ArticlePDF Available

An Upper Limit for Macromolecular Crowding Effects

Authors:

Abstract and Figures

Solutions containing high macromolecule concentrations are predicted to affect a number of protein properties compared to those properties in dilute solution. In cells, these macromolecular crowders have a large range of sizes and can occupy 30% or more of the available volume. We chose to study the stability and ps-ns internal dynamics of a globular protein whose radius is ~2 nm when crowded by a synthetic microgel composed of poly(N-isopropylacrylamide-co-acrylic acid) with particle radii of ~300 nm. Our studies revealed no change in protein rotational or ps-ns backbone dynamics and only mild (~0.5 kcal/mol at 37°C, pH 5.4) stabilization at a volume occupancy of 70%, which approaches the occupancy of closely packing spheres. The lack of change in rotational dynamics indicates the absence of strong crowder-protein interactions. Our observations are explained by the large size discrepancy between the protein and crowders and by the internal structure of the microgels, which provide interstitial spaces and internal pores where the protein can exist in a dilute solution-like environment. In summary, microgels that interact weakly with proteins do not strongly influence protein dynamics or stability because these large microgels constitute an upper size limit on crowding effects.
Content may be subject to copyright.
RESEARC H ARTIC L E Open Access
An upper limit for macromolecular crowding
effects
Andrew C Miklos
1
, Conggang Li
1,2
, Courtney D Sorrell
3,4,5
, L Andrew Lyon
3,4
and Gary J Pielak
1,6,7*
Abstract
Background: Solutions containing high macromolecule concentrations are predicted to affect a number of protein
properties compared to those properties in dilute solution. In cells, these macromolecular crowders have a large
range of sizes and can occupy 30% or more of the available volume. We chose to study the stability and ps-ns
internal dynamics of a globular protein whose radius is ~2 nm when crowded by a synthetic microgel composed
of poly(N-isopropylacrylamide-co-acrylic acid) with particle radii of ~300 nm.
Results: Our studies revealed no change in protein rotational or ps-ns backbone dynamics and only mild
(~0.5 kcal/mol at 37°C, pH 5.4) stabilization at a volume occupancy of 70%, which approaches the occupancy of
closely packing spheres. The lack of change in rotational dynamics indicates the absence of strong crowder-protein
interactions.
Conclusions: Our observations are explained by the large size discrepancy between the protein and crowders and
by the internal structure of the microgels, which provide interstitial spaces and internal pores where the protein
can exist in a dilute solution-like environment. In summary, microgels that interact weakly with proteins do not
strongly influence protein dynamics or stability because these large microgels constitute an upper size limit on
crowding effects.
Background
The cellular interior, where most biological processes
occur, is unlike the dilute solutions where most proteins
are studied. The large volume excluded by high macro-
molecule concentrations in cells, from 20-40% [1], is
predicted to change many protein properties compared
to dilute solution. We used a synthetic microgel com-
posed of poly(N-isopropylacrylamide-co-acrylic acid)
[p-NIPAm-co-AAc (Figure 1A)], as a crowding agent to
study the backbone dynamics and the stability of the
globular test protein, chymotrypsin inhibitor 2 (CI2).
p-NIPAm-co-AAc is of interest in pharmaceutical
applications because it forms environmentally sensitive
microgels [2]. Each microgel particle (Figure 1B) is a
lightly cross-linked single polymer molecule of molecu-
lar weight 10
9
Da with an average of 70 monomer units
between each cross link. The polymer absorbs a large
amount of water resulting in spherical particles of
300 nm radii that exclude large amounts of solution
volume. Their porosity arises from the balance between
the external (solution) osmotic pressure and the internal
osmotic pressure. This internal pressure is the result of
the solvated cations that neutralize the deprotonated
polymer side chains. We chose this crowding agent
because its status as a drug delivery molecule makes it
pharmaceutically relevant, and its ability to take up
water provides a model for volume exclusion by a mole-
cule much larger than our test protein.
CI2 is a small globular protein (7.4 kDa, PDB ID:
2CI2) that exhibits two-state folding [3]. NMR relaxa-
tion experiments [4] allowed us to assess backbone rota-
tional dynamics for CI2 in the presence and absence of
p-NIPAm-co-AAc. Amide proton exchange experiments
[5,6] allowed us to assess the stability of CI2 in dilute
and crowded conditions.
Globular proteins are often treated like hard spheres,
buttheyhavemeasurableamounts of internal motion.
Analysis of relaxation parameters from NMR experi-
ments - longitudinal and transverse relaxation times, T
1
and T
2
,andthe
15
N-
1
H nuclear Overhauser effect
* Correspondence: gary_pielak@unc.edu
1
Department of Chemistry, University of North Carolina, Chapel Hill, North
Carolina 27599, USA
Full list of author information is available at the end of the article
Miklos et al.BMC Biophysics 2011, 4:13
http://www.biomedcentral.com/2046-1682/4/13
© 2011 Miklos et al; licensee BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons
Attribution License (http://creative commons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in
any medium, pro vided the original work is properly cited.
(NOE) of backbone
15
N atoms - offers a residue-level
window into this ps-ns backbone motion. The analysis
involves a model-free method established by Lipari and
Szabo [7]. Analysis is performed by fitting the spectral
density function I(ω) as calculated from measured T
1
,
T
2
, and NOE values [8], to the equation [7]
I(ω)= 2
5S2τm
1+ω2τ2
m
+(1 S2)τ
1+ω2τ2
The overall correlation time τis linked to the correla-
tion time for isotropic tumbling, τ
m
, and internal motion
timescale, τ
e
, by the equation
1
τ=1
τ
e
+1
τ
m
with the internal motions faster than the overall iso-
tropic tumbling. The order parameter, S
2
,canhave
values between 0 and 1, and is related to the degree of
internal mobility for a particular
1
H-
15
Nvector.AnS
2
value of 0 corresponds to complete freedom of motion.
In this instance, relaxation is related solely to internal
motion. An S
2
value of 1 corresponds to complete
restriction of the vector with respect to overall molecule
motion, and relaxation is related solely to isotropic tum-
bling of the protein. These parameters can be linked to
models for motion, in our case, the wobble-in-a-cone
model [7]. Variations of Lipari-Szabo analysis exist for
cases involving ms timescale conformational exchange,
but no CI2 residue (except Thr40) has significant contri-
butions from slow exchange [9]. It is also possible to
study the equilibrium thermodynamic stability of globu-
lar proteins by using NMR.
Amide proton exchange experiments can be used
with NMR to assess protein stability. The technique
relies on the exchange of amide protons for deuterons
in a D
2
O solution. We have recently reviewed the
requirements for its application in crowded solution by
using NMR [10].
Exchange occurs via the scheme
cl 1H
k
op
k
cl
op 1Hkint
op 2H(Scheme 1
)
with opening rate k
op
,closingratek
cl
,andrateof
exchange from the open state k
int
. If the protein is stable
(k
cl
>>k
op
) and exchange from the open state is rate
limiting, the stability of an amide proton against
exchange (G0
o
p
) can be determined with the equation,
G0
op =RT ln kobs
k
int
where Ris the gas constant and Tis the absolute tem-
perature. The value of k
obs
, the overall rate of exchange
for any particular backbone amide proton, is assessed by
acquiring
1
H-
15
N heteronuclear single quantum correla-
tion (HSQC) spectra as a function of time after initiat-
ing exchange. As with dynamics, G0
o
p
can be quantified
on a per-residue basis. The largest G0
o
p
values match
the global protein stability values determined by other
methods (e.g., calorimetry, circular dichroism spectropo-
larimetry) [11].
Results
Experiments were performed by using samples compris-
ing 1 mM CI2 in 50 mM sodium acetate solution, pH
5.4 at 37°C. Crowded samples also contained 10 g/L
p-NIPAm-co-AAc microgels.
Polymer Characterization
The microgels composed of p-NIPAm-co-AAc have an
average hydrodynamic radius (R
H
) of 312 nm and an
average polydispersity of 7.4%. The molecular weight of
the microgels was estimated to be 1 GDa by multiple
angle laser light scattering [12].
Figure 1 Structure and Size of p-NIPAm-co-AAc:A)The
monomeric repeat of NIPAm. B) Overall shape and size of p-NIPAm-
co-AAc microgels.
Miklos et al.BMC Biophysics 2011, 4:13
http://www.biomedcentral.com/2046-1682/4/13
Page 2 of 7
Controls for Amide Proton Exchange
To determine whether exchange from the open state
(k
int
) is rate limiting, nuclear Overhauser enchancement
spectroscopy-detected amide proton exchange (NOESY-
HEX) experiments were performed [10]. The results are
given in Table 1, along with individual backbone residue
decay rates from HSQC-detected amide proton
exchange.
To determine whether k
int
values are changed by
crowding, phase-modulated clean exchange (CLEANEX-
PM) experiments [13] were used to determine k
int
for
residues on the extended loop region of CI2. For His37,
k
int
values were 11 ± 2 s
-1
in dilute solution and 8 ± 2 s
-1
in 10 g/L p-NIPAm-co-AAc.
Dynamics
Analysis of the T
1
,T
2
, and NOE data (Additional File 1)
acquired in dilute solution and in 10 g/L p-NIPAm-co-
AAc yielded the values for τ
m
,S
2
,andτ
e
.Thevalueofτ
m
was the same (4.1 ns) in dilute solution and in 10 g/L
p-NIPAm-co-AAc, and is consistent with the value
obtained by Shaw et al. in dilute solution [9]. Histograms
of S
2
and τ
e
versus residue number are shown in Figure 2.
Linear least squares analysis of a plot of S
2
in dilute solu-
tion versus S
2
in crowded solution gives a slope of 1.0 ±
0.1, a y-intercept of 0.1 ± 0.1 and an R
2
value of 0.80.
Amide Proton Exchange and Stability
Values for k
obs
were determined in triplicate for solutions
in the presence and absence of 10 g/L p-NIPAm-co-AAc.
Exchange was slowed in 10 g/L p-NIPAm-co-AAc com-
pared to dilute solution (Figure 3). Values of ΔG
0*op
were
determined by using values for k
int
calculated from
SPHERE [14] and k
obs
values from amide proton
exchange experiments. A listing of values is given in
Additional File 2. A histogram of ΔG
0*op
versus residue
number is shown in Figure 4.
Discussion
Thevolumeoccupancyofp-NIPAm-co-AAc solutions
defines the degree of crowding. Using a hydrodynamic
radius of 312 nm and a molecular weight of 1 GDa, the
microgel in a 10 g/L solution occupies ~70% of the solution
volume at pH 5.4 and 37°C (the conditions used in our
experiments). The practical limit of spherical packing is
64% volume occupancy [15], but soft materials such as
microgels can be overpacked[16]. Our solutions, however,
were still in the liquid state, meaning our value for volume
occupancy is likely an overestimate. The high value does,
however, suggest that experimental conditions were within
the realm of crowding, as other systems show crowding
effects at less than 20% volume occupancy [17,18].
Although the microgel slowed exchange (Figure 3), it
was necessary to perform control experiments to ensure
Table 1 NOESY-HEX results
Residue(s) k
obs
NOESY (s
-1
×10
5
)k
obs
HSQC (s
-1
×10
5
)
Leu8 3 3
Val9 2 2
Leu8 + Val9
a
55
Leu8, Val9
b
5 N/A
Lys17 52 40
Lys18 20 15
Lys17 + Lys18
a
72 55
Lys17, Lys18
b
50 N/A
Ala58 3 4
Glu59 3 3
Ala58 + Glu59
a
67
Ala58, Glu59
b
7 N/A
k
abs
values from NOESY-detected amide proton exchange and HSQC-detected
amide proton exchange for CI2 in 10 g/L p-NIPAm-co-AAc, 50 mM sodium
acetate, pH 5.4, 37°C.
a
Sum of values from individual crosspeak decays.
b
Exchange rate of amide-amide NOESY crosspeak.
Figure 2 CI2 dynamics in p-NIPAm-co-AAc: Order parameters
(upper panel) and timescales of internal motion (lower panel) for
CI2 in dilute solution and in 10 g/L NIPAm-AAc.
Miklos et al.BMC Biophysics 2011, 4:13
http://www.biomedcentral.com/2046-1682/4/13
Page 3 of 7
that stability values could be obtained under both sets of
conditions. First, we confirmed that amide proton
exchange from the open state (k
int
, Scheme 1) is rate
limiting. Under this condition, pairs of proximal amide
protons, A and B, open with the same frequency, but
with different k
int
values. That is, amide proton
exchanges for A and B are uncorrelated. By observing
the decay of an amide-amide NOESY crosspeak corre-
sponding to a resonance coupling between A and B, it is
possible to determine whether their exchange is corre-
lated or uncorrelated. If the exchange is uncorrelated,
the decay curve of the amide-amide crosspeak should
equal the product of the individual amide proton decay
curves [19,20],
I
AB
=I
A
·I
B
In this instance, the overall exchange rate of the
amide-amide crosspeak will correspond to the sum of
the individual exchange rate constants,
K
AB
=K
A
+K
B
All these rates can be assessed from a series of
15
N-fil-
tered
1
H-
1
H NOESY spectra acquired under exchange
conditions [10].
As shown in Table 1, the exchange rates observed for
the amide-amide crosspeaks for CI2 in both dilute solu-
tion and in 10 g/L p-NIPAm-co-AAc are, within the
uncertainty of the experiment, the sums of their respec-
tive individual exchange rates, indicating that the
exchanges are uncorrelated. We conclude that exchange
from the open state is rate limiting, allowing determina-
tion of stability from amide proton exchange rates.
Second, we must determine if the microgel changes
k
int
from the values determined in dilute solution. The
dilute solution value for each residue is calculated by
using the computer program, SPHERE [14] (http://www.
fccc.edu/research/labs/roder/sphere/). The program uses
values from the exchange of free peptides [21], and
relies solely on the primary structure of the test protein.
We assessed whether k
int
is affected by adding
p-NIPAm-co-AAc by using the CLEANEX-PM experi-
ment [13]. We measured the exchange rate of the His37
amide proton, which is fully exposed in the flexible loop
region of CI2 (residues 35-44). The data indicated that
the intrinsic rate of exchange in 10 g/L p-NIPAm-co-
AAc (8 ± 2 s
-1
) is within uncertainty of the value in
dilute solution (11 ± 2 s
-1
). These results suggest that
k
int
values can be used without alteration. Having shown
that it is valid to use k
obs
and k
int
values to obtain open-
ing free energies, we constructed histograms of ΔG
0*op
values versus residue number (Figure 4).
Dynamics and Stability
Crowding involves two different types of effects on pro-
tein stability: volume exclusion and chemical interactions.
Volume exclusion is expected to stabilize protein native
states, whereas chemical interactions can be stabilizing or
destabilizing [6]. Attractive chemical interactions are
expected to impede rotational dynamics, and the micro-
gel used here is known to have favorable electrostatic
interactions with proteins at low ionic strength [22]. Our
data were collected at pH 5.4, where the microgel is
negatively charged. The truncated form of CI2 we use
has an isoelectric point (pI) of 6. Therefore, the polymer
Figure 3 Exchange Curves: Amide proton exchange curves for
Lys24 in dilute solution (blue triangles) and in 10 g/L p-NIPAm-co-
AAc (green squares). Values for k
obs
are 7.53 ± 0.05 × 10
-5
±s
-1
in
dilute solution and 4.55 ± 0.05 × 10
-5
s
-1
in 10 g/L p-NIPAm-co-AAc.
These uncertainties are from non-linear least squares fitting and are
smaller than the uncertainty from triplicate analysis.
Figure 4 CI2 stability in p-NIPAm-co-AAc: Results are shown for
dilute solution (blue) and 10 g/L p-NIPAm-co-AAc (green). Error bars
reflect the standard error in k
obs
values from three trials. Colored
arrows indicate the average ΔG
0*op
values for globally exchanging
residues [20] in crowded (5.2 kcal/mol) and dilute (4.9 kcal/mol)
solution.
Miklos et al.BMC Biophysics 2011, 4:13
http://www.biomedcentral.com/2046-1682/4/13
Page 4 of 7
and CI2 are oppositely charged, and one might expect an
attractive interaction.
Our observation that the order parameters (S
2
), the
timescale of internal motion (τ
e
), and the rotational cor-
relation time (τ
m
), are unchanged by the polymer indi-
cates the absence of significant chemical interactions
between the polymer and CI2. The lack of interaction
probably arises because we used an ionic strength of 50
mM, which minimizes binding [22]. Therefore, we only
consider contributions from volume exclusion effects.
The patterns of ΔG
0*op
values along the amino acid
sequence (Figure 4) are the same in dilute solution as
they are in the microgel solution, suggesting that the
microgel does not alter the open states of CI2. The
ΔG
0*op
values in the microgel are uniformly larger than
the values for dilute solution, indicating the polymer sta-
bilizes the protein with a maximal stability increase of
approximately 0.4 kcal/mol. Averaging the ΔG
0*op
values
from residues known to be implicated in global unfold-
ing [20] show that the microgel increases the overall sta-
bility from 4.9 kcal/mol to 5.2 kcal/mol. We cannot
state with certainty that the increased stability arises
from the polymeric nature of the microgel because its
crosslinked nature makes determination of a suitable
monomer unit difficult.
Considering the volume fraction estimate of ~70%, a 0.3
kcal/mol stability increase is quite small. A modest
increase is anticipated, however, because the hydrodynamic
radius of CI2 is only 1% that of the p-NIPAm-co-AAc
microgels (Figure 5). In such a system, CI2 can occupy
interstitial spaces between p-NIPAm-co-AAc microgels,
putting CI2 in a dilute solution environment. Alternatively,
the microgel particles probably have pores large enough to
accommodate CI2 and water.
Next, we try to relate the stability change to the back-
bone dynamics data (Figure 2). The data indicate that
the increased stability does not alter the ps-ns backbone
dynamics. It has been proposed that stability changes
are associated with alterations of ps-ns backbone
dynamics [23,24]. Our results do not indicate a connec-
tion, because we observe increased stability without a
change in ps-ns timescale dynamics. The most straight-
forward conclusion is that stability is not linked to back-
bone ps-ns dynamics. It is possible, however, that
stability is reflected in slower (ms-s) motions [25].
Conclusions
Even though the 10 g/L solution of p-NIPAm-co-AAc
microgels occupy ~70% of solution volume, these condi-
tions do not affect the ps-ns timescale backbone
dynamics of CI2. The microgel, however, does have a
modest stabilizing effect on the protein. These conclu-
sions are explained by the fact that the majority of the
protein occupies a water-like environment in interstitial
spaces of the microgel particles. In the context of
p-NIPAm-co-AAc as a drug delivery tool, this is promising
information, supporting the notion that these microgels
are biocompatible materials. It seems likely, however, that
larger crowding agents such as p-NIPAm-co-AAc can
have more noticeable effects when present in mixed solu-
tions that also contain multiple sizes of crowders [26].
Methods
15
N-enriched CI2 was expressed and purified as
described by Miklos et al. [10].
Polymer Synthesis and Characterization
A general synthesis for NIPAm-AAc microgels is
described by Jones and Lyon [27], but variations yield
products with different properties (size, temperature, pH
dependence, etc.) [28-30]. The microgels used here were
prepared via aqueous, surfactant-free, free radical precipi-
tation polymerization using 70 mM total monomer con-
centration. Briefly, N-isopropylacrylamide (0.6973 g) and
N,N-methylenebis(acrylamide) (0.0215 g) were dissolved
in 99 mL of H2O and filtered through a 0.8 μmsyringe
filter into a round bottom flask. The mixture was
bubbled with N
2
(g) and heated to 70°C 2°C) over
~1 h. Acrylic acid (46 μL) was then added. Polymeriza-
tion was initiated by adding a solution of (NH
4
)
2
S
2
O
8
(0.0226 g) dissolved in 1 mL of H
2
O. This reaction was
stirredat70°C(±2°C)underablanketofN
2
(g) for 4 h
Figure 5 Interstitial Spaces in p-NIPAm-co-AAc: Depiction of the
scale of microgel sizes for p-NIPAm-co-AAc (green) and CI2 (red).
CI2 can exist in the spaces between crowder particles or within
pores (of unknown size) without experiencing a change in
environment compared to bulk water.
Miklos et al.BMC Biophysics 2011, 4:13
http://www.biomedcentral.com/2046-1682/4/13
Page 5 of 7
and was stirred and cooled overnight. The mixture was
filtered through Whatman #2 paper and stored. Ali-
quots of the resultant colloidal dispersion were purified
with centrifugation at 15,000 × g, decanted, and resus-
pended in H
2
O. This process was performed three
times. The particles were then lyophilized to yield a
white powder.
The microgels were characterized after suspension in
sodium acetate (pH 5.4) and passage through a 0.8 μm
filter. This solution was sonicated for 5 min, allowed to
equilibrate for 30 min, then analyzed by using multi-
angle laser light scattering (MALLS) [12].
NMR
HSQC-detected and NOESY-HEX experiments were
performed on a 500 MHz Varian Inova spectrometer
equipped with a triple-resonance HCN cold probe as
described by Miklos et al. [10]. CLEANEX-PM experi-
ments were conducted as described by Hwang et al.
[13] with a 600 MHz Varian Inova spectrometer
equipped with a triple-resonance HCN probe with
three-axis gradients system.
15
NT
1
and T
2
relaxation times and
15
N{
1
H} NOEs
were measured as described by Kay et al. [31]. Experi-
ments were performed on the 600 MHz spectrometer.
Lipari-Szabo model free analysis [7] was performed with
the software package Relaxn 2.2. [32]. The majority of
residues were fit with the original model-free formalism
[4] to yield τ
m
,S
2
and τ
e
.
Additional material
Additional file 1:
15
NT
1
,
15
NT
2
, and
1
H-
15
N NOEs for CI2. A table
containing
15
NT
1
,
15
NT
2
, and
1
H-
15
N NOE values for CI2 in dilute
solution and 10 g/L p-NIPAm-co-AAc at 37°C, pH 5.4.
Additional file 2: CI2 Stability Values. A table containing ΔG
0*op
values
and standard error from triplicate results for CI2 in 10 g/L p-NIPAm-co-
AAc at 37°C, pH 5.4.
Acknowledgements
This work was supported by the National Institutes of Health (5DP1OD783)
and the National Science Foundation (MCB-1051819). We thank Gregory B.
Young for spectrometer assistance and Elizabeth Pielak for helpful
comments on the manuscript.
Author details
1
Department of Chemistry, University of North Carolina, Chapel Hill, North
Carolina 27599, USA.
2
State Key Laboratory of Magnetic Resonance and
Molecular and Atomic Physics, Wuhan Institute of Physics and Mathematics,
Chinese Academy of Sciences, Wuhan, 430071, PR China.
3
School of
Chemistry & Biochemistry, Georgia Institute of Technology, Atlanta, GA
30332, USA.
4
Petit Institute for Bioengineering and Bioscience, Georgia
Institute of Technology, Atlanta, GA 30332, USA.
5
Department of Chemistry,
University of Alberta, Edmonton, AB, T6G 2G2, Canada.
6
Department of
Biochemistry and Biophysics, University of North Carolina, Chapel Hill, North
Carolina 27599, USA.
7
Lineberger Comprehensive Cancer Center, University
of North Carolina, Chapel Hill, North Carolina 27599, USA.
Authorscontributions
ACM, CL, LAL, and GJP designed the research; ACM and CL performed the
NMR experiments; CDS prepared and characterized microgels; ACM and GJP
wrote the manuscript; All authors read and approved the final manuscript.
Received: 15 December 2010 Accepted: 31 May 2011
Published: 31 May 2011
References
1. Zimmerman SB, Trach SO: Estimation of macromolecule concentrations
and excluded volume effects for the cytoplasm of Escherichia coli.J Mol
Biol 1991, 222(3):599-620.
2. Pelton R: Temperature-sensitive aqueous microgels. Adv Colloid Interface
Sci 2000, 85(1):1-33.
3. Jackson SE, Fersht AR: Folding of chymotrypsin inhibitor 2. 1. Evidence
for a two-state transition. Biochemistry 2002, 30(43):10428-10435.
4. Jarymowycz VA, Stone MJ: Fast time scale dynamics of protein
backbones: NMR relaxation methods, applications, and functional
consequences. Chemical Reviews 2006, 106(5):1624-1671.
5. Charlton LM, Barnes CO, Li C, Orans J, Young GB, Pielak GJ:
Macromolecular crowding effects on protein stability at the residue
level. J Am Chem Soc 2008, 130:6826-6830.
6. Miklos AC, Li CG, Sharaf NG, Pielak GJ: Volume exclusion and soft
interaction effects on protein stability under crowded conditions.
Biochemistry 2010, 49(33):6984-6991.
7. Lipari G, Szabo A: Model-free approach to the interpretation of nuclear
magnetic-resonance relaxation in macromolecules. 1. Theory and range
of validity. J Am Chem Soc 1982, 104(17):4546-4559.
8. Palmer AG: NMR probes of molecular dynamics: Overview and
comparison with other techniques. Annu Rev Biophys Biomol Struct 2001,
30:129-155.
9. Shaw GL, Davis B, Keeler J, Fersht AR: Backbone dynamics of
chymotrypsin inhibitor 2: Effect of breaking the active-site bond and its
implications for the mechanism of inhibition of serine proteases.
Biochemistry 1995, 34(7):2225-2233.
10. Miklos AC, Li C, Pielak GJ: Using NMR-detected backbone amide
1
H
exchange to assess macromolecular crowding effects on globular-
protein stability. Methods Enzymol 2009, 466:1-18.
11. Huyghues-Despointes BMP, Scholtz JM, Pace CN: Protein conformational
stabilities can be determined from hydrogen exchange rates. Nat Struct
Biol 1999, 6(10):910-912.
12. Sorrell CD, Lyon LA: Deformation controlled assembly of binary microgel
thin films. Langmuir 2008, 24(14):7216-7222.
13. Hwang TL, van Zijl PCM, Mori S: Accurate quantitation of water-amide
exchange rates using the phase-modulated CLEAN chemical EXchange
(CLEANEX-PM) approach with a fast-HSQC (FHSQC) detection scheme. J
Biomol NMR 1998, 11:221-226.
14. Zhang YZ: Protein and peptide structure and interactions studied by
hydrogen exchange and NMR. PA, USA: University of Pennsylvania; 1995.
15. Aste T, Weaire D: The Pursuit of Perfect Packing. New York: Taylor &
Francis;, 2 2008.
16. Lyon LA, Meng ZY, Singh N, Sorrell CD, John AS: Thermoresponsive
microgel-based materials. Chemical Society Reviews 2009, 38(4):865-874.
17. Homouz D, Stagg L, Wittung-Stafshede P, Cheung MS: Macromolecular
crowding modulates folding mechanism of α/βprotein apoflavodoxin.
Biophys J 2009, 96(2):671-680.
18. Ping GH, Yang GL, Yuan HM: Depletion force from macromolecular
crowding enhances mechanical stability of protein molecules. Polymer
2006, 47(7):2564-2570.
19. Wagner G: A novel application of nuclear Overhauser enhancement
(NOE) in proteins: Analysis of correlated events in the exchange of
internal labile protons. Biochem Biophys Res Commun 1980, 97(2):614-620.
20. Neira JL, Itzhaki LS, Otzen DE, Davis B, Fersht AR: Hydrogen exchange in
chymotrypsin inhibitor 2 probed by mutagenesis. J Mol Biol 1997,
270(1):99-110.
21. Bai Y, Milne JS, Mayne L, Englander SW: Primary structure effects on
peptide group hydrogen exchange. Proteins: Struct, Funct, Genet 1993,
17(1):75-86.
22. Chen XW, Chen SA, Wang JH: A pH-responsive poly(N-
isopropylacrylamide-co-acrylic acid) hydrogel for the selective isolation
of hemoglobin from human blood. Analyst 2010, 135(7):1736-1741.
Miklos et al.BMC Biophysics 2011, 4:13
http://www.biomedcentral.com/2046-1682/4/13
Page 6 of 7
23. Boyer JA, Lee AL: Monitoring aromatic picosecond to nanosecond
dynamics in proteins via C-13 relaxation: Expanding perturbation
mapping of the rigidifying core mutation, V54A, in eglin c. Biochemistry
2008, 47(17):4876-4886.
24. Vicky DN, Loria JP: The effects of cosolutes on protein dynamics: The
reversal of denaturant-induced protein fluctuations by trimethylamine
N-oxide. Protein Sci 2007, 16(1):20-29.
25. Henzler-Wildman K, Kern D: Dynamic personalities of proteins. Nature
2007, 450(7172):964-972.
26. Jyotica B, Ke X, Huan-Xiang Z: Nonadditive effects of mixed crowding on
protein stability. Proteins: Struct, Funct, Bioinf 2009, 77(1):133-138.
27. Jones CD, Lyon LA: Synthesis and characterization of multiresponsive
core-shell microgels. Macromolecules 2000, 33(22):8301-8306.
28. Serpe MJ, Jones CD, Lyon LA: Layer-by-layer deposition of
thermoresponsive microgel thin films. Langmuir 2003, 19(21):8759-8764.
29. Blackburn WH, Lyon LA: Size-controlled synthesis of monodisperse core/
shell nanogels. Colloid and Polymer Science 2008, 286(5):563-569.
30. Meng ZY, Smith MH, Lyon LA: Temperature-programmed synthesis of
micron-sized multi-responsive microgels. Colloid and Polymer Science
2009, 287(3):277-285.
31. Kay LE, Torchia DA, Bax A: Backbone dynamics of proteins as studied by
15
N inverse detected heteronuclear NMR spectroscopy: Application to
staphylococcal nuclease. Biochemistry 1989, 28(23):8972-8979.
32. Lee AL, Wand AJ: Assessing potential bias in the determination of
rotational correlation times of proteins by NMR relaxation. J Biomol NMR
1999, 13(2):101-112.
doi:10.1186/2046-1682-4-13
Cite this article as: Miklos et al.: An upper limit for macromolecular
crowding effects. BMC Biophysics 2011 4:13.
Submit your next manuscript to BioMed Central
and take full advantage of:
Convenient online submission
Thorough peer review
No space constraints or color figure charges
Immediate publication on acceptance
Inclusion in PubMed, CAS, Scopus and Google Scholar
Research which is freely available for redistribution
Submit your manuscript at
www.biomedcentral.com/submit
Miklos et al.BMC Biophysics 2011, 4:13
http://www.biomedcentral.com/2046-1682/4/13
Page 7 of 7
... There was a rapid need to reform the legislative norms for medical devices. Some of the notable points are: [21][22][23] Enhanced patient safety The primary motivation behind the MDR was to strengthen patient safety by addressing shortcomings in the existing regulatory framework. This involved strengthening the standards for clinical review, improving post-market surveillance, and boosting medical device traceability and transparency. ...
Article
Full-text available
The present review article emphasizes the pivotal role of ISO 13485:2016 in facilitating the global harmonization of medical device regulations. Compliance with this standard is crucial for manufacturers aiming to access international markets, including India, Europe, and the USA. Harmonization of regulations simplifies the process of obtaining licenses and approvals, reducing burdens on manufacturers and enhancing patient safety. By implementing effective quality management systems, manufacturers can navigate the complex regulatory landscape and contribute to the global healthcare industry. The review article also underscores the diversity of medical devices available and acknowledges the substantial expansion of the Indian market. It discusses the stringent regulations outlined in the Indian Medical Device Rules (IMDR) 2017 and the challenges faced by nations in accessing high-quality medical devices. Furthermore, it touches upon the European Medical Device Regulations and the dynamic regulatory environment in the USA. In conclusion, the paper underscores the importance of ISO 13485:2016 in achieving global and regional harmonization of medical device regulations, thereby facilitating market access for manufacturers and confirming the assurance of safe and effective medical device products to patients worldwide.
... It could serve as a reliable prediction marker for malaria severity where other well-practiced or expensive laboratory methodsare inaccessible. According to a previous study, a CRP level of <20 mg/L indicates uncomplicated malaria as it could differentiate malaria cases from healthy controls [46]. The cut-off value of 18.5 mg/L for CRP level with a sensitivity of 71.4 %, specificity of 68.7 %, and AUC of 78 % was reported in the study by Bhardwaj et al; they suggested CRP adequacy in differentiating uncomplicated malaria from healthy controls [9]. ...
Article
Inflammatory biomarkers; C-reactive protein (CRP), Interleukin 6 (IL-6), and tumor necrosis factor- alpha (TNF-α) play a very crucial role in disease pathogenesis. Studies conducted earlier showed the associativity of these biomarkers with malaria severity. Meta-analysis of individual biomarkers was done in many studies, while in a few others, all these candidates were estimated, but the findings were inconclusive. Therefore, a systematic review and meta-analyses were performed to evaluate differences in biomarkers mentioned above in complicated and uncomplicated malaria patients. Studies focussed on CRP, IL-6, and TNF-α with quantitative data on complicated and uncomplicated malaria patients were searched on PubMed, Scopus, and Google Scholar. The quality of the studies selected for this review was checked following Newcastle-Ottawa Scale guidelines. The standard mean difference and confidence interval of biomarkers in the targeted groups were calculated using the random effects model. Egger's test and funnel plot asymmetry were performed to assess the publication bias. Thirteen studies that qualified the inclusion criteria were considered for this meta-analysis. CRP levels were higher in complicated malaria patients than uncomplicated ones (P < 0.00001, pooled SMD: 0.90 mg/L, 95 % CI: 0.51 to 1.30 mg/L, I2: 80 %, six studies). IL-6 levels were elevated in complicated cases (P < 0.00001, pooled SMD: 0.89 pg/ml, 95 % CI: 0.66 to 1.12, I2: 99 %, four studies) and TNF-α also showed an increase in severe complicated patients (P < 0.00001, pooled SMD: 1.18 pg/ml, 95 % CI: 1 to 1.36, I2: 99 %, six studies). In most of the included studies, CRP, IL-6, and TNF-α were higher in complicated malaria patients. Nevertheless, the results of a few studies were not convincing. Due to the lack of specificity in all individual biomarkers, none had adequate diagnostic accuracy. Considering the role of pro-inflammatory cytokines in the CRP activation pathway in malaria progression, the combination of these biomarkers should be used in monitoring the disease severity.
... The model monomer was not studied. An examination of the effect of a microgel comprising 10 9 -Da poly(N-isopropylacrylamide-coacrylic acid) on the stability and dynamics of chymotrypsin inhibitor 2 reveals only a small stabilizing effect and no effect on nanosecond internal dynamics compared to buffer, showing the size limit of crowding effects (129). Another recent effort shows that polyacetate destabilizes Cu/Zn superoxide dismutase (192). ...
Article
Cells are crowded, but proteins are almost always studied in dilute aqueous buffer. We review the experimental evidence that crowding affects the equilibrium thermodynamics of protein stability and protein association and discuss the theories employed to explain these observations. In doing so, we highlight differences between synthetic polymers and biologically relevant crowders. Theories based on hard-core interactions predict only crowding-induced entropic stabilization. However, experiment-based efforts conducted under physiologically relevant conditions show that crowding can destabilize proteins and their complexes. Furthermore, quantification of the temperature dependence of crowding effects produced by both large and small cosolutes, including osmolytes, sugars, synthetic polymers, and proteins, reveals enthalpic effects that stabilize or destabilize proteins. Crowding-induced destabilization and the enthalpic component point to the role of chemical interactions between and among the macromolecules, cosolutes, and water. We conclude with suggestions for future studies. Expected final online publication date for the Annual Review of Biophysics, Volume 51 is May 2022. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
... 79,81 We note that this is not the first case where the cellular environment had no discernable effects on molecular structure. [84][85][86] A similar lack of structural effects was also reported for fluorescently-labeled polyethylene glycol in live cells and dilute aqueous buffer. 87 Thus, the effects of molecular crowding and the cellular environment are clearly molecule specific. ...
Article
Full-text available
Many proteins are composed of independently-folded domains connected by flexible linkers. The primary sequence and length of such linkers can set the effective concentration for the tethered domains, which impacts rates of association and enzyme activity. The length of such linkers can be sensitive to environmental conditions, which raises questions as to how studies in dilute buffer relate to the highly-crowded cellular environment. To examine the role of linkers in domain separation, we measured Fluorescent Protein-Fluorescence Resonance Energy Transfer (FP-FRET) for a series of tandem FPs that varied in the length of their interdomain linkers. We used discrete molecular dynamics to map the underlying conformational distribution, which revealed intramolecular contact states that we confirmed with single molecule FRET. Simulations found that attached FPs increased linker length and slowed conformational dynamics relative to the bare linkers. This makes the CLYs poor sensors of inherent linker properties. However, we also showed that FP-FRET in CLYs was sensitive to solvent quality and macromolecular crowding making them potent environmental sensors. Finally, we targeted the same proteins to the plasma membrane of living mammalian cells to measure FP-FRET in cellulo. The measured FP-FRET when tethered to the plasma membrane was the same as that in dilute buffer. While caveats remain regarding photophysics, this suggests that the supertertiary conformational ensemble of these CLY proteins may not be affected by this specific cellular environment.
... The model monomer was not studied. An examination of the effect of a microgel comprising 10 9 -Da poly(N-isopropylacrylamide-coacrylic acid) on the stability and dynamics of chymotrypsin inhibitor 2 reveals only a small stabilizing effect and no effect on nanosecond internal dynamics compared to buffer, showing the size limit of crowding effects (129). Another recent effort shows that polyacetate destabilizes Cu/Zn superoxide dismutase (192). ...
Article
Full-text available
Background: In the disability sector globally, and specifically in Australia, assessments of functioning have become key to diagnostic processes, and accessing therapy and funding. Over half of all individuals accessing support through Australia’s National Disability Insurance Scheme have a neurodevelopmental condition diagnosis. Little is known about assessments of functioning for this population. Methods: A mixed methods online survey was designed to understand the current assessment of functioning practices (including clinical contexts, concepts being assessed, and assessment methods) and barriers and facilitators to clinicians using best practice. Results were analysed descriptively, and differences between professions calculated where possible. Content analysis was used to explore qualitative comments. Results: Clinicians from various medical and allied health backgrounds completed the survey (n = 93), with varying ranges of age, experience, and education. Clinicians reported that they assessed functioning across age, setting, sector, funding body, and individuals with a wide variety of diagnoses. Missing from current practice is a clear transdisciplinary conceptualisation of functioning. The largest barriers to best practice were limited time, large caseloads, availability of appropriate tools, and lack of clarity from funding bodies. Conclusions: Missing from current practice is a clear transdisciplinary conceptualisation of functioning. These results will help inform steps forward to improve assessment of functioning practices to ensure that all individuals receive appropriate and sufficient support.
Article
Nearly all biological processes, including strictly regulated protein-protein interactions fundamental in cell signaling, occur inside living cells where the concentration of macromolecules can exceed 300 g/L. One such interaction is between a 7 kDa SH3 domain and a 25 kDa intrinsically disordered region of Son of Sevenless (SOS). Despite its key role in the mitogen activated protein kinase signaling pathway of all eukaryotes, most biophysical characterizations of this complex are performed in dilute buffered solutions where cosolute concentrations rarely exceed 10 g/L. Here, we investigate the effects of proteins, sugars, and urea, at high g/L concentrations, on the kinetics and equilibrium thermodynamics of binding between SH3 and two SOS-derived peptides using 19F NMR lineshape analysis. We also analyze the temperature-dependence, which enables quantification of the enthalpic and entropic contributions. The energetics of SH3-peptide binding in proteins differs from those in the small molecules we used as control cosolutes, demonstrating the importance of using proteins as physiologically-relevant cosolutes. Although the majority of the protein cosolutes destabilize the SH3-peptide complexes, the effects are non-generalizable and there are subtle differences, which are likely from weak nonspecific interactions between the test proteins and the protein crowders. We also quantify the effects of cosolutes on SH3 translational and rotational diffusion to rationalize the effects on association rate constants. The absence of a correlation between the SH3 diffusion data and the kinetic data in certain cosolutes suggests that the properties of the peptide in crowded conditions must be considered when interpreting energetic effects. These studies have implications for understanding protein-protein interactions in cells and show the importance of using physiologically-relevant cosolutes for investigating macromolecular crowding effects.
Article
Extra virgin olive oil (EVOO) polyphenols, including the secoiridoids oleocanthal (OLC) and oleacein (OLE), are attracting attention because of their beneficial effects on health. Data on OLC and OLE bioavailability are scarce, as most research on EVOO polyphenols has concentrated on hydroxytyrosol, tyrosol, and oleuropein. Consequently, relevant goals for future research are the elucidation of OLC and OLE bioavailability and finding evidence for their beneficial effects through pre-clinical and clinical studies. The aim of this review is to shed light on OLC and OLE, focusing on their precursors in the olive fruit and the impact of agronomic and processing factors on their presence in EVOO. Also discussed are their bioavailability and absorption, and finally, their bioactivity and health-promoting properties.
Article
Background Primary care providers encounter a large proportion of the population with depression. Yet, many primary care patients with depression remain undiagnosed and untreated. Objective This study aims to examine depression screening patterns and the role of screening in depression diagnosis and treatment in the outpatient primary care setting. Design This is a cross-sectional analysis of nationally representative survey data of visits to outpatient physician offices from the 2005 to 2015 National Ambulatory Medical Care Surveys. Participants The sample included the first visit in the past year to a primary care provider by patients 12 years and older (N = 16,887). Methods The associations of visit characteristics with depression screening and of depression screening with depression diagnosis and treatment during the visit were assessed using logistic regression. Logistic regression with propensity score weighting was used to estimate the odds of depression diagnosis and treatment under the counterfactual scenario in which patients who visited providers with lower depression screening rates had visited providers with higher screening rates instead. All models were adjusted for patient and visit characteristics. Key Results A small proportion of sample visits involved depression screening (3.0%). Visits by patients with depressive symptom complaints were associated with higher odds of depression screening than other visits. When visits were weighted to have similar demographic and clinical characteristics, visits to providers with higher screening rates had higher odds of diagnosis (OR = 1.99, p < 0.001) and treatment (OR = 1.61, p = 0.001) compared to visits to providers with lower screening rates. Conclusions Physicians appear to use depression screening selectively based on patients’ presenting symptoms. Higher screening rates were associated with higher odds of depression diagnosis and treatment, and even modest increases in screening rates could meaningfully increase population-level rates of depression identification and treatment in primary care. Future research is needed to identify barriers to depression care and implement systematic interventions to improve services and patient outcomes.
Article
Full-text available
A new synthetic protocol for the synthesis of large diameter (2.5 to 5μm), temperature-, and pH-responsive microgels via aqueous surfactant-free radical precipitation copolymerization is presented. We have found that in this size range, which is not typically attainable using traditional dispersion polymerization approaches, excellent monodispersity and size control are achieved when the synthesis employs a programmed temperature ramp from 45 to 65°C during the nucleation stage of the polymerization. A combined kinetic and thermodynamic hypothesis for large particle formation under these conditions is described. Particle sizes, volume phase transition temperatures, and pH responsivity were characterized by particle tracking and photon correlation spectroscopy to illustrate their similar behavior to particles made via more traditional routes. These particles have been enabling for various studies in our group where microscopic visualization of the particles is required.
Article
Hydrogen exchange is widely used, especially in the study of protein stability and dynamics, folding, and interactions. Most applications require quenching of the exchange. Some of the methods have limitations due to insufficient quench. We studied the quenching properties of solvent mixtures. The pH dependence of the exchange rates in DMSO/D$\sb2$O mixtures is the same as that in water, but with suppressed minimum rate, k$\sb{\rm min}$, and elevated optimal pH. The decrease of k$\sb{\rm min}$ is due to the dramatic reduction of the base catalyzed exchange caused mainly by the changes in the relevant pKa's. By solvent quenching, it is possible to measure exchange rates for peptides in situations where direct NMR measurement is not applicable due to overlap in 1D and exchange during 2D experiments. This new approach was used to measure the exchange rates of monomeric melittin in aqueous solution and when bound to calmodulin. Our results indicate that the C-terminal half of melittin binds calmodulin more tightly then the N-terminal half. The protection pattern is consistent with bound melittin being $\alpha$-helical. The new method allows more detailed mapping of the binding sites involved in protein-protein interactions since one can now observe the protected amide protons on protein surface, which exchange rapidly in the free protein.^ A synthetic peptide (CHIGPGRAFC), derived from the V3 region of gp120 of HIV-1, in aqueous solution was investigated by NMR. The spectrum of the reduced peptide is characteristic of an unstructured, highly flexible coil. Disulfide bond cyclization results in major changes in chemical shifts and gives rise to NOE's indicative of well-defined turns. The presence of these turns is supported by the low temperature coefficients of the chemical shift of the amide protons of Arg7 and Phe9. The conformation of the peptide is in good agreement with that reported for this sequence as part of a longer peptide bound to an HIV neutralizing antibody, which might be similar to that of the corresponding fragment on the surface of the intact virus. The artificially introduced disulfide bond appears to reinforce the inherent structural tendencies.
Article
The reversible folding and unfolding of barley chymotrypsin inhibitor 2 (CI2) appears to be a rare example in which both equilibria and kinetics are described by a two-state model. Equilibrium denaturation by guanidinium chloride and heat is completely reversible, and the data can be fitted to a simple two-state model involving only native and denatured forms. The free energy of folding in the absence of denaturant, DELTA-G(H2O) at pH 6.3, is calculated to be 7.03 +/- 0.16 and 7.18 +/- 0.43 kcal mol-1 for guanidinium chloride and thermal denaturation, respectively. Scanning microcalorimetry shows that the ratio of the van't Hoff enthalpy of denaturation to the calorimetric enthalpy of denaturation does not deviate from unity, the value observed for a two-state transition, over the pH range 2.2-3.5. The heat capacity change for denaturation is found to be 0.789 kcal mol-1 K-1. The rate of unfolding of CI2 is first order and increases exponentially with increasing guanidinium chloride concentration. Refolding, however, is complex and involves at least three well-resolved phases. The three phases result from heterogeneity of the unfolded form due to proline isomerization. The fast phase, 77% of the amplitude, corresponds to the refolding of the fraction of the protein that has all its prolines in a native trans conformation. The rate of this major phase decreases exponentially with increasing guanidinium chloride concentration. The unfolding and refolding kinetics can also be fitted to a two-state model. Importantly, DELTA-G(H2O) and m, the constant of proportionality of the free energy of folding with respect to guanidinium chloride concentration, calculated from the kinetic experiments, 7.24 +/- 0.22 kcal mol-1 and 1.86 +/- 0.05 kcal mol-1 M-1, respectively, agree, within experimental error, with the values measured from the equilibrium experiments. This is perhaps the strongest evidence that the unfolding of CI2 follows a simple two-state transition.
Article
This paper describes the use of novel two-dimensional nuclear magnetic resonance (NMR) pulse sequences to provide insight into protein dynamics. The sequences developed permit the measurement of the relaxation properties of individual nuclei in macromolecules, thereby providing a powerful experimental approach to the study of local protein mobility. For isotopically labeled macromolecules, the sequences enable measurement of heteronuclear nuclear Overhauser effects (NOE) and spin-lattice (Tâ) and spin-spin (Tâ) ¹⁵N or ¹³C relaxation times with a sensitivity similar to those of many homonuclear ¹H experiments. Because T⁠values and heteronuclear NOEs are sensitive to high-frequency motions (10⁸-10¹²s⁻¹) while Tâ values are also a function of much slower processes, it is possible to explore dynamic events occurring over a large time scale. The authors have applied these techniques to investigate the backbone dynamics of the protein staphylococcal nuclease (S. Nase) complexed with thymidine 3â²,5â²-bisphosphate (pdTp) and Ca{sup 2+} and labeled uniformly with ¹⁵N. Tâ, Tâ, and NOE values were obtained for over 100 assigned backbone amide nitrogens in the protein. Values of the order parameter (S), characterizing the extent of rapid ¹H-¹⁵N bond motions, have been determined. These results suggest that there is no correlation between these rapid small amplitude motions and secondary structure for S. Nase. In contrast, ¹⁵N line widths suggest a possible correlation between secondary structure and motions on the millisecond time scale. In particular, the loop region between residues 42 and 56 appears to be considerably more flexible on this slow time scale than the rest of the protein.
Article
ChemInform is a weekly Abstracting Service, delivering concise information at a glance that was extracted from about 200 leading journals. To access a ChemInform Abstract of an article which was published elsewhere, please select a “Full Text” option. The original article is trackable via the “References” option.
Article
We report the synthesis and characterization of temperature and pH responsive hydrogel particles (microgels) with core−shell morphologies. Core particles composed of cross-linked poly(N-isopropylacrylamide) (p-NIPAm) or poly(NIPAm-co-acrylic acid) (p-NIPAm-AAc) were synthesized via precipitation polymerization and then used as nuclei for subsequent polymerization of p-NIPAm-AAc and p-NIPAm, respectively. The presence of a core−shell morphology was confirmed by transmission electron microscopy (TEM). Thermally initiated volume phase transitions were interrogated via temperature-programmed photon correlation spectroscopy (TP-PCS) as a function of solution pH. The p-NIPAm-AAc core hydrogel displays both a strong temperature and pH dependence on swelling. However, both p-NIPAm-AAc (core)/p-NIPAm (shell) and p-NIPAm (core)/p-NIPAm-AAc (shell) particles display a more complex pH dependence than the homogeneous particles. Specifically, a multistep volume phase transition appears when the AAc component becomes highly charged at a high pH. It is apparent from the measured deswelling curves that a portion of the particle swelling behavior is dominated by p-NIPAm, regardless of its location in the particle. However, deswelling behavior that is due to a mixture of p-NIPAm-AAc and p-NIPAm is evident, as well as a regime that is largely attributed to p-NIPAm-AAc alone. Small differences in the effect of pH on the two core−shell particles indicate that the influence of p-NIPAm is somewhat greater when it is localized in the shell.
Article
We report investigations of the assembly of thermoresponsive, 4-acrylamidofluorescein-modified (indicated by an asterisk) poly(N-isopropylacrylamide-co-acrylic acid) (pNIPAm-co-AAc*) microgel thin films via a traditional alternate layer deposition protocol. In this work, pNIPAm-co-AAc* microgels have been synthesized and incorporated into thin films by alternatively exposing glass slides functionalized with 3-aminopropyltrimethoxysilane to pNIPAm-co-AAc* microgels, which act as polyanions, and poly(allylamine hydrochloride), which acts as a polycation. Using this method, pNIPAm-co-AAc* microgel thin films have been constructed as confirmed by fluorescence microscopy. Investigations were also conducted on the effect of the microgel deposition temperature on the film morphology and thermoresponsivity. In addition, the thermoresponsivities of the pNIPAm-co-AAc* microgel thin films were studied with respect to the microgel layer number, microgel deposition conditions, and pH. As expected from the solution behavior of the microgels, the films exhibited enhanced thermoresponsivity and hindered deswelling at pH values below and above the pKa of the microgel acrylic acid (AAc) groups, respectively. This result is expected due to charge repulsion of the AAc groups in the microgel network, which directly impacts the thermal deswelling of the hydrogel.
Article
In crowded solutions the presence of many cosolutes often affects the stability of compact polymers, such as globular proteins. Important examples of crowded environments are those inside some cells, where protein stability or aggregation rates are affected by the presence of co-existing bio-macromolecules. In the present article the concept of depletion force from colloidal physics and theoretical techniques developed for polymer science have been applied to study the effects of macromolecular crowding on protein stability. A continuous three-dimensional polymer model is used to simulate the behavior of protein under the conditions of macromolecular crowding and the depletion force based on such a model is calculated. Calculated results have been compared with the measured results in our laboratory, where the enhancement of the forces required to unfold ubiquitin molecules in a solution crowded with dextran has been measured using single-molecule atomic force microscopy techniques. Comparison between the calculated results and experimental observations shows that only qualitative agreement has been reached in the sense that both show a larger force required because of crowding as a protein molecule is mechanical stretched, but the magnitude of the enhancement of the unfolding force theoretically predicted is small compared to the measured value. Possible sources of discrepancy and improvements of the model are discussed.
Article
Two-dimensional NMR spectroscopy has been used to monitor hydrogen-deuterium exchange in chymotrypsin inhibitor 2. Application of two independent tests has shown that at pH 5.3 to 6.8 and 33 to 37°C, exchange occurs via an EX2 limit. Comparison of the exchange rates of a number of mutants of CI2 with those of wild-type identifies the pathway of exchange, whether by local breathing, global unfolding or a mixture of the two pathways. For a large number of residues, the exchange rates were unaffected by mutations which destabilised the protein by up to 1.9 kcal mol−1, indicating that exchange is occurring through local fluctuations of the native state. A small number of residues were found for which the mutations had the same effect on the rate constants for exchange as on the equilibrium constant for unfolding, indicating that these residues exchange by global unfolding. These are residues that have the slowest exchange rates in the wild-type protein. We see no correspondence between these residues and residues involved in the nucleation site for the folding reaction identified by protein engineering studies. Rather, the exchange behaviour of CI2 is determined by the native structure: the most protected amide protons are located in regions of hydrogen bonding, specifically the C terminus of the α-helix and the centre of the β-sheet. A number of the most slowly exchanging residues are in the hydrophobic core of the protein.