ArticlePDF Available

Fluorescent protein barrel fluctuations and oxygen diffusion pathways in mCherry

Authors:

Abstract and Figures

Fluorescent proteins (FPs) are valuable tools as biochemical markers for studying cellular processes. Red fluorescent proteins (RFPs) are highly desirable for in vivo applications because they absorb and emit light in the red region of the spectrum where cellular autofluorescence is low. The naturally occurring fluorescent proteins with emission peaks in this region of the spectrum occur in dimeric or tetrameric forms. The development of mutant monomeric variants of RFPs has resulted in several novel FPs known as mFruits. Though oxygen is required for maturation of the chromophore, it is known that photobleaching of FPs is oxygen sensitive, and oxygen-free conditions result in improved photostabilities. Therefore, understanding oxygen diffusion pathways in FPs is important for both photostabilites and maturation of the chromophores. In this paper, we use molecular dynamics calculations to investigate the protein barrel fluctuations in mCherry, which is one of the most useful monomeric mFruit variant. We employ implicit ligand sampling to determine oxygen pathways from the bulk solvent into the mCherry chromophore in the interior of the protein. We also show that these pathways can be blocked or altered and barrel fluctuations can be reduced by strategic amino acid substitutions.
Content may be subject to copyright.
THE JOURNAL OF CHEMICAL PHYSICS 135, 235101 (2011)
Fluorescent protein barrel fluctuations and oxygen diffusion pathways
in mCherry
Prem P. Chapagain,a) Chola K. Regmi, and William Castillo
Department of Physics, Florida International University, Miami, Florida 33199, USA
(Received 1 August 2011; accepted 19 October 2011; published online 15 December 2011)
Fluorescent proteins (FPs) are valuable tools as biochemical markers for studying cellular processes.
Red fluorescent proteins (RFPs) are highly desirable for in vivo applications because they absorb
and emit light in the red region of the spectrum where cellular autofluorescence is low. The naturally
occurring fluorescent proteins with emission peaks in this region of the spectrum occur in dimeric
or tetrameric forms. The development of mutant monomeric variants of RFPs has resulted in sev-
eral novel FPs known as mFruits. Though oxygen is required for maturation of the chromophore, it
is known that photobleaching of FPs is oxygen sensitive, and oxygen-free conditions result in im-
proved photostabilities. Therefore, understanding oxygen diffusion pathways in FPs is important for
both photostabilites and maturation of the chromophores. In this paper, we use molecular dynamics
calculations to investigate the protein barrel fluctuations in mCherry, which is one of the most useful
monomeric mFruit variant. We employ implicit ligand sampling to determine oxygen pathways from
the bulk solvent into the mCherry chromophore in the interior of the protein. We also show that these
pathways can be blocked or altered and barrel fluctuations can be reduced by strategic amino acid
substitutions. © 2011 American Institute of Physics. [doi:10.1063/1.3660197]
I. INTRODUCTION
Fluorescent proteins (FPs) are valuable tools in molec-
ular and cell biology and are used as biochemical markers
for studying cellular processes.13Red fluorescent proteins
(RFPs) are highly desirable for in vivo applications because
they absorb and emit light in the red region of the spectrum
where cellular autofluorescence is low.4However, the natu-
rally occurring fluorescent proteins with emission peaks in
this region of the spectrum occur in dimeric or tetrameric
forms,5,6which tend to oligomerize7,8and render them un-
suitable for fusion tagging.9The development of mutant
monomeric variants of RFPs to avoid these issues has re-
sulted in several novel monomeric FPs known as mFruits.10
Some of the most promising mFruits are mCherry, mOrange,
and mStrawberry11 and their names reflect the wavelengths of
their corresponding emission spectra.
Though oxygen is required for maturation of the chro-
mophore, it is known that photobleaching of FPs is oxy-
gen sensitive, and oxygen-free conditions result in improved
photostabilities.12 This poses limitations to the next genera-
tion of single-molecule spectroscopy and low-copy fluores-
cence microscopy experiments and therefore, improving the
photostability of the mFruits is highly desirable. The photo-
bleaching of the monomeric variants of RFPs has been at-
tributed to the lack of proper shielding against oxygen or
other molecules by the beta barrel surface.7Increasing evi-
dence suggests that protein flexibility plays a major role in
gas access into many proteins1318 and the dynamic fluctu-
ations in the size of transient cavities due to residues’ ther-
mal fluctuations are the determining factor in the pathways
a)Author to whom correspondence should be addressed. Electronic mail:
chapagap@fiu.edu.
of gas diffusion.1922 Oxygen diffusion in myoglobin’s dis-
tal pocket has been extensively studied, both experimentally
and by simulations, in light of the influence of different pro-
tein conformations or mutations.14,23 In FPs, the interaction
between the chromophore and the surrounding protein has
important implications for both structures.24 The electronic
molecular orbitals of the chromophore that are responsible
for its spectral properties may be modified by the surround-
ing protein. Also, protein barrel fluctuations and the motion
of residues can affect the spectral properties and the lifetime
of the fluorescence.25
Recent developments in efficient computational sampling
methods have allowed thorough scanning of the possible path-
ways for gas diffusion in the interior of proteins.2629 For ex-
ample, such computational investigations have proved use-
ful in understanding gas diffusion in many protein systems
such as molecular dioxygen pathways via dynamic oxygen
access channels in flavoproteins,3032 ammonia transport in
carbamoyl phosphate synthetase,3335 and gas diffusion and
channeling in hemoglobin.28,36 In this paper, we use explicit
solvent all-atom molecular dynamics (MD) simulations to in-
vestigate the protein barrel fluctuations in mCherry, which is
one of the most useful monomeric variants of RFP. We com-
pare the barrel fluctuations in mCherry to those in citrine,
a yellow variant (YFP) of green fluorescent protein (GFP).
Although citrine and mCherry belong to different FP fami-
lies and the photobleaching mechanisms can be different, we
compare the barrel structural integrity of these proteins due
to two main reasons. First, citrine is a natural monomer and
a GFP homologue of mCherry with a similar barrel structure
and second, it is the most useful FP among all YFPs due to
its reduced halide sensitivity and improved photstability.37,38
Other YFPs are very sensitive to halides due to easy ion
0021-9606/2011/135(23)/235101/6/$30.00 © 2011 American Institute of Physics135, 235101-1
235101-2 Chapagain, Regmi, and Castillo J. Chem. Phys. 135, 235101 (2011)
FIG. 1. (a) Superposition of ribbon structures of red (mCherry) and yellow
(citrine) fluorescent proteins. (b) The β7-β10 region is displayed with a space
filling model which shows that the gap in mCherry is larger than in citrine.
access via a solvent channel or cavity formed close to the
dimer interface.39,40 In citrine, this cavity is filled by a mu-
tation Gln69Met preventing the access to the ion.37 A simi-
lar effect is desired in mCherry. We employ implicit ligand
sampling (ILS) to determine oxygen pathways from the bulk
solvent into the mCherry active site. Using these results as
a guide, we show that the barrel fluctuations and the oxygen
pathways can be altered or blocked with strategic amino acid
substitutions.
Following the folding of the protein, the chromophore
formation involves cyclization of tripeptide and oxidation,
which requires molecular oxygen.41 Therefore, the maturation
times can depend on the accessibility of molecular oxygen.
For example, it is shown that a water-filled pore was essen-
tial for fast maturation of TurboGFP chromophore.42 The pore
that leads from outside of the barrel to the inside chromophore
possibly facilitates molecular oxygen entry. Upon comparing
the crystal structures of GFPs and mFruits, structural differ-
ences in the beta barrels are observed. The tetramer subunit
interactions present in the naturally occurring red fluores-
cent protein DsRed are not present in the mFruit monomeric
forms and therefore the latter is expected to have less struc-
tural integrity. The crystal structures show larger openings
in the mFruits’ protein structure, which may be transiently
increased further by more pronounced thermal fluctuations.
These larger openings may allow oxygen to pass more eas-
ily to the chromophore which may have implications to both
chromophore maturation speed as well as photobleaching due
to oxidation.
Figure 1(a) is a superposition of ribbon structures of the
RFP mCherry (PDB code 2H5Q) and its GFP homologue cit-
rine (PDB code 1HUY). Figure 1(b) displays a space-filling
model of the β7-β10 region and shows that the gap between
β7 and β10 is smaller in citrine. We show that differences in
this region create pathways in mCherry for dioxygen diffusion
through the barrel to the chromophore.
II. METHODS
A. Molecular dynamics computations
The time series trajectories were obtained from explicit
solvent, all-atom simulations using the CHARMM27 force
field.43 The initial x-ray crystallographic structures of RFP
mCherry (pdb code 2H5Q) and YFP citrine (pdb code 1HUY)
have a few missing amino acid residues. The missing amino
acid residues were completed using MODELLER.44 For the
chromophore representation, we used pre-cyclized forms that
are composed of three amino acids each, Met66-Tyr67-Gly68
for mCherry and Gly65-Tyr66-Gly67 for citrine. Although
the fluorescent spectral properties are only possible in ma-
ture chromophores, the simpler representative forms of chro-
mophores can be used for the purpose of investigating the
barrel fluctuations.45
The MMTSB toolset46 was used to set up the system
for simulations. The initial structures of mCherry and cit-
rine were separately solvated in octahedral boxes under pe-
riodic boundary conditions with TIP3P water molecules with
a box cut-off of 10 Å. For mCherry, 11 319 water molecules
were used, and for citrine, 9290 water molecules were used.
All water molecules overlapping with the protein were re-
moved. The particle mesh Ewald method47 was used to treat
long range interactions with a 9-Å non-bonded cutoff. Energy
minimization was performed using the adopted basis
Newton–Raphson (ABNR) method.43 Each system was then
neutralized by adding sodium counter ions: six sodium ions
for RFP and 8 sodium ions for YFP. Water molecules that
overlapped with the sodium ions were removed and ABNR
energy minimization was performed again. The systems were
then heated with a linear gradient of 50 K/ps from 50 K to
300 K. At 300 K, the systems were equilibrated for 2 ns with
a 2 fs integration time step in the NVT (constant number, vol-
ume, and temperature) ensemble. The SHAKE algorithm48
was used to constrain the bonds connected with hydrogen
atoms. This was followed by a 10 ns NVT dynamics simu-
lation with 2 fs time steps for each protein that was used for
analysis.
B. Implicit ligand sampling for molecular oxygen
The ILS (Ref. 28) is a computational method, which
computes the potential of mean force (PMF) corresponding
to the placement of a given small ligand such as O2and CO,
everywhere inside the protein. The calculated PMF describes
the Gibb’s free energy cost of having a particle located at
a given position, integrated over all degrees of freedom of
the system, except ligand position. Calculated PMF values
also show the area accessible to the ligand with the associ-
ated free energy cost. We applied PMF/ILS calculations to
the frames from our MD simulations to determine locations
in a protein that are especially important for blocking, or fa-
cilitating oxygen passage, and to quantify the differences at
these locations between FP variants. A total of 5000 protein
conformations from a 10-ns MD trajectory were used for lig-
and sampling. Therefore, the free-energy value at each of the
locations is the average obtained from ILS performed every
2 ps for the 10-ns MD simulation trajectory. For the free en-
ergy calculation at each location, 20 different rotational ori-
entations of molecular dioxygen were sampled at each gird
position with a volume element size of 1 Å3. The free energy
is compared to a dioxygen molecule placed outside the protein
in the surrounding water, where the free energy is defined to
be zero. In the figures, all locations with a free energy below
235101-3 Oxygen diffusion pathways in mCherry J. Chem. Phys. 135, 235101 (2011)
2.0 kcal/mol are colored red, and all locations with a free
energy above +10.0 kcal/mol are colored blue. The values
of the free energy as a function of reaction coordinate were
calculated for specific positions separated by 1 Å distance
along the pathways, extending from outside the protein in
the solvent, into the protein and leading to the chromophore.
The pathways were determined from visual inspection as well
as from the 3D grid data of free-energy values from ILS
simulations.
III. RESULTS AND DISCUSSION
Results from 10 ns MD simulations comparing the
monomeric mCherry RFP to the citrine YFP are shown in
Fig. 2. Figure 2(a) displays the root-mean-square fluctua-
tions (RMSF) of each amino acid. The β7 strand in mCherry
(amino acids 141–151) displays a significantly larger RMSF
than β7 in citrine (amino acids 147–157), whereas for strand
β10 the RMSF is approximately the same for mCherry (193–
204) and citrine (199–208).
To further investigate the possibility of oxygen access in
this region, we next focus specifically on the opening between
strands β7 and β10, which is the region that is postulated to
allow oxygen entry for mCherry. Figure 2(b) is a time series of
the separation between strands β7 and β10. The plots repre-
sent the separation r between a pair of residues, one residue
on β7 and the other on β10. In order to characterize the size
of the gap, an atom is chosen on each residue that is clos-
FIG. 2. Results from 10 ns MD simulations comparing the citrine YFP to
the monomeric mCherry RFP. (a) Root-mean-square fluctuations (RMSF) of
each amino acid. (b) Time series of the separation between strands β7and
β10. Strands β7andβ10 show a large and fluctuating separation in mCherry
as compared to the small, fixed separation in citrine.
est to the other residue across the gap. For mCherry, ris
the distance between the Cαof residues Ala145 and Lys198,
and for citrine, r is the distance between the Cαof residues
Ser147 and Gln204. Not only is the β7-β10 gap always big-
ger in mCherry than citrine throughout the 10 ns, the gap in
mCherry also exhibits larger fluctuations.
A. Dioxygen access routes to the chromophore
in mCherry
The large and fluctuating gap between strands β7 and
β10 displayed in Fig. 2for mCherry but not for citrine, makes
this region a prime candidate for dioxygen access in mCherry.
To determine which regions or pathways allow the molecular
oxygen to enter the protein barrel, we calculated the free en-
ergy of placing molecular oxygen in and around the entire
protein barrel using ILS, which uses PMF calculations to de-
termine the free energy of placing a small molecule such as
dioxygen at a specific location in a protein. Figure 3(a) dis-
plays ensemble-averaged free-energy diagrams for a dioxy-
gen molecule if it is placed in, or around, mCherry and com-
pare that with citrine in Fig. 3(b). We calculated the free en-
ergy of the dioxygen using the ILS routine implemented in
the VMD molecular dynamics package.49 In Fig. 3(a),wedis-
play a slice that includes the β7 region. The color red repre-
sents low free-energy (<2 kcal/mole) locations, white rep-
resents intermediate free energy locations, and blue represents
high free-energy (>+10 kcal/mol) locations. In order for the
chromophore to have access to molecular oxygen, a pathway
without substantial free-energy barriers must exist from the
region outside the protein, through the protein barrel, to the
chromophore location in the interior. Figure 3(a) shows that
mCherry displays two low free-energy routes through the bar-
rel: one through the β7-β10 gap (R1) and the other through
the β7-β8 gap (R2). These two entry routes for dioxygen
lead all the way to the chromophore. In contrast, citrine has
no easy pathway, including the β7-β10 region (Y1) or the
FIG. 3. Free-energy isosurfaces for molecular dioxygen in (a) citrine and (b)
mCherry. The free-energy slice shown includes the β7-β10 region as well as
the β7-β8 region. The color red represents low free-energy (<2 kcal/mole)
locations, blue represents high free-energy (>+10 kcal/mol) locations, and
white represents locations for which the oxygen has intermediate free energy.
Neither the β7-β10 region nor the β7-β8 region in citrine offers low free-
energy routes for dioxygen entry, whereas in mCherry both regions display
gaps representing low free-energy access routes.
235101-4 Chapagain, Regmi, and Castillo J. Chem. Phys. 135, 235101 (2011)
FIG. 4. Free-energy values of dioxygen at locations along the pathways
showninFig.3leading from the solvent outside the protein (9 Å) into the
chromophore (0 Å). The mCherry has two easy routes that can be accessed
by entering through either the β7-β10gap(R1)ortheβ7-β8 gap (R2). The
routes for citrine through either the β7-β10 (Y1) region or β7-β8(Y2)region
are blocked by a high free-energy barrier.
β7-β8 region (Y2) both of which involve high free-energy
(blue) barriers.
Figure 4quantifies the value of the free energy along the
pathways shown in Fig. 3. The reaction coordinate is the po-
sition of the oxygen molecule along the route. The oxygen
molecule is placed at steps along the path that are separated
by 1 Å. The coordinate 0 represents a location near the chro-
mophore and 9 Å represents a location outside the protein in
the solvent. It is seen in this figure that both routes in citrine
(Y1, Y2) face large free energy barriers due to the protein
barrel, whereas there is no substantial barrier for either of the
pathways in mCherry (R1, R2). The identification of these
pathways is used later to guide mutations of key residues in
order to create barriers in mCherry to block these routes and
prevent dioxygen access to the chromophore.
B. Importance of sidechains in controlling gap size
As discussed in the Introduction, the β7-β10 gap and the
β7- β8 gap in mCherry are large enough to allow dioxygen to
pass through the protein barrel to the chromophore. An aim of
this work is to determine if site-specific amino acid mutations
can alter these routes. In order to ascertain more details of the
structural fluctuations in the barrel, we determined if the large
fluctuations in the mCherry β7-β10 gap is due to motion of
strand β7 or strand β10, and similarly for the β7-β8 gap.
In comparing the time series of the fluctuations in the
size of the mCherry β7-β10 gap (Fig. 2(b)) and the β7-β8
gap, we found that the openings and closings of the gaps were
out of phase with each other. When the β7-β10 gap is large,
the β7-β8 is small, and vice versa, implying that the fluctu-
ations in the β7-β10 gap and the β7-β8 gap are mostly due
to movement of β7. In addition, to provide more information
for guiding the mutations, we wish to determine why Figs. 3
and 4both show that the dioxygen pathway through the β7-
β8 gap (R2) is not as easy as the pathway through the β7-β10
gap (R1) even though the backbone separations are similar. In
Fig. 5we show that amino acid sidechains play an important
role in closing the β7-β8 gap. Figure 5displays the results
FIG. 5. Fluctuations in the β7-β8 gap in mCherry determined by the distance
between Cαatoms of β7-Ala145 and β8-Gln163 (darker line) compared to
the separation determined by the distance between Cαon β7-Ala145 and
the N on the sidechain of β8-Gln163 (lighter line). The sidechain of Gln163
narrows the gap significantly more than the backbone.
of 10 ns MD simulations for the separation between strands
β7 and β8 determined in two different ways. For both time
series, the separation is measured between Ala145 on β7 and
Gln163 on β8. One plot shows the time series of fluctuations
in the separation between the amino acids’ Cαatoms. The
other time series displays the fluctuating distance between the
Cαof Ala145 (on β7) and the N on the sidechain of Gln163
(on β8). Figure 5shows that the sidechain of Gln163 narrows
the gap significantly more than the backbone.
Figure 6further quantifies the importance of sidechains
for determining gap sizes. In Fig. 6, we present a contour
plot of the free energy of Ala145 on β7 and Gln163 on β8
as a function of their separation, measured in the same two
ways that are used in Fig. 5. The vertical axis is the sepa-
ration between the Cαof residue β7-Ala145 and the Cαof
residue β8-Gln163 (dark line in Fig. 5) and the horizontal
axis is the separation between the Cαof β7-Ala145 and the
N on the sidechain of residue β8-Gln163 (light line in Fig. 5).
Figure 6shows that there are two islands of low free energy,
which implies that β7 is stable at two distinct separations
from β8, with the larger separation meaning that β7iscloser
to β10. Another important feature is that when β7 is further
FIG. 6. Free-energy plot (in kcal/mol) of the β7-β8 strands. The horizon-
tal axis is the separation between the Cαon β7-Ala145 and the N on the
sidechain of β8-Gln163, the vertical axis is the separation between the Cαof
β7-Ala145 and the Cαof β8-Gln163. There are two distinct islands of low
free energy, showing that the β7 strand spends most time in these two distinct
positions. Additionally, when the Cα-Cαseparation is large, the sidechain un-
dergoes larger fluctuations, which restricts the gap from opening widely.
235101-5 Oxygen diffusion pathways in mCherry J. Chem. Phys. 135, 235101 (2011)
from β8, the range of fluctuations in the sidechain of β8is
also larger. This allows the large side chain of β8 to partially
close the gap even when β7 is far away.
C. Amino acid mutations in mCherry to control
dioxygen access to the chromophore
Figures 3and 4show that the easiest pathway for oxygen
access in mCherry is through the gap between β7 and β10,
and Figs. 5and 6show that sidechains play a role in clos-
ing the β7-β8 gap. Therefore, our aim in making amino acid
mutations is to decrease the β7-β10 gap without substantially
increasing the size of the gap between β7 and β8.
On comparing amino acid properties of the residues in
the β7, β8, and β10 strands of mCherry and citrine, it is seen
that there are more charged residues in mCherry as compared
to just two in β8 of citrine. The citrine residues in the region of
interest are mostly polar (Ser, Tyr, Asn, His) or hydrophobic
(Ala, Val, Phe, Ile, Leu). We first attempted a few mutations
in mCherry such that the charged amino acids are replaced
by polar or hydrophobic residues, similar to the pattern in cit-
rine. This change, however, either made the fluctuations worse
or the β7-β10 gap increased even further. Since the β7-β10
gap fluctuates the most, a possible strategy to reduce this fluc-
tuation is to create appropriate ionic interactions. In this re-
gion of mCherry, the inter-strand charged amino acid residues
participate in inter-strand ionic interactions and give rise to
salt-bridges. For example, the attractive interaction between
Glu144(-) in β7 and 164Arg(+)inβ8 swings β7 towards β8
and helps to open the gap between β7 and β10. To reduce
the barrel flexibility in this region, we made two amino acid
replacements (one in β7 and other in β8): we replaced the
polar 143Trp in β7 by a positively charged 143Lys(+), and
the 164Arg(+)inβ8 was replaced by 164Glu(-). The mu-
tations introduce two possible electrostatic interactions that
might close the β7-β10 gap. The attraction between the mu-
tated β7 143Lys(+) and β10 200ASP(-) pulls β7 towards
β10, and the repulsion between β7 Glu144(-) and the mu-
tated β8 164Glu(-) pushes β7 away from β8 and towards β10.
Since the location of the gap is close to the original tetramer-
breaking mutations, the barrel folding in monomeric form can
be sensitive to new mutations such as the one presented here.
A new set of mutations must still allow the barrel to fold and
chromophore to mature. If this is achieved, the marked reduc-
tion in the barrel fluctuations that limit the oxygen entry may
result in a more photostable FP.
The success of the mutations in closing the β7-β10 gap
in the mutated RFP (mut-RFP) is shown in Fig. 7.Thetwo
curves display time series for mCherry (same curve as in
Fig. 2(b)) and mut-RFP for the β7-β10 gap. The β7-β10 gap
in Fig. 7for the mut-RFP starts out at 7.5 Å, which is as
large as it ever gets for mCherry. This is expected because
the initial positions for the atoms in our proposed mut-RFP
are given by the PDB file for mCherry. Within a short time,
Fig. 7(a) shows that new interactions in mut-RFP greatly re-
duce the β7-β10 gap compared to mCherry which signifi-
cantly reduces the ease of oxygen permeability as displayed
by the free-energy isosurface in Fig. 7(b).
FIG. 7. (a) Results from the 10 ns MD simulation of the β7-β10 gap for
mCherry (red, same curve as in Fig. 2(b)) compared to its mutant (purple).
The β7-β10 gap in mut-RFP is greatly reduced compared to mCherry. (b)
Free-energy isosurface for molecular dioxygen in mut-RFP. As compared to
the isosurface displayed in Fig. 3(b) for mCherry, mut-RFP isosurface shows
significantly less favorable pathways for oxygen entry with high free-energy
barriers (blue).
We have used MD calculations to determine the pathways
for molecular oxygen entry through the barrel of mCherry. We
have shown that specific point mutations can alter the oxygen
pathways in the RFPs. Blocking or altering these pathways
through the barrel can have an effect on FP maturation as
well as on its photostability. For example, easy oxygen access
may significantly reduce the photostability whereas it may be
useful for chromophore maturation, especially at low oxygen
conditions. The computational approach can provide impor-
tant insights for guiding efficient mutagenesis experiments to
improve the maturation speed and photostability of mFruits.
ACKNOWLEDGMENTS
This work was supported in part by Award No.
SC3GM096903 from the National Institute of General Medi-
cal Sciences. The content of this publication is solely the re-
sponsibility of the authors and does not necessarily represent
the official views of the National Institute of General Medical
Sciences or the National Institutes of Health.
We thank Professor Ralph Jimenez at the University of
Colorado at Boulder for bringing this topic to our attention.
We also thank Professor Jimenez for critical reading of the
paper and many helpful discussions.
1R. Y. Tsien, Annu. Rev. Biochem. 67, 509 (1998).
2M. Zimmer, Chem. Rev. 102(3), 759 (2002).
3M. Chalfie, Y. Tu, G. Euskirchen, W. W. Ward, and D. C. Prasher, Science
263(5148), 802 (1994).
4A. Goulding, S. Shrestha, K. Dria, E. Hunt, and S. K. Deo, Protein Eng.
Des. Sel. 21(2), 101 (2008).
5M. V. Matz, A. F. Fradkov, Y. A. Labas, A. P. Savitsky, A. G. Zaraisky,
M. L. Markelov, and S. A. Lukyanov, Nat. Biotechnol. 17(10), 969 (1999).
235101-6 Chapagain, Regmi, and Castillo J. Chem. Phys. 135, 235101 (2011)
6D. Yarbrough, R. M. Wachter, K. Kallio, M. V. Matz, and S. J. Remington,
Proc. Natl. Acad. Sci. U.S.A. 98(2), 462 (2001).
7R. E. Campbell, O. Tour, A. E. Palmer, P. A. Steinbach, G. S. Baird,
D. A. Zacharias, and R. Y. Tsien, Proc. Natl. Acad. Sci. U.S.A. 99(12),
7877 (2002).
8G. S. Baird, D. A. Zacharias, and R. Y. Tsien, Proc. Natl. Acad. Sci. U.S.A.
97(22), 11984 (2000).
9E. M. Merzlyak, J. Goedhart, D. Shcherbo, M. E. Bulina, A. S. Shcheglov,
A. F. Fradkov, A. Gaintzeva, K. A. Lukyanov, S. Lukyanov, T. W. Gadella,
and D. M. Chudakov, Nat. Methods 4(7), 555 (2007).
10N. C. Shaner, R. E. Campbell, P. A. Steinbach, B. N. Giepmans,
A. E. Palmer, and R. Y. Tsien, Nat. Biotechnol. 22(12), 1567 (2004).
11N. C. Shaner, P. A. Steinbach, and R. Y. Tsien, Nat. Methods 2(12), 905
(2005).
12N. C. Shaner, M. Z. Lin, M. R. McKeown, P. A. Steinbach, K. L. Hazel-
wood, M. W. Davidson, and R. Y. Tsien, Nat. Methods 5(6), 545 (2008).
13P. Amara, P. Andreoletti, H. M. Jouve, and M. J. Field, Protein Sci. 10(10),
1927 (2001).
14M. L. Carlson, R. M. Regan, and Q. H. Gibson, Biochemistry 35(4), 1125
(1996).
15C. Bossa, M. Anselmi, D. Roccatano, A. Amadei, B. Vallone, M. Brunori,
and A. Di Nola, Biophys. J. 86(6), 3855 (2004).
16D. C. Lamb, A. Arcovito, K. Nienhaus, O. Minkow, F. Draghi, M. Brunori,
and G. U. Nienhaus, Biophys. Chem. 109(1), 41 (2004).
17Q. H. Gibson, R. Regan, R. Elber, J. S. Olson, and T. E. Carver, J. Biol.
Chem. 267(31), 22022 (1992).
18M. Brunori and Q. H. Gibson, EMBO Rep. 2(8), 674 (2001).
19J. Cohen, K. Kim, P. King, M. Seibert, and K. Schulten, Structure 13(9),
1321 (2005).
20V. A. Feher, E. Baldwin, and F. W. Dahlquist, Nat. Struct. Biol. 3, 516
(1996).
21J. R. Lakowicz and G. Weber, Biochemistry 12(21), 4171 (1973).
22W. Nadler and D. L. Stein, Proc. Natl. Acad. Sci. U.S.A. 88(15), 6750
(1991).
23D. B. Calhoun, J. M. Vanderkooi, G. V. Woodrow III, and S. W. Englander,
Biochemistry 22(7), 1526 (1983).
24B. R. Branchini, A. R. Nemser, and M. Zimmer, J. Am. Chem. Soc. 120(1),
1 (1998).
25X. Shu, N. C. Shaner, C. A. Yarbrough, R. Y. Tsien, and S. J. Remington,
Biochemistry 45(32), 9639 (2006).
26R. G. Justin, B. Rosemary, and S. Klaus, J. Comput. Phys. 151(1), 190
(1999).
27B. Roux, Comp. Phys. Commun. 91(1–3), 275 (1995).
28J. Cohen, A. Arkhipov, R. Braun, and K. Schulten, Biophys. J. 91(5), 1844
(2006).
29Y. Wang, J. Cohen, W. F. Boron, K. Schulten, and E. Tajkhorshid,
J. Struct. Biol. 157(3), 534 (2007); B. J. Johnson, J. Cohen, R. W. Welford,
A. R. Pearson, K. Schulten, J. P. Klinman, and C. M. Wilmot, J. Biol. Chem.
282(24), 17767 (2007).
30J. Saam, I. Ivanov, M. Walther, H. G. Holzhutter, and H. Kuhn, Proc. Natl.
Acad. Sci. U.S.A. 104(33), 13319 (2007).
31J. Saam, E. Rosini, G. Molla, K. Schulten, L. Pollegioni, and S. Ghisla,
J. Biol. Chem. 285(32), 24439 (2010).
32R. Baron, C. Riley, P. Chenprakhon, K. Thotsaporn, R. T. Winter,
A. Alfieri, F. Forneris, W. J. van Berkel, P. Chaiyen, M. W. Fraaije,
A. Mattevi, and J. A. McCammon, Proc. Natl. Acad. Sci. U.S.A. 106(26),
10603 (2009).
33Y. Fan, L. Lund, Q. Shao, Y. Q. Gao, and F. M. Raushel, J. Am. Chem. Soc.
131(29), 10211 (2009).
34L. Lund, Y. Fan, Q. Shao, Y. Q. Gao, and F. M. Raushel, J. Am. Chem. Soc.
132(11), 3870 (2010).
35X. S. Wang, A. E. Roitberg, and N. G. Richards, Biochemistry 48(51),
12272 (2009).
36R. Daigle, M. Guertin, and P. Lague, Proteins 75(3), 735 (2009).
37O. Griesbeck, G. S. Baird, R. E. Campbell, D. A. Zacharias, and
R. Y. Tsien, J. Biol. Chem. 276(31), 29188 (2001).
38A. Miyawaki, T. Nagai, and H. Mizuno, Adv. Biochem. Eng. Biotechnol.
95, 1 (2005).
39R. M. Wachter, D. Yarbrough, K. Kallio, and S. J. Remington, J. Mol. Biol.
301(1), 157 (2000).
40A. Rekas, J. R. Alattia, T. Nagai, A. Miyawaki, and M. Ikura, J. Biol. Chem.
277(52), 50573 (2002).
41T. D. Craggs, Chem. Soc. Rev. 38(10), 2865 (2009).
42A. G. Evdokimov, M. E. Pokross, N. S. Egorov, A. G. Zaraisky, I. V. Yam-
polsky, E. M. Merzlyak, A. N. Shkoporov, I. Sander, K. A. Lukyanov, and
D. M. Chudakov, EMBO Rep. 7(10), 1006 (2006).
43B. R. Brooks, R. E. Bruccoleri, B. D. Olafson, D. J. States, S. Swaminathan,
and M. Karplus, J. Comp. Chem. 4(2), 187 (1983).
44N. Eswar, B. Webb, M. A. Marti-Renom, M. S. Madhusudhan, D. Eramian,
M. Y. Shen, U. Pieper, and A. Sali, Curr. Protoc. Bioinf. 15, 5.6-1 (2006).
45B. T. Andrews, S. Gosavi, J. M. Finke, J. N. Onuchic, and P. A. Jennings,
Proc. Natl. Acad. Sci. U.S.A. 105(34), 12283 (2008).
46M. Feig, J. Karanicolas, and C. L. Brooks 3rd, J. Mol. Graphics Modell.
22(5), 377 (2004).
47E. Ulrich, P. Lalith, L. B. Max, D. Tom, L. Hsing, and G. P. Lee, J. Chem.
Phys. 103(19), 8577 (1995).
48W. F. van Gunsteren and H. J. C. Berendsen, Mol. Phys. 34(5), 1311
(1977).
49W. Humphrey, A. Dalke, and K. Schulten, J. Mol. Graphics 14(1), 33
(1996).
... Additionally, owing to their monomeric nature, they can be genetically fused with other proteins without undergoing dimerization or tetramerization, which could affect their function. 54,55 The structure of known FPs consists of a β-barrel fold encompassing a chromophore-containing α-helix 56 (Figure 1a). The α-helix is threaded through the center of the barrel, 56 and additional helices and coils cross over the top and bottom of the cylindrical barrel. ...
... This gap is a vestige of the missing tetrameric interactions disrupted by mutations introduced to design a monomeric FP. 85 Previous studies have shown that the β7-β10 gap undergoes structural transitions between closed and open states, 61 allowing molecular oxygen to diffuse into the barrel and affect the photostability of the chromophore. 55,85 To determine the width of the β7-β10 gap, we measured the distance between three residues on β7 (E144, A145, S146) and three residues on β10 (I197, K198, L199). Consistent with previous studies, 55,61 we observe that the β7-β10 gap undergoes transitions between two highly populated states, as seen in Figure 3b. ...
... 55,85 To determine the width of the β7-β10 gap, we measured the distance between three residues on β7 (E144, A145, S146) and three residues on β10 (I197, K198, L199). Consistent with previous studies, 55,61 we observe that the β7-β10 gap undergoes transitions between two highly populated states, as seen in Figure 3b. The state where the gap distance is small corresponds to the closed state, and the state where the gap distance is large corresponds to the open state. ...
Preprint
Full-text available
The dense cellular environment influences bio-macromolecular structure, dynamics, interactions and function. Despite advancements in understanding protein-crowder interactions , predicting their precise effects on protein structure and function remains challenging. Here, we elucidate the effects of PEG-induced crowding on the fluorescent protein mCherry using molecular dynamics simulations and fluorescence-based experiments. We identify and characterize specific PEG-induced structural and dynami-cal changes in mCherry. Importantly, we find interactions in which PEG molecules wrap around specific surface-exposed residues in a binding mode previously observed 1. CC-BY-NC-ND 4.0 International license available under a was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made The copyright holder for this preprint (which this version posted May 7, 2024. ; https://doi.org/10.1101/2024.05.07.592799 doi: bioRxiv preprint in protein crystal structures. Fluorescence correlation spectroscopy experiments capture PEG-induced changes, including aggregation, suggesting a potential role for the specific PEG-mCherry interactions identified in simulations. Additionally, mCherry fluorescence lifetimes are influenced by PEG and not by the bulkier crowder dextran, highlighting the importance of crowder-protein soft interactions. This work augments our understanding of macromolecular crowding effects on protein structure and dynamics .
... 48 To facilitate comparisons with published results, force field parameters for the deprotonated, anionic form of the mature chromophore in the ground state from previous work were used. 49,50 The protonation state of titratable residues at pH ∼ 7.4 (as for in vitro measurements) was based on previous MD studies of mCherry variants detailed in refs 49 and 50. Based on comparable pK a values and sensitivity to basic pH, it was assumed the E215 residue was protonated and was modified using a patch for all RFPs in this study. ...
... Previous investigations of the interstrand dynamics of β7, β8, and β10 in mCherry revealed nanosecond time scale fluctuations of the gap between strands β7 and β10 as quantified by the distance between the α-C atoms of A145 and K198 (Figures 2 and S4.1). 49,50 Our current findings (Figures 2 and S4.1) are consistent with our previous results. This gap was identified as an entry point for O 2 into the barrel, which can subsequently lead to chromophore photodegradation. ...
... This gap was identified as an entry point for O 2 into the barrel, which can subsequently lead to chromophore photodegradation. 49,50 Kriek (mCherry W143I, I161M, and Q163V) a highly photostable mCherry The Journal of Physical Chemistry B pubs.acs.org/JPCB Article variant was rationally designed by targeting residues that could effectively reduce these fluctuations. ...
Article
The 3-fold higher brightness of the recently developed mCherry-XL red fluorescent protein (FP) compared to its progenitor, mCherry, is due to a significant decrease in the nonradiative decay rate underlying its increased fluorescence quantum yield. To examine the structural and dynamic role of the four mutations that distinguish the two FPs and closely related variants, we employed microsecond time scale, all-atom molecular dynamics simulations. The simulations revealed that the I197R mutation leads to the formation of multiple hydrogen-bonded contacts and increased rigidity of the β-barrel. In particular, mCherryXL showed reduced nanosecond time scale breathing of the gap between the β7 and β10-strands, which was previously shown to be the most flexible region of mCherry. Together with experimental results, the simulations also reveal steric interactions of residue 161 and a network of hydrogen-bonding interactions of the chromophore with residues at positions 59, 143, and 163 that are critical in perturbing the chromophore electronic structure. Finally, we shed light on the conformational dynamics of the conserved residues R95 and S146, which are hydrogen-bonded to the chromophore, and provide physical insights into the observed photophysics. To the best of our knowledge, this is the first study that evaluates the conformational space for a set of closely related FPs generated by directed evolution.
... Molecular dynamics (MD) computer simulations can be used to disclose oxygen d fusion pathways in proteins [10][11][12][13][14][15] and to obtain starting structures for prospective qua tum-based calculations of reactions of proteins with molecular oxygen [8,9,[15][16][17][18]. T present work aims at predicting oxygen-binding sites in the miniSOG variants using M simulations with force-field interaction potentials and with quantum mechanics/molec lar mechanics (QM/MM) potentials. ...
... Molecular dynamics (MD) computer simulations can be used to disclose oxygen diffusion pathways in proteins [10][11][12][13][14][15] and to obtain starting structures for prospective quantum-based calculations of reactions of proteins with molecular oxygen [8,9,[15][16][17][18]. The present work aims at predicting oxygen-binding sites in the miniSOG variants using MD simulations with force-field interaction potentials and with quantum mechanics/molecular mechanics (QM/MM) potentials. ...
Article
Full-text available
Interaction of molecular oxygen 3O2 with the flavin-dependent protein miniSOG after light illumination results in creation of singlet oxygen 1O2 and superoxide O2●−. Despite the recently resolved crystal structures of miniSOG variants, oxygen-binding sites near the flavin chromophore are poorly characterized. We report the results of computational studies of the protein−oxygen systems using molecular dynamics (MD) simulations with force-field interaction potentials and quantum mechanics/molecular mechanics (QM/MM) potentials for the original miniSOG and the mutated protein. We found several oxygen-binding pockets and pointed out possible tunnels bridging the bulk solvent and the isoalloxazine ring of the chromophore. These findings provide an essential step toward understanding photophysical properties of miniSOG—an important singlet oxygen photosensitizer.
... It is well established that the rigidity of the β-barrel structure of fluorescent proteins shields the chromophore from the solvent and quenchers. In this way, a favorable environment for fluorescence emission is provided [42,43]. Disruption of barrel integrity leads in general to enhanced contact of solvent molecules to the chromophore, which results in enhanced fluorescence quenching. ...
... Both processes compete with the radiative deactivation channel and reduce the fluorescence brightness. A well-investigated example is the introduction of an aperture in the barrel, as in the H148G GFP mutant, which strongly affects the protonation state [42,44] but also the oxygen accessibility [43,45] of the chromophore. We propose that the enlargement of the cysteine side chains upon S-nitrosylation (Scheme 1) leads to decreasing rigidity of the βbarrel and a subsequently increased fluorescence quenching. ...
Article
Full-text available
S-Nitrosylation of cysteine residues is an important molecular mechanism for dynamic, post-translational regulation of several proteins, providing a ubiquitous redox regulation. Cys residues are present in several fluorescent proteins (FP), including members of the family of Aequorea victoria Green Fluorescent Protein (GFP)-derived FPs, where two highly conserved cysteine residues contribute to a favorable environment for the autocatalytic chromophore formation reaction. The effect of nitric oxide on the fluorescence properties of FPs has not been investigated thus far, despite the tremendous role FPs have played for 25 years as tools in cell biology. We have examined the response to nitric oxide of fluorescence emission by the blue-emitting fluorescent protein mTagBFP2. To our surprise, upon exposure to micromolar concentrations of nitric oxide, we observed a roughly 30% reduction in fluorescence quantum yield and lifetime. Recovery of fluorescence emission is observed after treatment with Na-dithionite. Experiments on related fluorescent proteins from different families show similar nitric oxide sensitivity of their fluorescence. We correlate the effect with S-nitrosylation of Cys residues. Mutation of Cys residues in mTagBFP2 removes its nitric oxide sensitivity. Similarly, fluorescent proteins devoid of Cys residues are insensitive to nitric oxide. We finally show that mTagBFP2 can sense exogenously generated nitric oxide when expressed in a living mammalian cell. We propose mTagBFP2 as the starting point for a new class of genetically encoded nitric oxide sensors based on fluorescence lifetime imaging.
... Using real-time in vivo fluorescence imaging combined with antimicrobial activity, we determined the impact of promoter arrangement and culture conditions (including choice of growth media and inducer agents) on the heterologous expression of nisin, clausin and epidermin. The use of mCherry as a fusion partner provides several advantages beneficial for lanthipeptide production and bioprocess engineering [47,48]. Of significance is enhanced lanthipeptide solubility in the cytosol, reduced potential toxic effects to the producing host and the ability to track the tuned expression of fused lanthipeptides in real-time and in vivo prior to purification and characterization steps. ...
Article
Full-text available
Background Lanthipeptides are a rapidly expanding family of ribosomally synthesized and post-translationally modified natural compounds with diverse biological functions. Lanthipeptide structural and biosynthetic genes can readily be identified in genomic datasets, which provides a substantial repository for unique peptides with a wide range of potentially novel bioactivities. To realize this potential efficiently optimized heterologous production systems are required. However, only a few class I lanthipeptides have been successfully expressed using Escherichia coli as heterologous producer. This may be attributed to difficulties experienced in the co-expression of structural genes and multiple processing genes as well as complex optimization experiments. Results Here, an optimized modular plasmid system is presented for the complete biosynthesis for each of the class I lanthipeptides nisin and clausin, in E. coli. Genes encoding precursor lanthipeptides were fused to the gene encoding the mCherry red fluorescent protein and co-expressed along with the required synthetases from the respective operons. Antimicrobially active nisin and clausin were proteolytically liberated from the expressed mCherry fusions. The mCherry-NisA expression system combined with in vivo fluorescence monitoring was used to elucidate the effect of culture media composition, promoter arrangement, and culture conditions including choice of growth media and inducer agents on the heterologous expression of the class I lanthipeptides. To evaluate the promiscuity of the clausin biosynthetic enzymes, the optimized clausin expression system was used for the heterologous expression of epidermin. Conclusion We succeeded in developing novel mCherry-fusion based plug and play heterologous expression systems to produce two different subgroups of class I lanthipeptides. Fully modified Pre-NisA, Pre-ClausA and Pre-EpiA fused to the mCherry fluorescence gene was purified from the Gram-negative host E. coli BL21 (DE3). Our study demonstrates the potential of using in vivo fluorescence as a platform to evaluate the expression of mCherry-fused lanthipeptides in E. coli. This allowed a substantial reduction in optimization time, since expression could be monitored in real-time, without the need for extensive and laborious purification steps or the use of in vitro activity assays. The optimized heterologous expression systems developed in this study may be employed in future studies for the scalable expression of novel NisA derivatives, or novel genome mined derivatives of ClausA and other class I lanthipeptides in E. coli.
... [29][30][31][32] Several novel mutant monomeric variants known as mFruits have been designed to address this issue; among the designed variants, mCherry, mOrange, and mStrawberry are the most promising. 28,33,34 Compared to its progenitor, mCherry has the advantages of a favorable red-shift of absorption and emission spectra, higher expression and fast chromophore maturation, and lower phototoxicity. 35 Unfortunately, the fluorescence quantum yield of mcherry is only one-third, rendering it suubstantially dimmer than DsRed. ...
Preprint
Full-text available
In this study, we explore the electron transition mechanism and optical properties of the popular red fluorescent protein mCherry. By examining the charge transfer spectrum and combining it with the mCherry hole-electron distribution, we identify that the charge transfer between the phenolate and imidazolinone loops significantly contributes to the absorption spectrum. Quantitative analysis of charge transfer shows that, overall, the electrons are transferred to the C16 atom in the middle of phenolate and the imidazolinone loops during absorption. We speculate that C16 may also absorb protons to enable the photoconversion of mCherry in the excited state, similar to the blinking mechanism of IrisFP. In addition, we further investigated the optical properties of mcherry in the external field by polarizability (hyperpolarizability), showing the anisotropy of the polarization, the first hyperpolarization and the second hyperpolarization by unit spherical representation. Our results suggest that significant polarization and second hyperpolarizability occur when the field direction and electron transfer direction are aligned. We also analyzed the polarizability and first hyperpolarizabilities for different external fields. The polarizability mutated when the external field satisfies the S_0,min-> S_1 transition. Finally, the study of the first hyperpolarizability shows that adjusting the appropriate field can lead to a linear photoelectric effect or second harmonic generation of mCherry. These studies have certain reference values for various red fluorescent protein correlation simulations and experiments because of the similarity of the red fluorescent protein.
... Since strong fluorescence was observed on top of the construct opposite to the LB agar, RFP heterogeneity is not caused by nutrient limitation preventing growth or gene expression. Rather, oxygen limitation is likely to occur at the center of the structure, reducing bacterial proliferation and RFP maturation [42]. ...
Article
Full-text available
The intertwined adoption of synthetic biology and 3D bioprinting has the potential to improve different application fields by fabricating engineered living materials (ELMs) with unnatural genetically-encoded sense & response capabilities. However, efforts are still needed to streamline the fabrication of sensing ELMs compatible with field use and improving their functional complexity. To investigate these two unmet needs, we adopted a workflow to reproducibly construct bacterial ELMs with synthetic biosensing circuits that provide red pigmentation as visible readout in response to different proof-of-concept chemical inducers. We first fabricated single-input/single-output ELMs and we demonstrated their robust performance in terms of longevity (cell viability and evolutionary stability >15 days, and long-term storage >1 month), sensing in harsh, non-sterile or nutrient-free conditions compatible with field use (soil, water, and clinical samples, including real samples from Pseudomonas aeruginosa infected patients). Then, we fabricated ELMs including multiple spatially-separated biosensor strains to engineer: level-bar materials detecting molecule concentration ranges, multi-input/multi-output devices with multiplexed sensing and information processing capabilities, and materials with cell-cell communication enabling on-demand pattern formation. Overall, we showed successful field use and multiplexed functioning of reproducibly fabricated ELMs, paving the way to a future automation of the prototyping process and boost applications of such devices as in-situ monitoring tools or easy-to-use sensing kits.
... Cells chronically exposed to hyperoxic O 2 levels (18 kPa O 2 ) showed a robust geNOps expression, yet the biosensor lacked sensitivity for exogenous NO administration (Fig. 1). Cells adapted to physiological normoxia (5 kPa O 2 ) displayed lower basal fluorescence in both HEK293T and EA.hy926 cells, as expected due to the requirement for O 2 in the maturation of the fluorophore [47]. A recent study investigated chromophore maturation under certain oxygen conditions including 9, 12, 15, and 21 kPa O 2 with differently colored purified pre-mature FPs. ...
Article
Full-text available
Lighting systems with circularly polarized luminescence (CPL) are an emerging field with high hopes in, for example, neural cell circuits and encoding applications. The major challenges that forfeits their real‐world application are i) the design of chiroptical materials (CMs) with high CPL brightness (BCPL; today's record is Eu‐based compounds with average 287 M⁻¹cm⁻¹, while 90% of other CMs show <150 M⁻¹cm⁻¹ in solution) and ii) how to keep CPL response in films/coatings of technological relevance. Since natural evolution is driven by chiral selectivity at the supramolecular level, fluorescent proteins (FPs) are ideal candidates to provide large BCPL spanning visible and near‐infrared regions. This hypothesis is confirmed for all the known FP classes, demonstrating high emission intensities (photoluminescence quantum yields (ϕ) up to 76%) and record average BCPL of |200| M⁻¹cm⁻¹ (solution). What is more, the CPL response is also kept in polymer coatings. It is rationalized that structural factors (chromophore rigidity, surrounding amino acids, supramolecular packaging, and exciton coupling) hold a significant influence, regardless of the ϕ values. Finally, proof‐of‐concept CPL‐encoded signals in monochromatic/white hybrid light‐emitting diodes with FP‐polymer filters show exceptional stabilities. Overall, this work stands out FPs toward a new CM family, in general, and biogenic CPL‐encoded lighting systems, in particular.
Article
Full-text available
Iron is an essential metal for cellular metabolism and signaling, but it has adverse effects in excess. The physiological consequences of iron deficiency are well established, yet the relationship between iron supplementation and pericellular oxygen levels in cultured cells and their downstream effects on metalloproteins has been less explored. This study exploits the metalloprotein geNOps in cultured HEK293T epithelial and EA.hy926 endothelial cells to test the iron-dependency in cells adapted to standard room air (18 kPa O2) or physiological normoxia (5 kPa O2). We show that cells in culture require iron supplementation to activate the metalloprotein geNOps and demonstrate for the first time that cells adapted to physiological normoxia require significantly lower iron compared to cells adapted to hyperoxia. This study establishes an essential role for recapitulating oxygen levels in vivo and uncovers a previously unrecognized requirement for ferrous iron supplementation under standard cell culture conditions to achieve geNOps functionality.
Article
Full-text available
The previously developed particle mesh Ewald method is reformulated in terms of efficient B‐spline interpolation of the structure factors. This reformulation allows a natural extension of the method to potentials of the form 1/r p with p≥1. Furthermore, efficient calculation of the virial tensor follows. Use of B‐splines in place of Lagrange interpolation leads to analytic gradients as well as a significant improvement in the accuracy. We demonstrate that arbitrary accuracy can be achieved, independent of system size N, at a cost that scales as N log(N). For biomolecular systems with many thousands of atoms this method permits the use of Ewald summation at a computational cost comparable to that of a simple truncation method of 10 Å or less.
Article
Full-text available
Since the cloning of Aequorea victoria green fluorescent protein (GFP) in 1992, a family of known GFP-like proteins has been growing rapidly. Today, it includes more than a hundred proteins with different spectral characteristics cloned from Cnidaria species. For some of these proteins, crystal structures have been solved, showing diversity in chromophore modifications and conformational states. However, we are still far from a complete understanding of the origin, functions and evolution of the GFP family. Novel proteins of the family were recently cloned from evolutionarily distant marine Copepoda species, phylum Arthropoda, demonstrating an extremely rapid generation of fluorescent signal. Here, we have generated a non-aggregating mutant of Copepoda fluorescent protein and solved its high-resolution crystal structure. It was found that the protein beta-barrel contains a pore, leading to the chromophore. Using site-directed mutagenesis, we showed that this feature is critical for the fast maturation of the chromophore.
Article
Full-text available
Molecular dynamics simulations and implicit ligand sampling methods have identified trajectories and sites of high affinity for O2 in the protein framework of the flavoprotein d-amino-acid oxidase (DAAO). A specific dynamic channel for the diffusion of O2 leads from solvent to the flavin Si-side (amino acid substrate and product bind on the Re-side). Based on this, amino acids that flank the putative O2 high affinity sites have been exchanged with bulky residues to introduce steric constraints. In G52V DAAO, the valine side chain occupies the site that in wild-type DAAO has the highest O2 affinity. In this variant, the reactivity of the reduced enzyme with O2 is decreased ≥100-fold and the turnover number ≈1000-fold thus verifying the concept. In addition, the simulations have identified a chain of three water molecules that might serve in relaying a H+ from the product imino acid =NH2+ group bound on the flavin Re-side to the developing peroxide on the Si-side. This function would be comparable with that of a similarly located histidine in the flavoprotein glucose oxidase.
Article
The application of the computer simulation method of molecular dynamics to macromolecules is investigated. The protein trypsin inhibitor (BPTI), consisting of 454 united atoms, is used as an example. Different algorithms for integrating the equations of motion are compared, both theoretically and in practice. It is examined to what extent the chain structure of a macromolecule allows a reduction of the computational effort by the introduction of constraints in the dynamics of the chain.A calculational scheme is proposed, by which constraints can be incorporated in predictor-corrector algorithms. The optimum choice of an algorithm depends on the desired accuracy of the solution and on the character of the forces acting on the molecule, viz. whether these are noisy or not. For nonconstraint dynamics a Gear predictor-corrector algorithm yields the best results, whereas for constraint dynamics the Gear and Verlet algorithms produce comparable results. The application of bond-length constraints reduces the required computer time by a factor of 3. The inclusion of bondangle constraints is not recommended.
Article
A thorough conformational search of the chromophore-forming region of immature green fluorescent protein (GFP) revealed that it is preorganized in a unique conformation required for chromophore formation. This "tight turn" conformation has an i carbonyl carbon to i + 2 amide nitrogen distance of less than 2.90 Angstrom with phi = 60 +/- 30 degrees and psi = 30 +/- 15 degrees. Less than 1.00% of the residues of the 50 representative proteins examined adopt this conformation. The tight turn conformation is predominately located on the periphery of the proteins or in flexible areas, except in GFP. Molecular dynamics simulations and Ramachandran plots show the chromophore-forming region in immature GFP can only adopt the tight turn conformational family. Moreover, this conformation is ideally suited for the cyclization necessary for chromophore formation, i.e., for nucleophilic attack of the amino group of Gly67 on the carbonyl group of Ser67. The 11 beta sheets of GFP force the chromophore-forming peptide fragment to adopt a conformation that has an exceptionally short interatomic distance between the carbonyl carbon of Ser65 and the amide nitrogen of Gly67 and lock it into this conformation. Several mutant GFPs have been expressed that exhibit greater solubility and thermostability than wild-type GFP. These properties are linked to protein folding and chromophore formation. Our calculations show that the mutations cause a shortening of the distance between the carbonyl carbon of Ser65 and the amide nitrogen df Gly67, and therefore enhance chromophore formation.
Article
CHARMM (Chemistry at HARvard Macromolecular Mechanics) is a highly flexible computer program which uses empirical energy functions to model macromolecular systems. The program can read or model build structures, energy minimize them by first- or second-derivative techniques, perform a normal mode or molecular dynamics simulation, and analyze the structural, equilibrium, and dynamic properties determined in these calculations. The operations that CHARMM can perform are described, and some implementation details are given. A set of parameters for the empirical energy function and a sample run are included.
Article
Atomic force microscopy (AFM) experiments and steered molecular dynamics (SMD) simulations have revealed much about the dynamics of protein-ligand binding and unbinding, as well as the stretching and unfolding of proteins. Both techniques induce ligand unbinding or protein unfolding by applying external mechanical forces to the ligand or stretched protein. However, comparing results from these two techniques, such as the magnitude of forces required to unbind ligands, has remained a challenge since SMD simulations proceed six to nine orders of magnitude faster due to limitations in computational resources. Results of simulations and experiments can be compared through a potential of mean force (PMF). We describe and implement three time series analysis techniques for reconstructing the PMF from position and applied force data gathered from SMD trajectories. One technique, based on the WHAM theory, views the unbinding or stretching as a quasi-equilibrium process; the other two techniques, one based on van Kampen's Ω-expansion, the second on a least squares minimization of the Onsager–Machlup action with respect to the choice of PMF, assume a Langevin description of the dynamics in order to account for the nonequilibrium character of SMD data. The latter two methods are applied to SMD data taken from a simulation of the extraction of a lipid from a phospholipid membrane monolayer.
Article
The problem of unbiasing and combining the results of umbrella sampling calculations is reviewed. The weighted histogram analysis method (WHAM) of S. Kumar et al. (J. Comp. Chem. 13 (1992) 1011) is described and compared with other approaches. The method is illustrated with molecular dynamics simulations of the alanine dipeptide for one-and two-dimensional free energy surfaces. The results show that the WHAM approach simplifies considerably the task of recombining the various windows in complex systems.
Article
The transport of carbamate through the large subunit of carbamoyl phosphate synthetase (CPS) from Escherichia coli was investigated by molecular dynamics and site-directed mutagenesis. Carbamate, the product of the reaction involving ATP, bicarbonate, and ammonia, must be delivered from the site of formation to the site of utilization by traveling nearly 40 A within the enzyme. Potentials of mean force (PMF) calculations along the entire tunnel for the translocation of carbamate indicate that the tunnel is composed of three continuous water pockets and two narrow connecting parts, near Ala-23 and Gly-575. The two narrow parts render two free energy barriers of 6.7 and 8.4 kcal/mol, respectively. Three water pockets were filled with about 21, 9, and 9 waters, respectively, and the corresponding relative free energies of carbamate residing in these free energy minima are 5.8, 0, and 1.6 kcal/mol, respectively. The release of phosphate into solution at the site for the formation of carbamate allows the side chain of Arg-306 to rotate toward Glu-25, Glu-383, and Glu-604. This rotation is virtually prohibited by a barrier of at least 23 kcal/mol when phosphate remains bound. This conformational change not only opens the entrance of the tunnel but also shields the charge-charge repulsion from the three glutamate residues when carbamate passes through the tunnel. Two mutants, A23F and G575F, were designed to block the migration of carbamate through the narrowest parts of the carbamate tunnel. The mutants retained only 1.7% and 3.8% of the catalytic activity for the synthesis of carbamoyl phosphate relative to the wild type CPS, respectively.
Article
Glutamine 5'-phosphoribosylpyrophosphate amidotransferase (GPATase) catalyzes the synthesis of 5'-phosphoribosylamine in a reaction that involves the translocation of ammonia along an intramolecular tunnel linking the two active sites of the enzyme. We now report a locally enhanced sampling (LES) strategy for modeling ammonia transfer between the active sites of Escherichia coli GPATase in its active conformation. Our calculations demonstrate that the ammonia channel in GPATase is best regarded as a "pipe" through which ammonia travels in the absence of an external "driving" potential. This combined LES/PMF computational approach, which offers a straightforward alternative to steered molecular dynamics simulations in studies of substrate channeling, also provides new insights into the molecular basis of the reduced ammonia transfer efficiency exhibited by the L415A GPATase mutant.