 1113 1117.
ical mapping of the active site of glucoamylase of Aspergil-
[83] K. Hiromi, Kinetic studies on the interaction of glucoamy-
lus niger, Can. J. Chem. 74 (1996) 319 335.
lase with substrates and analogues, in: Y. Takehiko (Ed.),
[97] B. Svensson, K.S. Bak-Jensen, M.T. Jensen, J. Sauer, K.W.
Enzyme Chemistry and Molecular Biology of Amylases and
Rodenburg, Studies on structure, function, and protein en-
Related Enzymes, CRC Press, Boca Raton, FL, 1995, pp.
gineering of starch-degrading enzymes, J. Appl. Glycosci.
46 (1999) 51 65.
[84] L. Pauling, Chemical achievement and hope for the future,
[98] B.C. Saha, J.G. Zeikus, Microbial glucoamylases: Bio-
Am. Scientist 36 (1948) 51 58.
chemical and biotechnological features, Sta«rke 41 (1989)
[85] M.R. Sierks, B. Svensson, Catalytic mechanism of gluco-
amylase probed by mutagenesis in conjunction with hydro-
[99] C. Ford, Improving operating performance of glucoamy-
lysis of K-D-glucopyranosyl £uoride and maltooligosaccha-
lase by mutagenesis, Curr. Opin. Biotechnol. 10 (1999)
rides, Biochemistry 35 (1996) 1865 1871.
[86] B.W. Sigurskjold, C.R. Berland, B. Svensson, Thermody-
[100] H.M. Chen, Y. Li, T. Panda, F.U. Buehler, C. Ford, P.J.
namics of inhibitor binding to the catalytic site of glucoa-
Reilly, E¡ect of replacing helical glycine residues with ala-
mylase from Aspergillus niger determined by displacement
nines on reversible and irreversible stability and production
titration calorimetry, Biochemistry 33 (1994) 10191 10199.
of Aspergillus awamori glucoamylase, Protein Eng. 9 (1996)
[87] E.J. Toone, Structure and energetics of protein-carbohydrate
complexes, Curr. Opin. Struct. Biol. 4 (1994) 719 728.
[101] H.M. Chen, C. Ford, P.J. Reilly, Identi¢cation and elimi-
[88] T. Weimar, B.O. Petersen, B. Svensson, B.M. Pinto, Deter-
nation by site-directed mutagenesis of thermolabile aspartyl
mination of the solution conformation of D-gluco-dihydro-
bonds in Aspergillus awamori glucoamylase, Protein Eng. 8
acarbose, a high a¤nity inhibitor bound to glucoamylase by
(1995) 575 582.
transferred NOE NMR spectroscopy, Carbohydr. Res. 326
[102] H.M. Chen, C. Ford, P.J. Reilly, Substitution of asparagine
(2000) 50 55.
residues in Aspergillus awamori glucoamylase by site-di-
[89] J.S. Andrews, T. Weimar, T.P. Frandsen, B. Svensson, B.M.
rected mutagenesis to eliminate N-glycosylation and inacti-
Pinto, Novel disaccharides containing sulfur in the ring and
vation by deamidation, Biochem. J. 301 (1994) 275 281.
nitrogen in the interglycosidic linkage. Conformation of 5P-
[103] M.J. Allen, P.M. Coutinho, C.F. Ford, Stabilization of
thio-4-N-maltoside bound to glucoamylase and its activity as
Aspergillus awamori glucoamylase by proline substitution
competitive inhibitor, J. Am. Chem. Soc. 117 (1995) 10799
and combining stabilizing mutations, Protein Eng. 11
(1998) 783 788.
[90] K.D. Randell, T.P. Frandsen, B. Sto¡er, M.A. Johnson, B.
[104] Y. Li, P.M. Coutinho, C. Ford, E¡ect on thermostability
Svensson, B.M. Pinto, Synthesis and glycosidase inhibitory
and catalytic activity of introducing disul¢de bonds into
activity of 5-thioglucopyranosylamines. Molecular modeling
Aspergillus awamori glucoamylase, Protein Eng. 11 (1998)
of complexes with glucoamylase, Carbohydr. Res. 321 (1999)
[105] L.R.O. Tosi, H.F. Francisco, J.A. Jorge, Puri¢cation an
[91] E. Mirgorodskaya, H.-P. Fierobe, B. Svensson, P. Roep-
characterization of an extracellular glucoamylase from the
stor¡, Mass spectrometric identi¢cation of a stable catalytic
thermophilic fungi Humicola grisea var. thermoidea, Can. J.
cysteinesul¢nic acid residue in an enzymatically active chemi-
Microbiol. 39 (1993) 846 852.
cally modi¢ed glucoamylase mutant, J. Mass Spectrom. 34
[106] W. Brandani, R.M. Peralta, Puri¢cation and characteriza-
(1999) 952 957.
tion of a thermostable glucoamylase from Aspergillus fumi-
[92] T. Weimar, B. Sto¡er, B. Svensson, B.M. Pinto, Complexes
gatus, Can. J. Microbiol. 44 (1998) 493 497.
of glucoamylase with maltoside heteroanalogues: Bound li-
[107] D.-C. Li, Y.-J. Yang, Y.-L. Peng, C.-Y. Shen, Puri¢cation
gand conformations by use of transferred NOE experiments
and characterization of extracellular glucoamylase from the
and molecular modeling, Biochemistry 39 (2000) 300 306.
thermophilic Thermomyces lanuginosus, Mycol. Res. 102
[93] M.R. Sierks, K. Bock, S. Refn, B. Svensson, Active site
(1998) 568 572.
similarities of glucose dehydrogenase, glucose oxidase, and
[108] H. Ohnishi, H. Kitamura, T. Minowa, H. Sakai, T. Ohta,
BBAPRO 36304 12-12-00
J. Sauer et al. / Biochimica et Biophysica Acta 1543 (2000) 275 293
Molecular cloning of a glucoamylase gene from a thermo-
Scott, N.S.M. Geoghagen, J.C. Venter, Complete genome
philic Clostridium and kinetics of the cloned enzyme, Eur. J.
sequence of the methanogenic archaeon, Methanococcus
Biochem. 207 (1992) 413 418.
jannaschii, Science 273 (1996) 1058 1073.
[109] C.J. Bult, O. White, G.J. Olsen, L. Zhou, R.D. Fleisch-
[110] A. Ducki, O. Grundmann, L. Konermann, F. Mayer, M.
mann, G.G. Sutton, J.A. Blake, L.M. FitzGerald, R.A.
Hoppert, Glucoamylase from Thermoanaerobacterium ther-
Clayton, J.D. Gocayne, A.R. Kerlavage, B.A. Dougherty,
mosaccharolyticum: Sequence studies and analysis of mac-
J.F. Tomb, M.D. Adams, C.I. Reich, R. Overbeek, E.F.
romolecular architecture of the enzyme, J. Gen. Appl. Mi-
Kirkness, K.G. Weinstock, J.M. Merrick, A. Glodek, J.L.
crobiol. 44 (1998) 327 335.
BBAPRO 36304 12-12-00
Source: http://www.sut.ac.th/iat/biotech/Montarop/amylase/glucoamylase.pdf
Matrons Board Report – Quarter 3 2012 Paper for Board of Directors on 27 March 2013 Director Lisa Knight, Director of Patient Care/Chief Nurse Paper prepared by: Jane Naish, Deputy Chief Nurse/Head of Quality Jon White, Senior Nurse (Practice Development and Rapid Response)
Int J High Dilution Res 2014; 13(48): 207-226 Original article ‘Paradoxical pharmacology': therapeutic strategy used by the ‘homeopathic pharmacology' for more than two centuries Marcus Zulian Teixeira School of Medicine, University of São Paulo, São Paulo, Brazil ABSTRACT