ArticlePDF Available

Structural, Functional and Dynamical Responses of Protein in Restricted Environment Imposed by Macromolecular Crowding

Authors:

Abstract and Figures

The inter-cellular environment is known to be very different from the environment where most of the elementary biological processes are studied in the laboratory. As a result, there was a considerable effort on cell mimicking either by confinement or by introducing macromolecular crowding. In the present study dextran of varying sizes has been used to crowd the environment of a protein, human serum albumin (HSA), and its structure, dynamics, and activity were studied as a function of crowder concentration. By employing bulk and single molecular level spectroscopic measurement, we elucidate the overall structure and local microsecond dynamics of HSA. Further, we have attempted to correlate these structural changes with its activity.
Content may be subject to copyright.
Structural, Functional, and Dynamical Responses of a Protein in a
Restricted Environment Imposed by Macromolecular Crowding
Nilimesh Das and Pratik Sen*
Department of Chemistry, Indian Institute of Technology Kanpur, Kanpur, Uttar Pradesh 208 016, India
*
SSupporting Information
ABSTRACT: The intercellular environment is known to be very
dierent from the environment where most of the elementary
biological processes are studied in the laboratory. As a result, there
was a considerable eort on cell mimicking either by connement
or by introducing macromolecular crowding. In the present study,
dextran of varying sizes has been used to crowd the environment
of a protein, human serum albumin (HSA), and its structure,
dynamics, and activity were studied as a function of crowder
concentration. By employing bulk and single molecular level
spectroscopic measurements, we elucidate the overall structure
and local microsecond dynamics of HSA. Further, we have
attempted to correlate these structural changes with its activity.
Most of the biological studies are being carried out in vitro
as performing experiments inside a living cell is always a
challenge.
1,2
Although researchers maintain a physiological pH
for in vitro studies, for every practical purpose, the cellular
environment is dierent from the dilute buer solutions.
3
One
of the main dierences is the presence of high concentrations
(up to 300400 g L1) of macromolecules in the biological
environment, creating a restricted environment.
4,5
It has been
estimated that up to 30% of the cellular volume is occupied by
such macromolecules that make the free space inside the cell
rather limited.
68
This is known as the excluded volume eect.
Also, inside a cell, the viscosity is about three times higher than
that of the dilute buer solutions.
13,14
For all of these reasons,
it is necessary to modify the buer solution to closely resemble
the cell environment. This could be done either by oering a
conned environment to the system or by adding some
synthetic substance, commonly known as a crowder, to the
solution. There have been several theoretical and experimental
investigations in the past on how crowders can aect the
protein structure and stability,
1519
foldingunfolding path-
way,
20,21
binding kinetics,
22
aggregation behavior,
23
etc.
Usually, the eccentric behavior of proteins in the crowded
environment is explained on the basis of this excluded volume
eect. However, recent works suggest that the protein
crowder interaction may also be responsible for the unique
structural, dynamical, and functional response of protein in the
crowded environment.
5,9,10
While the excluded volume eect is
purely stabilizing, the chemical interaction may be stabilizing
or destabilizing. Using E. coli cytoplasm as a model crowding
agent and chymotrypsin inhibitor 2 (CI2) as the protein, the
Pielak group suggests that the destabilizing chemical
interactions between the cytoplasm of E. coli and CI2
overcome the stabilizing hard-core repulsions.
11
Molecular
dynamics simulation by Feig and co-workers on the same
system showed that the eect of crowding primarily depends
on the nature of proteincrowder interactions.
12
Biological macromolecules are of various sizes, shapes, and
properties, which have a profound eect on the behavior of a
protein.
24,25
Therefore, it is essential to investigate the eect of
crowding on protein as exerted by biological macromolecules
in in vivo conditions. Note that the cell environment is
crowded with macromolecules of dierent sizes. Ribosomes,
molecular chaperons, GroEL, GroES, etc. are of comparatively
higher sizes, whereas substances like nucleic acids, carbohy-
drates, etc. are small in size. On the other hand, the size of
proteins present in the cellular environment may vary from a
few kDa to several hundreds of kDa. Thus, to understand their
eect, crowding by dierent sized crowders is necessary. Feigs
group reported the inuence of crowders of various sizes on
the hydration structure and dynamics of proteins.
26
In another
report, they have shown that the size distributions of
macromolecules are important factors for in vivo crowding
eects.
27
In this context, we intend to investigate the behavior
(structure, dynamics, and function) of proteins in the presence
of crowders of various sizes, taking human serum albumin
(HSA) as the model protein and dierently sized dextrans as
the model crowder. Dextran is one of the most commonly used
synthetic crowder, which is a rod-shaped polysaccharide with
multiple numbers of glucose molecules.
28,29
The molecular
weight of dextran depends on the number of constituent sugar
moieties, and dextran is usually named based on the molecular
weight.
Received: May 28, 2018
Revised: September 10, 2018
Published: September 28, 2018
Article
pubs.acs.org/biochemistry
Cite This: Biochemistry 2018, 57, 60786089
© 2018 American Chemical Society 6078 DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
Downloaded via INDIAN INST OF TECHNOLOGY KANPUR on October 27, 2018 at 15:06:03 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
HSA is the most abundant transport protein in the human
body.
30,31
It is a large multidomain protein with three distinct
domains, namely, domains I, II, and III.
32
All of these domains
are responsible for binding and transporting dierent classes of
molecules like drugs, enzymes, hormones, carbohydrates,
etc.
3337
There are several reports on the behavior of HSA
in the crowded environment in recent years.
3844
Singh et al.
monitored the tryptophan uorescence intensity of HSA as a
function of the extent of crowding.
38
They have concluded that
the quenching of tryptophan uorescence by a crowder is
mainly static in nature and that dextran-6 exhibits maximum
quenching owing to its highest excluded volume by dextran-6.
Biswas et al. have studied the domain movement of HSA
through FRET analysis as a function of crowder concentration
and observed a gradual decrease of the interdomain distance
between domain-I and domain-II and between domain-II and
domain-III for dextran-40 and dextran-70.
39,40
On the other
hand, for dextran-6, there was a surprising increase in the
distance between domain-I and domain-II.
39
In another
contribution, they reported that the solvation dynamics of
domain-I of HSA depends highly on the concentration of the
crowder, owing mainly to the stiness of the protein matrix.
41
In a recent article, Samanta et al. concluded that in the
presence of polyethylene glycol (PEG), another type of
popular crowder, the α-helicity of HSA increases to a
considerable extent.
42
Examining the intrinsic tryptophan
uorescence, they also concluded that HSA becomes more
prone to unfold in the presence of PEG.
43
In spite of the considerable amount of studies on the
structural features of protein in the crowded environment, the
conformational uctuation dynamics and activity of HSA in the
presence of a crowder are not explored. To note here, it is well-
accepted that the structure of a protein is not rigid; rather, it is
an average of several closely related structures known as
dierent conformational states of the protein.
45
The
interconversion between these states is known to be conforma-
tional uctuation dynamics of the protein.
4547
This conforma-
tional uctuation dynamics serves a key role in protein
functions like catalysis, drug binding, hormonal activities,
etc.
48,49
Naturally, it is also important to study such structural
dynamics of a protein in the crowded environment. As this
conformational motion of the proteins is uncorrelated, a bulk
measurement cannot provide information about the dynamics,
and one has to perform the single molecular level study. In our
previous contributions, we have reported the conformational
dynamics of domain-III of HSA has a time constant of 8 μs,
50
and that of domain-I has a time scale of 27 μs using single
molecular level uorescence correlation spectroscopy
(FCS).
51,52
We have also studied the eect of thermal and
chemical denaturation on the dynamics using the same
technique.
50
In the present contribution, we have employed FCS to
investigate the eect of crowders on the structural transition
and conformational uctuation dynamics and attempted to
relate it to the activity of HSA inside the crowding milieu. For
this purpose, we have selectively tagged tyrosine-411 residue
located in domain-III of HSA by a uorescent marker p-
nitrophenyl coumarin ester (NPCE). We have used three
dierent macromolecular crowders, namely, dextran-6, dex-
tran-40, and dextran-70, for the present study.
MATERIALS AND METHODS
Materials. We purchased human serum albumin (HSA,
fatty acid-free), coumarin-343, 4-dimethylamino pyridine
(DMAP), N,N-dicyclohexylcarbodiimide (DCC), 4-nitrophe-
nol, 4-nitrophenyl acetate, dextran-6, dextran-40, and dextran-
70 from Sigma-Aldrich and used as received. Analytical grade
disodium hydrogen phosphate and sodium dihydrogen
phosphate were purchased from Merck, India, and used to
prepare 50 mM buer (pH 7.4). Dialysis membrane tubing (14
kDa cuto) was obtained from Sigma-Aldrich and used after
removing the glycerol and sulfur compounds according to the
procedure given by Sigma-Aldrich. Centrifugal lter units
(Amicon Ultra, 10 kDa cuto) have been purchased from
Merck Millipore, Germany. HPLC grade dimethyl sulfoxide
(DMSO) and dichloromethane (DCM) were purchased from
S. D. Fine Chemicals Limited, India, and used after distillation.
Synthesis of p-Nitrophenyl Coumarin Ester (NPCE).
The synthesis of NPCE was done according to the reported
procedure
51
following common esterication methodology.
53
Briey, an equimolar amount (0.38 mmol) of coumarin-343, 4-
nitrophenol, and 4-dimethylamino pyridine (DMAP) was
taken in 5 mL of dichloromethane, and the mixture was
stirred in an ice bath for 10 min under a nitrogen atmosphere.
Then 0.38 mmol of N,N-dicyclohexylcarbodiimide (DCC) was
added slowly. The solution was stirred at 0 °C in an ice bath
for an additional 20 min and then at 20 °C for 24 h. The
organic layer was then washed rst with 35 mL of 1.2 M HCl
and then with 35 mL of saturated NaHCO3solution. It was
then dried over MgSO4. The residue was suspended in
methanol, and the precipitate was washed with methanol. The
precipitate was collected in dichloromethane, dried under a
vacuum, and characterized by 1H NMR, IR, and mass
spectrometric methods.
51
Protein Labeling and Sample Preparation. NPCE was
tagged to HSA following the procedure by Wang et al. (see
Scheme 1).
54
Briey, 36 mg of HSA was dissolved in 9 mL of
50 mM phosphate buer (pH = 8.0), and 0.3 mg of NPCE was
dissolved in a minimum volume of DMSO. Both were mixed
slowly under stirring conditions. The reaction mixture was
then stirred at the room temperature for 30 h. Dialysis was
done against 500 mL of 1:15 (v/v) DMSO/buer (pH 7.4, 50
mM phosphate buer) solution at 5 °C for 4 days and then
against only phosphate buer for another 4 days to remove any
untagged NPCE. The dialysis medium was replaced every 12 h.
Scheme 1. Tagging of Human Serum Albumin (HSA) by
NPCE
a
a
The molecular structure of NPCE is also shown. The structure le of
HSA is downloaded from the Protein Data Bank (PDB ID: 1ha2).
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6079
The labeled protein was concentrated by centrifuging the
sample using a 10 kDa cutofffiltration unit. For steady-state
and lifetime measurements, the protein concentration was
maintained around 5 μM, and for circular dichroism
measurements, the concentration was about 3 μM. For single
molecular level measurements, the protein concentration was
kept around 50 nM. To thermal denaturation, HSA has been
kept at 75 °C for 30 min.
Instrumentation. A commercial UVvisible spectropho-
tometer (UV-2450, Shimadzu, Japan) and spectrouorimeter
(FluoroMax-4, Jobin-Yvon, USA) were used for recording
steady-state absorption and emission spectra, respectively.
Circular dichroism spectra were recorded on a commercial CD
spectrometer (J-815, Jasco, Japan). Time-resolved uorescence
was collected using a commercial TCSPC setup (LifeSpec II,
Edinburgh Instruments, U.K.). For lifetime measurements,
peak counts of 8000 were collected at the magic angle
condition by exciting the samples with a 442 nm diode laser
(EPL-445, Edinburgh Instruments, U.K.). Instrument response
function (IRF), which is measured to be 110 ps, is used for
deconvolution of the uorescence transients. All of the
experiments were done at 298 K unless stated otherwise.
The transients are tted with Fast software (Edinburgh
Instruments, U.K.) using the following equation:
55
I
tf t
() exp
i
n
ii
1τ
=
i
k
j
j
j
j
j
y
{
z
z
z
z
z(1)
In eq 1,I(t) is the uorescence intensity at the time t, and fiis
the relative contribution to the lifetime component τi. The
average uorescence lifetime is calculated using the following
equation:
f
f
i
n
ii
i
n
i
avg 1
1
τ
τ
=
Σ
Σ
=
=(2)
For the activity measurement, we have monitored the
formation of p-nitrophenol (λabs = 400 nm, ε400 = 17 700 M1
cm1)
5658
as a function of time as produced by the enzymatic
action of the HSA on p-nitrophenyl acetate. One unit of
activity is dened as the liberation of 1.0 μMp-nitrophenol per
minute for a 3 min experiment.
Fluorescence Correlation Spectroscopic Measure-
ments. A home-built FCS set up has been used for the single
molecular level uorescent measurement. The detailed
description of the set up can be found in our previous
reports.
50,51,59,60
The excitation source is a 405 nm continuous
wave laser (5 mW, Optoelectronics Tech. Co. Ltd., China). A
60×water immersion objective with a numerical aperture
(NA) of 1.2 was used to focus the light into the sample placed
on a clean coverslip onto the sample platform of an inverted
microscope (IX-71, Olympus, Japan). For all of the measure-
ments, the light was focused at a distance of 40 μm from the
surface of the coverslip. The same objective collects the
emitted photons, and after passing through a dichroic mirror
(ZT405rdc, Omega Optical Inc., USA) and an emission lter
(FSQ-GG455, Newport, USA), these photons are focused on a
multimode ber patch chord of 25 μm diameter (M67L01 25
mm 0.10 NA, ThorLabs, USA). The optical ber carries the
photons to the single photon counting avalanche photodiode
(SPCM-AQRH-13-FC, Excelitas, USA). These photon counts
are then sent to the autocorrelator card (FLEX990EM-12D,
Correlator.com, USA) to generate the autocorrelation
function, G(τ), and displayed on a LabView platform.
The autocorrelation is the self-similarity of uorescence
intensity at dierent times and can be mathematically
described as
GFt Ft
Ft
() () ( )
()2
τδδ τ
=⟨+
⟨⟩ (3)
In the above equation, F(t)is the average uorescence
intensity, and δF(τ) and δF(t+τ) are the uctuations in
uorescence intensity around the mean value at time tand (t+
τ).
55,61
For a single component system, where all of the
particles have the same diusion coecient, the diusion time
(τD) can be obtained by tting the autocorrelation function
G(τ) using the following equation:
55
GN
() 111
D
1
2D
1/2
ττ
τ
τ
ωτ
=+ +
i
k
j
j
j
j
j
y
{
z
z
z
z
z
i
k
j
j
j
j
j
y
{
z
z
z
z
z(4)
In the above equation, Nis the number of particles in the
observation volume, and ω= l/r is the depth to diameter ratio
of the three-dimensional Gaussian volume. If the diusing
species undergoes any other process leading to an additional
uorescence uctuation having an amplitude, A, and a time
scale, τR, the modied correlation function can be written as
55
GNA() 111 1exp
DDR
1
2
1/2
ττ
τ
τ
ωτ
τ
τ
=+ + +·
i
k
j
j
j
j
j
y
{
z
z
z
z
z
i
k
j
j
j
j
j
y
{
z
z
z
z
z
i
k
j
j
j
j
j
i
k
j
j
j
j
j
y
{
z
z
z
z
z
y
{
z
z
z
z
z
(5)
From the diusion time (τD) and radius of the observation
volume (ωxy), the diusion coecient (Dt) and hydrodynamic
radius (rH) of the molecule can be calculated using the
following couple of equations:
D4
xy
t
2
D
ω
τ
=(6)
r
kT
D6
HB
t
πη
=(7)
Here, kBis the Boltzmann constant, Tis the temperature (298
K in the present study), and ηis the viscosity of the solution.
The structural parameter (ω) associated with the detection
volume was calibrated using a sample of a known diusion
coecient (rhodamine 6G, R6G, in water, Dt= 4.14 ×106
cm2s1).
62
The detection volume of the present setup is
estimated to be 0.5 fL with a transverse radius of 280 nm.
Under the application of external agents, crowders in the
present case, the solution conditions such as refractive index
and viscosity may change signicantly. In turn, the diusion
properties of the molecule therein will also change. The eect
of the viscosity change has been nullied by doing control
experiments at every crowder concentration taking R6G as the
probe. R6G is a rigid molecule and will not undergo any
structural change when exposed to the crowder. So, any change
in its diusion time with a varying crowder concentration will
be solely due to the change of viscosity of the medium. Using
this information and the reported value of the hydrodynamic
radius of R6G, we have calculated the hydrodynamic radius of
HSA at every crowder concentration according to eq 8.
63
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6080
r
r
HHSA
HR6G
DHSA
DR6G
τ
τ
=
(8)
The change in the refractive index is taken care of by
changing the objective collar position and setting it to have the
minimum diusion time for each sample. This way we
maintain the lowest detection volume attainable for each
sample.
RESULTS
Circular Dichroism Spectroscopy. To proceed with the
NPCE-tagged HSA for any further experiment, it is necessary
to conrm that the tagging has not perturbed the secondary
structure of the protein, which we have veried by CD
spectroscopy (see gure S1 of the Supporting Information).
We have analyzed our CD data using CDNN software (http://
gerald-boehm.de) to get the information about the secondary
structural parameters of HSA and listed them in table S1 of the
Supporting Information. The CD spectra of untagged HSA was
recorded in the presence of dierent crowder concentrations as
shown in the gure S2 of the Supporting Information. At high
crowder concentrations (125 g L1), the CD data cannot be
recorded beyond 210 nm because the output voltage becomes
greater than the instrument threshold due to the scattering.
From the measured CD spectra, the structural parameters, i.e.,
the α-helicity, β-sheet, β-turn, and random coil content, of
HSA have been estimated, which are depicted in Figure 1.
With an increasing concentration of dextran-6, we have
observed an almost 88% decrease in α-helicity of HSA from
67.5% to 8%. This decrease in α-helicity is primarily
compensated by the increase of random coil content, which
increases almost 66% from 15% in the absence of dextran-6 to
43.5% in the presence of 200 g L1of dextran-6 (the highest
crowder concentration used in the study). On the contrary, for
both dextran-40 and dextran-70, the change of secondary
structural content of HSA with an increasing crowder
concentration is not that prominent. For both of the cases,
there is a slight but gradual increase of α-helicity with an
increasing crowder concentration. For dextran-40, there is
about a 7% increase in the α-helicity; whereas, for dextran-70,
there is about a 15% increase in α-helicity, mainly at the
expense of the random coil.
Steady-State Absorption and Fluorescence Spectro-
scopic Study. Free NPCE shows absorption and emission
maxima at 447.0 and 489.0 nm, respectively. Upon tagging to
HSA, the absorption and emission maxima of NPCE blue-
shifted to 440.0 and 480.0 nm (gure S3 of the Supporting
Information). These values are in accordance with our
previous report.
44
The emission spectra of NPCE-tagged
HSA at dierent crowder concentrations are shown in gure
S4 of the Supporting Information. As can be seen, the emission
maximum does not change in the presence of crowders.
Fluorescence Lifetime Measurement. All of the
uorescence transients (in the presence and absence of a
crowder) are best tted by a three exponential function. For
native HSA, the average lifetime of NPCE is found to be 3.59
ns with components of 0.95 ns (6.9%), 2.9 ns (47.9%), and
4.74 ns (45.2%). With the increase of crowder concentration,
the lifetime gradually decreases for all of the crowders used in
this study (see gure S5 of the Supporting Information for the
raw data and Figure 2 for the variation). In the presence of 200
gL
1crowder, for all three cases, the average lifetime decreases
to 3.43 ns. Thus, for all cases, the extent of decrease is
negligible (less than 5%).
Fluorescence Correlation Spectroscopic Study. The
uorescence autocorrelation curve of free NPCE in the buer
is satisfactorily tted with a single diusion model (eq 4), and
the associated diusion time is observed to be 24.3 μs (see
gure S6 of the Supporting Information). However, the
uorescence autocorrelation of NPCE-tagged HSA cannot be
tted satisfactorily with eq 4. This implies that some additional
process is contributing to the uorescence uctuation apart
from the simple diusion. Upon incorporation of an additional
relaxation term (eq 5), the tting quality improves signicantly
(see Figure 3). The diusion time of NPCE-tagged HSA is
found to be much higher (140 μs) compared to the free NPCE
in the buer, as expected. The uorescence autocorrelation
curves for NPCE-tagged HSA were recorded in the presence of
dierent crowder concentrations, and all of them were tted
Figure 1. Variation of secondary structural parameters of HSA as a
function of (a) dextran-6, (b) dextran-40, and (c) dextran-70
concentrations. The lled circles represent the α-helicity. The open
circles represent the β-turn. The lled squares represent the random
coil. The open squares represent the β-sheet.
Figure 2. Variation of the average lifetime of NPCE-tagged HSA with
increasing concentrations of (a) dextran-6, (b) dextran-40, and (c)
dextran-70.
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6081
with eq 5. Some representative uorescence autocorrelation
curves and the comparison of tting by two dierent models
are shown in Figure 3. Such analysis gives two important
parameters about the system: the rst one is the diusion time,
and the other one is the time scale of the additional uctuation.
From the diusion time of NPCE-tagged HSA, the hydro-
dynamic radius of the native HSA is calculated to be 38.4 Å,
which is in good agreement with the previous reports.
34,64
It is
found that with an increasing dextran-6 concentration there
was almost no change of hydrodynamic radii of HSA. For
dextran-40, as we increase its concentration, the hydrodynamic
radius of HSA remains almost constant up to 75 g L1,
followed by a sharp decrease to 29.3 Å at 200 g L1. On the
otherhand,fordextran-70,agradualdecreaseinthe
hydrodynamic radius of HSA from 38.9 to 29.7 Å was
observed. The variation of hydrodynamics radii of HSA as a
function of crowder concentration is shown in Figure 4.
The additional uctuation time component observed for
native HSA is found to be 8.3 μs, which is assigned as the
conformational uctuation time of domain-III of HSA.
50
There
is almost no change of this time component with the addition
of dextran-6. For both dextran-40 and dextran-70, a gradual
increase of conformational uctuation time was observed, as
shown in Figure 5. For dextran-40, the value reaches 16.9 μsat
200 g L1, and for dextran-70, the value reaches 15.4 μs at 200
gL
1.
Activity Measurement. The activity of native HSA in the
absence of a crowder is found to be 3.60 ±0.15 (see gure S7
of the Supporting Information and Figure 6). With the increase
in crowder concentration, a gradual decrease of activity was
observed for all three cases. At the highest concentration of
crowders used in this study, the activities of HSA are found to
be 1.2, 1.27, and 1.05 units for dextran-6, dextran-40, and
dextran-70, respectively (see Figure 6).
Figure 3. Normalized autocorrelation curve for NPCE-tagged HSA
(represented by circles) in (a) dextran-6, (b) dextran-40, and (c)
dextran-70 at dierent crowder concentrations and the comparison of
tting of these uorescence autocorrelation curves with a single
diusion equation (in green) and after incorporation of a relaxation
time component (in black).
Figure 4. Variation of hydrodynamic radii of HSA with increasing
concentrations of (a) dextran-6, (b) dextran-40, and (c) dextran-70.
Solid red lines represent the best t using eq 14.
Figure 5. Variation of conformational uctuation time of HSA with
increasing concentrations of (a) dextran-6, (b) dextran-40, and (c)
dextran-70. Solid red lines represent the best t using eq 14.
Figure 6. Estimated activity of HSA from a 3 min experiment with
increasing concentrations of (a) dextran-6, (b) dextran-40, and (c)
dextran-70.
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6082
DISCUSSION
The uorescent labeling of NPCE to HSA leads to a 9 nm blue
shift in its emission maximum due to the more hydrophobic
environment inside the protein matrix, thus conrming the
tagging. With increasing crowder concentrations, the emission
maxima of NPCE in HSA does not show any appreciable
change. Also, for all three dierent crowders used in this study,
we observed a gradual decrease in the uorescence lifetime in
the experimental concentration range. However, the decrease is
very small for all cases (less than 5%). Taken together, these
two ensemble-averaged measurements hint that the local
environment around NPCE does not change signicantly in
the presence of a crowder.
The increased diusion time of NPCE-tagged HSA as
compared to the untagged NPCE also proves the tagging of
NPCE to HSA. The almost similar value of secondary
structural parameters for the tagged and untagged HSA
(table S1 of the Supporting Information) conrms a minimal
change of secondary structure of HSA upon tagging with
NPCE. The CD spectra of HSA in the presence of various
dextran molecules were recorded to have an idea on the change
in the overall secondary structure of HSA with increasing
crowder concentrations. We observed a huge decrease in the α-
helical content of HSA with increasing dextran-6 concen-
trations with a concomitant increase in the random coil
contribution. However, for dextran-40 and dextran-70, the
secondary structural parameters remain almost the same,
except for a slight increase in the α-helicity of HSA is observed
as the concentration of the crowder was increased. The general
expectation is that, with increasing crowder concentrations, the
exerted force on the protein surface would increase, making the
protein take a more compact structure. In fact, for larger
dextrans, this is the case observed. However, for the smallest
dextran, i.e., dextran-6, the observation is completely opposite,
suggesting the presence of some dierent type of interaction
between HSA and dextran-6. This interaction possibly
interferes with the stabilizing interactions that hold the
secondary structural content of HSA together.
FCS measurements reveal almost no change of hydro-
dynamic radii of HSA as the dextran-6 concentration is
increased from 0200 g L1, while, for dextran-40 and dextran-
70, there is an appreciable decrease in the hydrodynamic radius
with increasing crowder concentrations. The pressure on the
protein surface may modulate the protein structure, and we
have tried to make a rough estimation of the relative pressure
on the protein surface by various crowder molecules. To do
that, we have made some assumptions. First of all, we assumed
the ideal behavior of the solution and used the following
equation for calculating the pressure (P):
65
P
mnc
V
1
3
2
=(9)
where mand nare respectively the mass of a crowder
molecule and the total number of crowder molecules, cis the
root-mean-square speed of a crowder, and Vis the volume.
Second, we have assumed that there are no interactions
between the crowder molecules. The number of various
dextrans at a xed concentration will be dierent owing to
their dierent molecular weights. In this study, the highest
concentration that we have used is 200 g L1for all of the
crowders. For ensemble-averaged measurements, at this
concentration, the number of dextran-6, dextran-40, and
dextran-70 molecules is 6600, 1000, and 560, respectively,
for each HSA. For single molecular level experiments, this ratio
is 1:660 000, 1:100 000, and 1:56 000, respectively, for dextran-
6, dextran-40, and dextran-70. From these numbers, it is
evident that the number density of dextran-6 molecule is the
highest followed by dextran-40 and dextran-70, and the
number of crowder molecules is much higher as compared
to the protein molecule. Thus, we may visualize that a protein
is surrounded by a large number of dextran molecules. At this
point, we assume that only the rst layer of the crowder around
the protein will be eective to give a pressure on the protein
surface. Thus, in eq 9,nwill be the number of dextran
molecules that form the immediate rst layer around an HSA
molecule. Hydrodynamic radii of dextran-6, dextran-40, and
Scheme 2. Rough Estimation of Pressure on an HSA Molecule by the Crowders
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6083
dextran-70 (rc) are reported by Masuelli et al., and the values
are 18.6 Å, 47.8 Å, and 64.9 Å, respectively.
66
The
hydrodynamic radius of HSA (rHSA) is 38.9 Å. The radius
(rt) and the volume (vt) of the overall system comprising one
HSA surrounded by a single layer of crowder molecule are
dened by the following equation (see Scheme 2):
r
rr
tHSA
c
=+ (10)
v
r
4
3
tt
3
π=
(11)
The volume occupied by the crowder molecules in the rst
layer around HSA will be given by
Vv v
tHS
A
=− (12)
Also, the approximate number of crowder molecules forming
the rst layer around HSA (Nc) is therefore
N
V
v
c
c
=(13)
Using eq 13 in calculating the approximate numbers of
crowders in the rst layer around HSA suggests that we are
neglecting the presence of any void space. Using eq 9, the ratio
of the eective pressure on HSA by dierent dextrans are
estimated to be Pdextran40/Pdextran6/Pdextran70 = 1.00:0.38:0.27.
The exerted pressure is found to be largest for dextran-6
followed by dextran-40 and dextran-70, which suggests that
HSA should take a more compact structure in the presence of
dextran-6 compared to dextran-40 and dextran-70. However,
our FCS result shows a completely reversed picture. The
reason could be the smallest size of dextran-6 as discussed
above. The volume of one dextran-6 molecule is only 11% of
that of one HSA, whereas the volume of the other two
crowders is higher as compared to that of HSA. Owing to the
small volume, some of the dextran-6 molecule may go inside
the protein matrix. We know that dextran is a branched
polysaccharide chain, and this branching increases with an
increasing molecular weight. Lower molecular weight dextrans,
like dextran-6, are more like rod-shaped, and with an increasing
molecular weight, the dextran molecule tends to be more and
more spherical. This dierence of shape may also cause a
dextran-6 molecule to penetrate the protein matrix, while the
other two cannot. From our CD data, it is clear that, in the
presence of dextran-6, the secondary structure of HSA is
disrupted. Due to the breakdown of the secondary structure of
HSA, the polypeptide chain may unfold and the size may tend
to increase. This tendency is neutralized by the pressure
exerted on the surface of HSA by dextran-6, which tries to
decrease the overall size of the HSA molecule. These two
forces balance themselves in such a way that there is almost no
change in the size of the HSA when increasing the dextran-6
concentration from 0 to 200 g L1. For dextran-40 and
dextran-70, we observed a decrease of the hydrodynamic radius
of HSA, proposed solely due to the exerted pressure eect by
these crowders. Dextran-40 causes an approximate 25%
decrease in the hydrodynamic radius, and dextran-70 causes
a 23% decrease in the hydrodynamic radius at the highest
concentration of the crowder used. The relative change of the
hydrodynamic radii of HSA with three dierent dextrans is
plotted in Figure 7. A slightly higher eect of dextran-40 is
attributed to the higher pressure exerted by it on the HSA
surface as compared to dextran-70 as we have estimated.
To further illustrate the mechanism of crowder-induced
changes, we have done a companion experiment involving
thermally and chemically denatured HSA. Thermally dena-
tured HSA (at 75 °C) and chemically denatured HSA
(denatured by 7.5 M of urea) have α-helicities of 38% and
16.6% and hydrodynamic radii of 59.3 and 57.7 Å, respectively.
The relative change of the α-helicity and hydrodynamic radius
of such a way denatured HSA in the presence of three dierent
crowders are shown in Figure 8. In the presence of 100 g L1
dextran-6, the α-helicity of thermally and chemically denatured
HSA decreases 58% and 42.5% from its original value, whereas
for both dextran-40 and dextran-70, at the same concentration
of crowder, the α-helicity increases. For thermally denatured
HSA, this increase is minimal, whereas for chemically
denatured HSA, the increase is signicant. On the other
hand, all of the three crowders used in this study compact the
structure of the denatured HSA. For dextran-6, the decrease in
the hydrodynamic radius is only 10.5% for thermally denatured
HSA and 1.5% for chemically denatured HSA, whereas, for
dextran-40 and dextran-70, the changes are, respectively, 24.1%
Figure 7. Relative change of the hydrodynamic radius of HSA in the
presence of three dextrans with an increasing concentration.
Figure 8. Relative change of the α-helicity and hydrodynamic radius
of (a) thermally denatured HSA at 75 °C and (b) chemically
denatured HSA at 7.5 M urea, in the presence of 100 g L1dextran-6,
dextran-40, and dextran-70.
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6084
and 19% for thermally denatured HSA and 22.8% and 18% for
urea denatured HSA. The extent of change in α-helicity and
hydrodynamic radius in the presence of crowders for thermally
and chemically denatured HSA can be exploited to understand
how dextrans of dierent molecular weights interact with HSA.
Considering only the crowder exerted pressure, dextran-6
should impose a maximum decrease of the hydrodynamic
radius, but the result is just the opposite. This is the same trend
that we have observed for the native protein. Thus, it is
obvious that there is some other way in which dextran-6
interacts with HSA. Moreover, the % α-helicity of HSA, either
in the native or denatured state, decreases sharply with the
addition of dextran-6. These observations unambiguously
prove that dextran-6 induces a structural change in HSA not
only through the pressure eect but also through some soft
chemical interaction, which is destabilizing in nature. At this
point, we hypothesize that, owing to the smaller size of
dextran-6, it may access the protein interior that facilitates the
destabilizing interactions within HSA. This causes the
breakdown of weak chemical forces that stabilize HSA. For
dextran-40 and dextran-70, the changes are in line with what is
expected from the pressure eect only.
The FCS experiment reveals an additional time component
of 8.3 μs apart from the diusion time scale for the native HSA.
The origin of this extra time component arises because of the
local environmental change around the uorescent probe due
to the conformational uctuation of the protein.
51
Such
conformational uctuation consists of the breathing motion
and the motion of the side chains of the protein.
55
The time
required for this type of conformational uctuation is termed
as conformational uctuation time, which is typically in the
range 1100 μs.
6769
This uctuation is a very local
phenomenon and represents the ease of the concerted chain
motion dynamics. The conformational uctuation time in the
present case is 8.3 μs, which is in good agreement with
previous reports.
51
The eect of the crowder on this
uctuation time is more prominent for dextran-40 and
dextran-70 than for dextran-6. For dextran-6, there is almost
no change of the conformational uctuation time even at the
highest concentration. The protein interior can be accessed
through dextran-6, and the higher pressure exerted by this
probably nullied its eect on the conformational dynamics.
On the other hand, for dextran-40 and dextran-70, there is
almost a 2-fold increase in the conformational uctuation time,
which is probably because of the pressure eect by these
crowders on HSA. The relative change of the conformational
uctuation time of domain-III of HSA with an increasing
crowder concentration is shown in Figure 9. One interesting
point to note is that ensemble-averaged bulk measurement
cannot detect this local environmental change and thus
signies the importance of the single molecular level
measurement over the bulk measurement.
FCS data have been tted to estimate the thermodynamic
parameters related to crowder-induced structural changes of
HSA. Fitting of hydrodynamic radius data gives us an insight
into the thermodynamics of the overall structural change of
HSA, whereas the tting of the conformational uctuation time
sheds light on the thermodynamics of the crowder-induced
change of the domain-III of HSA. All transitions related to
structural and dynamical changes are tted with a two-state
model as
NCF
where Nstands for the native and represents the protein in
the absence of any crowder and Cstands for the crowder and
represents the protein in the presence of 200 g L1crowder.
The variation of RHand τRwith an increasing crowder
concentration was tted using the following two-state model:
70
YYe
1e
x
x
NC
=
+
(14)
In the above equation, Yis the spectroscopic signal at
some crowder concentration, YNand YCare the spectroscopic
signals for native and crowded states, respectively. Here, xis
dened as
70
x
Gm
RT
(crowder)
=Δ−[ ]
°
(15)
Here, ΔG0is the free energy change associated with the
concerned transition, mdenotes the slope of the free energy
change plotted against crowder concentration, and Rand Tare
the universal gas constant and temperature in K, respectively.
The midpoint of the transition was calculated using
71
G
m
C
1/2
[
]=
Δ
°
(16)
Both the hydrodynamic radius and conformational uctua-
tion time of HSA remain almost the same with the changing
dextran-6 concentration. Thus, in this case, Nand Cstates
are virtually the same, and therefore, the tting of dextran-6-
induced structural and dynamical change is meaningless. The
most important observation from this analysis is that in the
case of the overall structural change, dextran-40 shows a higher
value of ΔG0than that of dextran-70, whereas, for the
dynamical transition, which is a local phenomenon within the
domain-III of HSA, the order is reversed (see Table 1),
suggesting that, for a big multidomain protein, all domains and
the overall protein may not respond in a similar fashion.
From the activity measurements, we have observed that the
activity of HSA decreases with increasing crowder concen-
trations for all three crowders under investigation. There are
conicting reports on the origin of the catalytic activity of
HSA. Some reports suggest that Tyr-411 in domain-III is
mainly responsible for this esterase activity of HSA.
72,73
A
more recent article reports that the pseudo esterase activity of
HSA is a result of multiple irreversible chemical modications
rather than of the catalytic activity occurring at a single reactive
Figure 9. Relative change of conformational uctuation time of
domain-III of HSA for three dextrans with increasing concentrations.
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6085
site.
74
Whatever the case may be, the structural change of HSA
should have a profound eect on this activity. Additionally, the
conformational uctuation dynamics of domain-III could have
a major role to play.
75,76
It has also been reported that, during
the initial hydrolysis, only Tyr-411 is important.
74,77
In our
study, we have estimated the activity during the initial
hydrolysis period (up to 3 min) and have considered the
only involvement of Tyr-411. Naturally, we expect a direct
correlation between the esterase activity of HSA and
conformational uctuation time of domain-III of HSA. For
all three crowders used in this study, we observe a gradual
decrease of esterase activity of HSA with the increase in
crowder concentration. Here it is to be noted that the activity
study has been performed with untagged HSA where the Tyr-
411 residue is free. This decrease may be attributed to several
factors. First, with the increase in crowder concentration, the
viscosity of the solution increases. This makes the acetylation
of the Tyr-411 residue by PNPA more dicult. The second
factor that could be attributed to the retardation of the activity
of HSA is that with the increasing crowder concentration the
number of crowder molecules around a protein molecule
increases, making the accessibility of the Tyr-411 residue more
dicult. Third, for the two higher dextrans used in this study,
i.e., dextran-40 and dextran-70, with an increasing crowder
concentration, the size of the protein molecule decreases, and
because of this gradual compactness of HSA, it becomes
progressively harder for the PNPA molecule to reach the Tyr-
411 site. However, the most astonishing observation is the
decrease of the activity of HSA with an increasing dextran-6
concentration. The general expectation is exactly the opposite;
i.e., with a high dextran-6 concentration, HSA is denatured,
and as a consequence, Tyr-411 should be accessible. However,
our data suggests that it is indeed inaccessible even in this
denatured state. Note that a denatured state of a protein might
take such a conrmation that a particular amino acid sequence
does not need to be more exposed compared to that in its
native state.
7880
Thermal denaturation of HSA is one of such
example. In thermally denatured HSA, Tyr-411 is more
buried.
7880
In the present case, it is also possible that the
disruption of α-helicity does not lead to the increase in the
accessibility of the Tyr-411 residue.
For all of the cases, the extent of decrease is almost similar,
although the extent of size compactness and conformational
uctuation are not same. This may be due to the rst two
factors that we have stated as the possible reason for the
decrease of activity of HSA. In such a scenario, it is very hard
to nd any correlation between the conformational uctuation
time and the activity. In our next study, we will try to establish
a clear relation between the activity of a protein and its
structure and dynamics by eliminating the rst two factors
while measuring the activity.
CONCLUSION
The eect of macromolecular crowding on the structure,
function, and dynamics of HSA has been investigated by bulk
and single molecular level studies, which led us to conclude the
following. (i) For higher molecular weight dextrans, i.e., for
dextran-40 and dextran-70, with an increasing concentration of
the crowder, the overall size of the protein decreases with an
increase in the helical content of the protein (though slightly).
Thus, the secondary structural change and overall structural
change are in line with each other. However, for dextran-6, the
helical content of the protein decreases signicantly with no
change of the hydrodynamic radius. For dextran-40 and
dextran-70, the eect can solely be explained in terms of the
pressure exerted by the crowders. The disparity of the behavior
of dextran-6 compared to that of dextran-40 and dextran-70
conrms the presence of a destabilizing soft chemical
interaction between dextran-6 and HSA. We hypothesized
that the smaller size of dextran-6 allows it to go inside the
protein matrix and facilitates the destabilizing interaction. Due
to a ne balance between the denaturating eect and pressure
exerted by the dextran-6 onto the protein surface, there is
almost no change of the hydrodynamic radius with increasing
dextran-6 concentrations. (ii) To the best of our knowledge,
this is the very rst report on the modulation of conforma-
tional uctuation dynamics in a macromolecular crowded
environment. Smaller sized dextran-6 can not impose a
signicant constraint in the movement of the side chains of
HSA; on the other hand, relatively large crowders can
signicantly hinder the side-chain movement of HSA, thus
increasing the conformational uctuation time. The dierence
in the spatial distribution of these three dierent crowders
around HSA and their eect to modulate the structure and
dynamics of protein is summarized in Scheme 3. (iii) The
presence of crowders aect the activity of HSA, which
decreases with an increasing crowder concentration and is
probably due to the lower accessibility of the Tyr-411 residue
in the restricted environment imposed by macromolecular
crowding. As a whole, the global and local changes of HSA
Table 1. Thermodynamic Parameters for the Structural and
Dynamical Change of NPCE-Tagged HSA in Phosphate
Buer (pH 7.4, 50 mM) by Dextran-40 and Dextran-70
a
analysis of rHchange analysis of τRchange
ΔG0m[C]1/2 ΔG0m[C]1/2
dextran-40 3700 26 140 1400 16 90
dextran-70 970 12 80 2100 16 130
a
ΔG0is in cal mol1;mis in cal mol1g1L, and [C]1/2 is in g L1.
Scheme 3. Prole for the Hydrodynamic Radius and
Conformational Fluctuation Time of HSA in the Presence
of Various Crowders
a
a
Our hypothesis of the asymmetry in the spatial distribution of
dextran-6 around HSA has been emphasized.
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6086
induced by the macromolecular crowding have been
summarized, and the plausible explanations have been related
to the dierence in size of the crowders.
ASSOCIATED CONTENT
*
SSupporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.bio-
chem.8b00599.
Circular dichroism (CD) spectra of untagged and
NPCE-tagged HSA, CD spectra of HSA with increasing
concentrations of dextran-6, dextran-40, and dextran-70,
absorption and emission spectra of free NPCE and
NPCE-tagged HSA, emission spectra and uorescence
transient of NPCE-tagged HSA with varying concen-
trations of dextran-6, dextran-40, and dextran-70,
uorescence autocorrelation curve of free NPCE and
NPCE-tagged HSA in buer, activity of HSA with
increasing concentrations of dextran-6, dextran-40, and
dextran-70, CD signal of the thermally denatured HSA
at 75 °C in the absence and presence of dextran-6,
dextran-40, and dextran-70, and uorescence autocorre-
lation curve for the thermally denatured HSA at 75 °Cin
the absence of a crowder and in the presence of dextran-
6, dextran-40, and dextran-70 (PDF)
AUTHOR INFORMATION
Corresponding Author
*E-mail: psen@iitk.ac.in. Fax: +91-512-259-6806.
ORCID
Pratik Sen: 0000-0002-8202-1854
Funding
N.D. acknowledges the Council of Scientic and Industrial
Research (CSIR, Government of India) for providing a
fellowship. This work is nancially supported by the Science
and Engineering Research Board, Government of India (EMR/
2016/006555), and IIT Kanpur.
Notes
The authors declare no competing nancial interest.
REFERENCES
(1) Zhou, H., Rivas, G., and Minton, A. (2008) Macromolecular
Crowding and Confinement: Biochemical, Biophysical, and Potential
Physiological Consequences. Annu. Rev. Biophys. 37, 375397.
(2) Gnutt, D., and Ebbinghaus, S. (2016) The macromolecular
crowding effect from in vitro into the cell. Biol. Chem. 397,3744.
(3) Ralston, G. B. (1990) Effects of crowdingin protein solutions.
J. Chem. Educ. 67, 857860.
(4) Rivas, G., Ferrone, F., and Herzfeld, J. (2004) Life in a crowded
world. EMBO Rep. 5,2327.
(5) Wang, Y., Sarkar, M., Smith, A. E., Krois, A. S., and Pielak, G. J.
(2012) Macromolecular Crowding and Protein Stability. J. Am. Chem.
Soc. 134, 1661416618.
(6) Ellis, R. J. (2001) Macromolecular crowding: obvious but
underappreciated. Trends Biochem. Sci. 26, 597604.
(7) Zimmerman, S. B., and Trach, S. O. (1991) Estimation of
macromolecule concentrations and excluded volume effects for the
cytoplasm of Escherichia coli. J. Mol. Biol. 222, 599620.
(8) Zimmerman, S. B., and Minton, A. P. (1993) Macromolecular
Crowding: Biochemical, Biophysical, and Physiological Consequen-
ces. Annu. Rev. Biophys. Biomol. Struct. 22,2765.
(9) Sarkar, M., Li, C., and Pielak, G. (2013) Soft interactions and
crowding. Biophys. Rev. 5, 187194.
(10) Yu, I., Mori, T., Ando, T., Harada, R., Jung, J., Sugita, Y., and
Feig, M. (2016) Biomolecular interactions modulate macromolecular
structure and dynamics in atomistic model of a bacterial cytoplasm.
eLife 5, e19274.
(11) Sarkar, M., Smith, A., and Pielak, G. (2013) Impact of
reconstituted cytosol on protein stability. Proc. Natl. Acad. Sci. U. S. A.
110, 1934219347.
(12) Feig, M., and Sugita, Y. (2012) Variable Interactions between
Protein Crowders and Biomolecular Solutes Are Important in
Understanding Cellular Crowding. J. Phys. Chem. B 116 (1), 599
605.
(13) Puchkov, E. O. (2013) Intracellular viscosity: Methods of
measurement and role in metabolism. Biochem. (Moscow) Supp. Series
A: Membrane and Cell Biol. 7, 270279.
(14) Schreiber, G., Haran, G., and Zhou, H. X. (2009) Fundamental
Aspects of ProteinProtein Association Kinetics. Chem. Rev. 109,
839860.
(15) Zhou, H., and Dill, K. (2001) Stabilization of Proteins in
Confined Spaces. Biochemistry 40, 1128911293.
(16) Senske, M., Tork, L., Born, B., Havenith, M., Herrmann, C., and
Ebbinghaus, S. (2014) Protein Stabilization by Macromolecular
Crowding through Enthalpy Rather Than Entropy. J. Am. Chem. Soc.
136, 90369041.
(17) Malik, A., Kundu, J., Mukherjee, S., and Chowdhury, P. (2012)
Myoglobin Unfolding in Crowding and Confinement. J. Phys. Chem. B
116, 1289512904.
(18) Yuan, J., Chyan, C., Zhou, H., Chung, T., Peng, H., Ping, G.,
and Yang, G. (2008) The effects of macromolecular crowding on the
mechanical stability of protein molecules. Protein Sci. 17, 21562166.
(19) Politi, R., and Harries, D. (2010) Enthalpically driven peptide
stabilization by protective osmolytes. Chem. Commun. 46, 64496451.
(20) Zhou, H. (2004) Protein folding and binding in confined
spaces and in crowded solutions. J. Mol. Recognit. 17, 368375.
(21) Gnutt, D., Ahlers, J., Konig, B., Heyden, M., and Ebbinghaus, S.
(2018) SOD1 Folding Modulation in the Crowded Cell. Biophys. J.
114,5253.
(22) Zhou, H. (2013) Influence of crowded cellular environments
on protein folding, binding, and oligomerization: Biological
consequences and potentials of atomistic modeling. FEBS Lett. 587,
10531061.
(23) Gao, M., Estel, K., Seeliger, J., Friedrich, R., Dogan, S., Wanker,
E., Winter, R., and Ebbinghaus, S. (2015) Modulation of human IAPP
fibrillation: cosolutes, crowders and chaperones. Phys. Chem. Chem.
Phys. 17, 83388348.
(24) Harada, R., Sugita, Y., and Feig, M. (2012) Protein Crowding
Affects Hydration Structure and Dynamics. J. Am. Chem. Soc. 134,
48424849.
(25) Charlton, L., Barnes, C., Li, C., Orans, J., Young, G., and Pielak,
G. (2008) Residue-Level Interrogation of Macromolecular Crowding
Effects on Protein Stability. J. Am. Chem. Soc. 130, 68266830.
(26) Wang, P., Yu, I., Feig, M., and Sugita, Y. (2017) Influence of
protein crowder size on hydration structure and dynamics in
macromolecular crowding. Chem. Phys. Lett. 671,6370.
(27) Ando, T., Yu, I., Feig, M., and Sugita, Y. (2016)
Thermodynamics of Macromolecular Association in Heterogeneous
Crowding Environments: Theoretical and Simulation Studies with a
Simplified Model. J. Phys. Chem. B 120, 1185611865.
(28) Venturoli, D., and Rippe, B. (2005) Ficoll and dextran vs.
globular proteins as probes for testing glomerular permselectivity:
effects of molecular size, shape, charge, and deformability. Am. J.
Physiol-Renal Physiol. 288, F605F613.
(29) Shahid, S., Hassan, M. I., Islam, A., and Ahmad, F. (2017) Size-
dependent studies of macromolecular crowding on the thermody-
namic stability, structure and functional activity of proteins: in vitro
and in silico approaches. Biochim. Biophys. Acta, Gen. Subj. 1861, 178
197.
(30) He, X. M., and Carter, D. C. (1992) Atomic structure and
chemistry of human serum albumin. Nature 358, 209215.
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6087
(31) Larsen, M., Kuhlmann, M., Hvam, M., and Howard, K. (2016)
Albumin-based drug delivery: harnessing nature to cure disease. Mol.
Cell. Ther. 4,112.
(32) Dockal, M., Carter, D. C., and Ruker, F. (1999) The Three
Recombinant Domains of Human Serum Albumin: Structural
characterization and ligand binding properties. J. Biol. Chem. 274,
2930329310.
(33) Sudlow, G., Birkett, D. J., and Wade, D. N. (1975) The
Characterization of Two Specific Drug Binding Sites on Human
Serum Albumin. Mol. Pharmacol. 11, 824832.
(34) Ghuman, J., Zunszain, P., Petitpas, I., Bhattacharya, A., Otagiri,
M., and Curry, S. (2005) Structural Basis of the Drug-binding
Specificity of Human Serum Albumin. J. Mol. Biol. 353,3852.
(35) Abou-Zied, O., and Al-Lawatia, N. (2011) Exploring the Drug-
Binding Site Sudlow I of Human Serum Albumin: The Role of Water
and Trp214 in Molecular Recognition and Ligand Binding.
ChemPhysChem 12, 270274.
(36) Yamasaki, K., Chuang, V. T. G., Maruyama, T., and Otagiri, M.
(2013) Albumindrug interaction and its clinical implication.
Biochim. Biophys. Acta, Gen. Subj. 1830, 54355443.
(37) Abou-Zied, O., and Al-Shihi, O. (2008) Characterization of
Subdomain IIA Binding Site of Human Serum Albumin in its Native,
Unfolded, and Refolded States Using Small Molecular Probes. J. Am.
Chem. Soc. 130, 1079310801.
(38) Singh, P., and Chowdhury, P. K. (2013) Crowding-Induced
Quenching of Intrinsic Tryptophans of Serum Albumins: A Residue-
Level Investigation of Different Conformations. J. Phys. Chem. Lett. 4,
26102617.
(39) Biswas, S., and Chowdhury, P. K. (2015) Unusual domain
movement in a multidomain protein in the presence of macro-
molecular crowders. Phys. Chem. Chem. Phys. 17, 1982019833.
(40) Biswas, S., and Chowdhury, P. K. (2016) Correlated and
Anticorrelated Domain Movement of Human Serum Albumin: A
Peek into the Complexity of the Crowded Milieu. J. Phys. Chem. B
120, 48974911.
(41) Biswas, S., Mukherjee, S. K., and Chowdhury, P. K. (2016)
Crowder-Induced Rigidity in a Multidomain Protein: Insights from
Solvation. J. Phys. Chem. B 120, 1250112510.
(42) Samanta, N., Luong, T. Q., Das Mahanta, D., Mitra, R. K., and
Havenith, M. (2016) Effect of Short Chain Poly(ethyleneglycol)s on
the Hydration Structure and Dynamics around Human Serum
Albumin. Langmuir 32, 831837.
(43) Samanta, N., Mahanta, D. D., Hazra, S., Kumar, G. S., and
Mitra, R. K. (2014) Short Chain Polyethylene Glycols Unusually
Assist Thermal Unfolding of Human Serum Albumin. Biochimie 104,
8189.
(44) Biswas, S., Kundu, J., Mukherjee, S., and Chowdhury, P. (2018)
Mixed Macromolecular Crowding: A Protein and Solvent Perspective 3,
43164330.
(45) Kahsai, A. W., Rajagopal, S., Sun, J., and Xiao, K. (2014)
Monitoring protein conformational changes and dynamics using
stable-isotope labeling and mass spectrometry (CDSiL-MS). Nat.
Protoc. 9, 13011319.
(46) Guo, J., and Zhou, H.-X. (2016) Protein Allostery and
Conformational Dynamics. Chem. Rev. 116, 65036515.
(47) Gangupomu, V. K., Wagner, J. R., Park, I.-H., Jain, A., and
Vaidehi, N. (2013) Mapping Conformational Dynamics of Proteins
Using Torsional Dynamics Simulations. Biophys. J. 104, 19992008.
(48) Peters, T., Jr. (1985) Serum albumin. Adv. Protein Chem. 37,
161245.
(49) Peters, T., Jr. (1996) All about albumin: Biochemistry, Genetics,
and Medical Applications, Academic press, San Diego.
(50) Sengupta, B., Das, N., and Sen, P. (2016) Elucidation of μs
dynamics of domain-III of human serumalbumin duringthe chemical
and thermal unfolding: fluorescence correlationspectroscopic inves-
tigation. Biophys. Chem. 221,1725.
(51) Sengupta, B., Acharya, A., and Sen, P. (2016) Elucidation of the
local dynamics of domain-III of human serum albumin over the psμs
time regime using a new fluorescent label. Phys. Chem. Chem. Phys. 18,
1435014358.
(52) Yadav, R., Sengupta, B., and Sen, P. (2014) Conformational
Fluctuation Dynamics of Domain I of Human Serum Albumin in the
Course of Chemically and Thermally Induced Unfolding Using
Fluorescence Correlation Spectroscopy. J. Phys. Chem. B 118, 5428
5438.
(53) Aujard, I., Benbrahim, C., Gouget, M., Ruel, O., Baudin, J. B.,
Neveu, P., and Jullien, L. (2006) o-Nitrobenzyl Photolabile Protecting
Groups with Red-Shifted Absorption: Syntheses and Uncaging Cross-
Sections for One- and Two-Photon Excitation. Chem. - Eur. J. 12,
68656879.
(54) Wang, R., Sun, S., Bekos, E., and Bright, F. V. (1995) Dynamics
surrounding chemically denatured, serum albumin Cys-34 in native,
and silicam-Adsorbed bovine serum albumin. Anal. Chem. 67, 149
159.
(55) Lakowicz, J. R. (2006) Principles of uorescence spectroscopy, 3rd
ed, Springer, New York, NY.
(56) Edwards, J. S., Kumbhar, A., Roberts, A., Hemmert, A. C.,
Edwards, C. C., Potter, P. M., and Redinbo, M. R. (2011)
Immobilization of Active Human Carboxylesterase 1 in Biomimetic
Silica Nanoparticles. Biotechnol. Prog. 27, 863869.
(57) Darkoh, C., Brown, E. L., Kaplan, H. B., and Dupont, H. L.
(2013) Bile Salt Inhibition of Host Cell Damage by Clostridium
Difficile Toxins. PLoS One 8, e79631.
(58) Sett, R., Ganguly, A., and Guchhait, N. (2016) Effect of the
Binding Interaction of an Emissive Niacin Derivative on the
Conformation and Activity of a Model Plasma Protein: A
Spectroscopic and Simulation-Based Approach. J. Photochem. Photo-
biol., B 164, 141150.
(59) Sengupta, B., Chaudhury, A., Das, N., and Sen, P. (2017) Single
Molecular Level Investigation of Structure and Dynamics of Papain
under Denaturation. Protein Pept. Lett. 24, 10731081.
(60) Sengupta, B., Das, N., and Sen, P. (2018) Monomerization and
aggregation of β-lactoglobulin under adverse condition: A fluores-
cence correlation spectroscopic investigation. Biochim. Biophys. Acta,
Proteins Proteomics 1866, 316326.
(61) Elson, E. L. (2013) Chapter two brief introduction to
fluorescence correlation spectroscopy. Methods Enzymol. 518,1141.
(62) Muller, C. B., Loman, A., Pacheco, V., Koberling, F., Willbold,
D., Richtering, W., and Enderlein, J. (2008) Precise measurement of
diffusion by multi-color dual-focus fluorescence correlation spectros-
copy. Europhys. Lett. 83, 46001.
(63) Sherman, E., Itkin, A., Kuttner, Y. Y., Rhoades, E., Amir, D.,
Haas, E., and Haran, G. (2008) Using Fluorescence Correlation
Spectroscopy to Study Conformational Changes in Denatured
Proteins. Biophys. J. 94, 48194827.
(64)Chowdhury,R.,Chattoraj,S.,Mojumdar,S.S.,and
Bhattacharyya, K. (2013) FRET between a Donor and an Acceptor
Covalently Bound to Human Serum Albumin in Native and Non-
Native States. Phys. Chem. Chem. Phys. 15, 1628616293.
(65) Atkins, P., and Paula, J. D. (2009) Physical Chemistry, 9th ed.,
Oxford University Press.
(66) Masuelli, M. A. (2013) Dextrans in aqueous solution.
Experimental Review on intrinsic viscosity measurements and
temperature effect. J. Polymer Biopolymer Phys. Chem. 1,1321.
(67) Das, G., Chattoraj, S., Nandi, S., Mondal, P., Saha, A.,
Bhattacharyya, K., and Ghosh, S. (2018) Probing the conformational
dynamics of photosystem I in unconfined and confined spaces. Phys.
Chem. Chem. Phys. 20, 449455.
(68) Sasmal, D. K., Mondal, T., Mojumdar, S. S., Choudhury, A.,
Banerjee, R., and Bhattacharyya, K. (2011) An FCS Study of
Unfolding and Refolding of CPM-Labeled Human Serum Albumin:
Role of Ionic Liquid. J. Phys. Chem. B 115, 1307513083.
(69) Pabbathi, A., Patra, S., and Samanta, A. (2013) Structural
Transformation of Bovine Serum Albumin Induced by Dimethyl
Sulfoxide and Probed by Fluorescence Correlation Spectroscopy and
Additional Methods. ChemPhysChem 14, 24412449.
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6088
(70) Naidum, K. T., and Prabhu, P. N. (2011) Proteinsurfactant
interaction: sodium dodecyl sulfate induced unfolding of ribonuclease
A. J. Phys. Chem. B 115, 1476014767.
(71) Al-Soufi, W., Reija, B., Felekyan, S., Seidel, C., and Novo, M.
(2008) Dynamics of Supramolecular Association Monitored by
Fluorescence Correlation Spectroscopy. ChemPhysChem 9, 1819
1827.
(72) Kurono, Y., Kushida, I., Tanaka, H., and Ikeda, K. (1992)
Esterase-like activity of human serum-albumin. VIII. Reaction with
amino-acid para-nitrophenyl esters. Chem. Pharm. Bull. 40, 2169
2172.
(73) Watanabe, H., Tanase, S., Nakajou, K., Maruyama, T., Kragh-
Hansen, U., and Otagiri, M. (2000) Role of Arg-410 andTyr-411 in
human serum albumin for ligand binding and esterase-like activity.
Biochem. J. 349, 813819.
(74) Ascenzi, P., Gioia, M., Fanali, G., Coletta, M., and Fasano, M.
(2012) Pseudo-enzymatic hydrolysis of 4-nitrophenyl acetate by
human serum albumin: pH-dependence of rates of individual steps.
Biochem. Biophys. Res. Commun. 424, 451455.
(75) Yon, J., Perahia, D., and Ghelis, C. (1998) Conformational
dynamics and enzyme activity. Biochimie 80,3342.
(76) Gagne, D., French, R., Narayanan, C., Simonovic, M., Agarwal,
P., and Doucet, N. (2015) Perturbation of the Conformational
Dynamics of an Active-Site Loop Alters Enzyme Activity. Structure 23,
22562266.
(77) Goncharov, N., Belinskaia, D., Shmurak, V., Terpilowski, M.,
Jenkins, R., and Avdonin, P. (2017) Serum Albumin Binding and
Esterase Activity: Mechanistic Interactions with Organophosphates.
Molecules 22, 12011227.
(78) Flora, K., Brennan, J., Baker, G., Doody, M., and Bright, F. V.
(1998) Unfolding of Acrylodan-Labeled Human Serum Albumin
Probed by Steady-State and Time-Resolved Fluorescence Methods.
Biophys. J. 75, 10841096.
(79) Pico, G. (1997) Thermodynamic features of the thermal
unfolding of human serum albumin. Int. J. Biol. Macromol. 20,6373.
(80) Shaw, A., and Pal, S. (2008) Spectroscopic studies on the effect
of temperature on pH-induced folded states of human serum albumin.
J. Photochem. Photobiol., B 90,6977.
Biochemistry Article
DOI: 10.1021/acs.biochem.8b00599
Biochemistry 2018, 57, 60786089
6089
... The labeling of HSA with NPCE can be found in prevous publications 73,74 and in section S8 of the SM. ...
... We performed the fluorescence correlation spectroscopic (FCS) measurements on an instrument built in our laboratory, and the details can be found in our previous publications. 74,75,76,78 FCS is a unique single molecular level technique, where fluorescence intensity fluctuation from a tiny observation volume is analysed to extract the reason for such a fluctuation. 15 The major factor for such fluctuation is translational diffusion. ...
Article
Many current methods for detecting molecular-level heterogeneity are complex and require stringent data analysis, limiting their widespread use. Herein, we propose a novel edge effect, i.e., the shift of excitation spectra at the blue edge of emission (which we termed as Blue Edge Emission Shift, BEEmS), to perceive the structural heterogeneity. This method is simple and can be easily implemented with commonly available fluorimeter. Red edge excitation shift (REES), a related technique, is already in use, but like most other known techniques, its usefulness is constrained by its dependence on environmental rigidity. Our method does not suffer from this drawback significantly. We showed the generality of the proposed method taking various chemically and biologically heterogenous systems including molecular liquid, deep eutectic solvents, organic cavitand, micelle and protein. BEEmS certainly comes out as a more effective sensor of heterogeneity than REES in certain cases (like denatured protein, hydrophobic deep eutectic solvent and SDS micelle) where solvation time is not sufficiently slow to be detected by REES, but can be measured though BEEmS. Furthermore, unlike most existing techniques, domain-specific heterogeneity of a model multi-domain protein is successfully measured.
... 27 The local and global conformations, and concentrations of biomolecules, like nucleic acids and proteins, can be vividly modulated in a crowded environment created by an inert crowder, like PEG8000. [28][29][30] Very recently, it has been found that different ordered and disordered proteins can undergo LLPS in the presence of a crowder, like PEG. 21 The formation of the phase-separated liquid droplets occurred through non-covalent interactions between the proteins, and such droplets were characterized using various spectroscopic and microscopic techniques. ...
Article
Full-text available
Owing to the significant role in the subcellular organization of biomolecules, physiology, and the realm of biomimetic materials, studies related to biomolecular condensates formed through liquid–liquid phase separation (LLPS) have emerged as a growing area of research. Despite valuable contributions of prior research, there is untapped potential in exploring the influence of phase separation on the conformational dynamics and enzymatic activities of native proteins. Herein, we investigate the LLPS of β-lactoglobulin (β-LG), a non-intrinsically disordered protein, under crowded conditions. In-depth characterization through spectroscopic and microscopic techniques revealed the formation of dynamic liquid-like droplets, distinct from protein aggregates, driven by hydrophobic interactions. Our analyses revealed that phase separation can alter structural flexibility and photophysical properties. Importantly, the phase-separated β-LG exhibited efficient enzymatic activity as an esterase; a characteristic seemingly exclusive to β-LG droplets. The droplets acted as robust catalytic crucibles, providing an ideal environment for efficient ester hydrolysis. Further investigation into the catalytic mechanism suggested the involvement of specific amino acid residues, rather than general acid or base catalysis. Also, the alteration in conformational distribution caused by phase separation unveils the latent functionality. Our study delineates the understanding of protein phase separation and insights into the diverse catalytic strategies employed by proteins. It opens exciting possibilities for designing functional artificial compartments based on phase-separated biomolecules.
... In fact, there are several studies, including that from our group, showing that the solvation time around the protein decreases upon its denaturation. 80,84,[86][87][88] Therefore, our result follows the expected trend. However, as exchange is the primary channel between associated and bulk water, any structural modulation will have a prominent effect not only on protein stability, but on water structure also. ...
Article
Full-text available
Modulation of protein associated water might decide protein–osmolyte interaction, where the rigidity and flexibility of associated water induce stabilization and destabilization, respectively.
... We performed the fluorescence correlation spectroscopic (FCS) measurements on an instrument built in our laboratory. The details can be found in our previous publications [47,[55][56][57] and section S2 of the Supplementary Material. For a single component system, assuming Gaussian detection volume, fluorescence intensity autocorrelation function (ACF) can be written as [51,58] (6) ...
Article
Deep eutectic solvents (DESs) are new-generation solvents with exquisite and tuneable properties. Molecular-level heterogeneity has been identified as an intriguing feature of such solvents. Herein, we examined the spatio-temporal heterogeneity of a potential non-ionic biocatalytic DES, acetamide/urea/sorbitol (0.5Ac/0.3Ur/0.2Sor), and compared the result with corresponding binary acetamide/urea (0.6Ac/0.4Ur) DES, and another related non-ionic ternary DES (0.55Ac/0.36Ur/0.09PEG). The effect of the addition of a third component on the spatio-temporal heterogeneity of a DES was investigated. The excitation wavelength-dependent emission measurement suggests an induction of spatial heterogeneity in acetamide/urea/sorbitol compared to spatially homogenous acetamide/urea and acetamide/urea/PEG. The dynamic heterogeneity measurements in terms of solvation dynamics, dielectric relaxation, and rotational/translational diffusion indicate a length and timescale dependency. Overall, acetamide/urea/sorbitol is found to be dynamically more heterogenous than the other two related DESs.
... The real protein environment in living systems significantly influences their conformational, functional, and dynamic properties [1,2]. The in vivo water compartments are saturated by a large number of macromolecules with concentrations from 80 to 400 g/L, occupying up to 40% of the water volume and creating macromolecular crowding [3,4]. Molecular crowding leads to an increase in excluded volume effects, a rise of viscosity, and the growth of specific and non-specific intermolecular interactions [5][6][7][8][9][10][11]. Significant changes in protein structure, diffusion transfer, and functioning were revealed under crowding conditions [12][13][14][15]. ...
Article
Full-text available
Intracellular environment includes proteins, sugars, and nucleic acids interacting in restricted media. In the cytoplasm, the excluded volume effect takes up to 40% of the volume available for occupation by macromolecules. In this work, we tested several approaches modeling crowded solutions for protein diffusion. We experimentally showed how the protein diffusion deviates from conventional Brownian motion in artificial conditions modeling the alteration of medium viscosity and rigid spatial obstacles. The studied tracer proteins were globular bovine serum albumin and intrinsically disordered α-casein. Using the pulsed field gradient NMR, we investigated the translational diffusion of protein probes of different structures in homogeneous (glycerol) and heterogeneous (PEG 300/PEG 6000/PEG 40,000) solutions as a function of crowder concentration. Our results showed fundamentally different effects of homogeneous and heterogeneous crowded environments on protein self-diffusion. In addition, the applied "tracer on lattice" model showed that smaller crowding obstacles (PEG 300 and PEG 6000) create a dense net of restrictions noticeably hindering diffusing protein probes, whereas the large-sized PEG 40,000 creates a "less restricted" environment for the diffusive motion of protein molecules.
Article
Deep eutectic solvents (DESs) are potential biocatalytic media due to their easy preparation, fine-tuneability, biocompatibility, and most importantly, due to their ability to keep protein stable and active. However, there are many unanswered questions and gaps in our knowledge about how proteins behave in these alternate media. Herein, we investigated solvation dynamics, conformational fluctuation dynamics, and stability of human serum albumin (HSA) in 0.5 Acetamide/0.3 Urea/0.2 Sorbitol (0.5Ac/0.3Ur/0.2Sor) DES of varying concentrations to understand the intricacy of protein behaviour in DES. Our result revealed a gradual decrease in the side-chain flexibility and thermal stability of HSA beyond 30 % DES. On the other hand, the associated water dynamics around domain-I of HSA decelerate only marginally with increasing DES content, although viscosity rises considerably. We propose that even though macroscopic solvent properties are altered, a protein feels only an aqueous type of environment in the presence of DES. This is probably the first experimental study to delineate the role of the associated water structure of the enzyme for maintaining its stability inside DES. Although considerable effort is necessary to generalize such claims, it might serve as the basis for understanding why proteins remain stable and active in DES.
Article
Enzyme function is governed by a complex network of conformational changes and internal dynamics, with the same getting more convoluted in the crowded cellular environment. Here, we have explored an intricate interplay amongst activity, structure, conformation, and dynamics of a multidomain enzyme, AK3L1 (UniProtKB: Q9UIJ7) in the crowded milieu. We have monitored changes in the enzyme landscape in response to the chemical denaturant, urea, under the influence of different concentrations of macromolecular crowders. Extensive experimental analyses using FRET-based domain displacement measurements, sub-nanosecond time scale local dynamics, and global structural changes, along with enzymatic activity studies, has been carried out to get deeper insights into the factors that may modulate the functional landscape of adenylate kinase (AK3L1). It was observed that AK3L1 gets activated at low urea concentrations, whereas higher urea concentrations unfold and thereby deactivate the enzyme. A sequential response of AK3L1 is observed towards external perturbation (urea) occurring through a series of well-defined steps. Incorporation of crowders not only shift the maximum activity of enzyme to a higher urea concentration, but also enhance domain compaction, as revealed by FRET studies. The modulation in enzyme activity and solvation dynamics acting as local response, precede global unfolding of the enzyme, indicating that the structural alterations around the active site are quite decoupled from the large amplitude global transitions.
Article
Full-text available
In the living cell, biomolecules perform their respective functions in the presence of not only one type of macromolecules but rather in the presence of various macromolecules with different shapes and sizes. In this study, we have investigated the effects of five single macromolecular crowding agents, Dextran 6, Dextran 40, Dextran 70, Ficoll 70, and PEG 8000 and their binary mixtures on the modulation in the domain separation of human serum albumin using a Förster resonance energy transfer-based approach and the translational mobility of a small fluorescent probe fluorescein isothiocyanate (FITC) using fluorescence correlation spectroscopy (FCS). Our observations suggest that mixed crowding induces greater cooperativity in the domain movement as compared to the components of the mixtures. Thermodynamic analyses of the same provide evidence of crossovers from enthalpy-based interactions to effects dominated by hard-sphere potential. When compared with those obtained for individual crowders, both domain movements and FITC diffusion studies show significant deviations from ideality, with an ideal solution being considered to be that arising from the sum of the contributions of those obtained in the presence of individual crowding agents. Considering the fact that domain movements are local (on the order of a few angstroms) in nature while translational movements span much larger lengthscales, our results imply that the observed deviation from simple additivity exists at several possible levels or lengthscales in such mixtures. Moreover, the nature and the type of deviation not only depend on the identities of the components of the crowder mixtures but are also influenced by the particular face of the serum protein (either the domain I–II or the domain II–III face) that the crowders interact with, thus providing further insights into the possible existence of microheterogeneities in such solutions.
Article
Full-text available
ELife digest Much of the work that has been done to understand how cells work has involved studying parts of a cell in isolation. This is particularly true of studies that have examined the arrangement of atoms in large molecules with elaborate structures like proteins or DNA. However, cells are densely packed with many different molecules and there is little proof that proteins keep the same structures inside cells that they have when they are studied alone. To really understand how cells work, new ways to understand how molecules behave inside cells are needed. While this cannot be achieved directly, technology has now reached the stage where we can, to some extent, study living cells by recreating them virtually. Simulated cells can copy the atomic details of all the molecules in a cell and can estimate how different molecules might behave together. Yu et al. have now developed a computer simulation of part of a cell from the bacterium, Mycoplasma genitalium, one of the simplest forms of life on Earth. This model suggested new possible interactions between molecules inside cells that cannot currently be studied in real cells. The model shows that some proteins have a much less rigid structure in cells than they do in isolation, whilst others are able to work together more closely to carry out certain tasks. Finally, the model predicted that small molecules such as food, water and drugs would move more slowly through cells as they become stuck or trapped by larger molecules. These results could be particularly important in helping to improve drug design. Currently the simulations are limited, and can only model parts of simple cells for less than a thousandth of a second. However, in future it should be possible to recreate larger and more complex cells, including human cells, for longer periods of time. These could be used to better study human diseases and help to design new treatments. The ultimate goal is to simulate a whole cell in full detail by combining all the available experimental data. DOI: http://dx.doi.org/10.7554/eLife.19274.002
Article
Full-text available
Introduction: Papain is a cysteine protease enzyme present in papaya and known to help in digesting peptide. Thus the structure and function of the active site of papain is of interest. Objective: The objective of present study is to unveil the overall structural transformation and the local structural change around the active site of papain as a function of chemical denaturant. Methods: Papain has been tagged at Cys-25 with a thiol specific fluorescence probe N-(7-dimethylamino-4-methylcoumarin-3-yl) iodoacetamide (DACIA). Guanidine hydrochloride (GnHCl) has been used as the chemical denaturant. Steady state, time-resolved, and single molecular level fluorescence techniques was applied to map the change in the local environment. Results: It is found that papain undergoes a two-step denaturation in the presence of GnHCl. Fluorescence correlation spectroscopic (FCS) data indicate that the size (hydrodynamic diameter) of native papain is ~36.8 Å, which steadily increases to ~53 Å in the presence of 6M GnHCl. FCS study also revealed that the conformational fluctuation time of papain is 6.3 μs in its native state, which decreased to 2.7 μs in the presence of 0.75 M GnHCl. Upon further increase in GnHCl concentration the conformational fluctuation time increase monotonically till 6 M GnHCl, where the time constant is measured as 14 μs. On the other hand, the measurement of ellipticity, hence the helical structure, by circular dichroism spectroscopy is found to be incapable to capture such structural transformation. Conclusion: It is concluded that in the presence of small amount of GnHCl the active site of papain takes up a more compact structure (although the overall size increases) than in the native state, which has been designated as the intermediate state.
Article
Full-text available
The albumin molecule, in contrast to many other plasma proteins, is not covered with a carbohydrate moiety and can bind and transport various molecules of endogenous and exogenous origin. The enzymatic activity of albumin, the existence of which many scientists perceive skeptically, is much less studied. In toxicology, understanding the mechanistic interactions of organophosphates with albumin is a special problem, and its solution could help in the development of new types of antidotes. In the present work, the history of the issue is briefly examined, then our in silico data on the interaction of human serum albumin with soman, as well as comparative in silico data of human and bovine serum albumin activities in relation to paraoxon, are presented. Information is given on the substrate specificity of albumin and we consider the possibility of its affiliation to certain classes in the nomenclature of enzymes.
Article
Fluorescence dynamics of Photosystem I (PSI) in bulk water as well as inside a confined environment like liposome have been investigated using time resolved confocal microscopy. In bulk water, PSI exhibits a major emission peak at ~680 nm while in the liposome it exhibits markedly blue shifted emission maxima at ~485 nm. This is indicative of conformational changes due to entrapment and emergence of a stressed conformation of PSI inside liposome. The observed time constants for the fluorescence lifetime of PSI inside the liposome are significantly high as opposed to the PSI in bulk water. More interestingly, fluorescence intensity of PSI in bulk water exhibits strong fluctuations with many high intensity jumps and these are anti-correlated with fluorescence lifetime of PSI. In contrast, inside the liposome, no such anti-correlated behaviour is observed. We further demonstrated that PSI exhibits at least two conformational states in bulk water whereas a single conformation is observed inside the liposome indicating conformational rigidity and locking of PSI complex inside a liposome.
Article
β-lactoglobulin is one of the major components of bovine milk and it remains in a dimeric form under physiological conditions. The present contribution elucidates the structural change of β-lactoglobulin at pH7.4 under the action of guanidine hydrochloride (GnHCl) and heat at the single molecular level. The only free cysteine (Cys-121) of β-lactoglobulin has been tagged with 7-Diethylamino-3-(4-maleimidophenyl)-4-methylcoumarin (CPM) for this purpose. The dimeric structure of β-lactoglobulin found to undergoes a monomerization prior to the unfolding process upon being subjected to GnHCl. The hydrodynamic diameter of the native dimer, native monomer and the unfolded monomer has been estimated as ~55Å, ~29Å and ~37Å, respectively. The free energy change for the monomerization and denaturation are respectively 1.57kcal mol(-1) and 8.93kcal mol(-1). With change in temperature, development of two types of aggregates (small aggregates and large aggregates) was observed, which is triggered by the formation of the monomeric structure of β-lactoglobulin. The hydrodynamic diameters of the smaller and larger aggregates has been estimated to be ~77Å and ~117Å, respectively. The formation of small aggregates turns out to be reversible whereas that of larger aggregates is irreversible. The free energy associated with these two steps are 0.69kcal mol(-1) and 9.09kcal mol(-1). Based on the size parameters, the smaller and larger aggregates have been proposed to contain three and four monomeric units. It has also been concluded that the monomeric subunits retain their native like secondary structure in these aggregates.
Article
We investigate the effects of protein crowder sizes on hydration structure and dynamics in macromolecular crowded systems by all-atom MD simulations. The crowded systems consisting of only small proteins showed larger total surface areas than those of large proteins at the same volume fractions. As a result, more water molecules were trapped within the hydration shells, slowing down water diffusion. The simulation results suggest that the protein crowder size is another factor to determine the effect of macromolecular crowding and to explain the experimental kinetic data of proteins and DNAs in the presence of crowding agents.
Article
The local structural dynamics and denaturation profile of domain-III of HSA against guanidine hydrochloride (GnHCl) and temperature has been studied using a coumarin based solvatochromic fluorescent probe p-nitrophenyl coumarin ester (NPCE), covalently tagged to Tyr-411 residue. By the steady state, time-resolved and single molecular level fluorescence studies it has been established that the domain-III of HSA is very sensitive to GnHCl but somewhat resistant to temperature and the domain specific unfolding proceeds in an altered way as compared to the overall unfolding of HSA. While the overall denaturation of HSA is a two-state process for both GnHCl and heat, domain-III adopts two intermediate states for GnHCl induced denaturation and one intermediate state for temperature induced denaturation. Fluorescence correlation spectroscopic investigation divulges the conformational dynamics of domain-III of HSA in the native, intermediates and denatured state.