ArticlePDF Available

Multiple active site residues are important for photochemical efficiency in the light-activated enzyme protochlorophyllide oxidoreductase (POR)

Authors:

Abstract and Figures

Protochlorophyllide oxidoreductase (POR) catalyzes the light-driven reduction of protochlorophyllide (Pchlide), an essential, regulatory step in chlorophyll biosynthesis. The unique requirement of the enzyme for light has provided the opportunity to investigate how light energy can be harnessed to power biological catalysis and enzyme dynamics. Excited state interactions between the Pchlide molecule and the protein are known to drive the subsequent reaction chemistry. However, the structural features of POR and active site residues that are important for photochemistry and catalysis are currently unknown, because there is no crystal structure for POR. Here, we have used static and time-resolved spectroscopic measurements of a number of active site variants to study the role of a number of residues, which are located in the proposed NADPH/Pchlide binding site based on previous homology models, in the reaction mechanism of POR. Our findings, which are interpreted in the context of a new improved structural model, have identified several residues that are predicted to interact with the coenzyme or substrate. Several of the POR variants have a profound effect on the photochemistry, suggesting that multiple residues are important in stabilizing the excited state required for catalysis. Our work offers insight into how the POR active site geometry is finely tuned by multiple active site residues to support enzyme-mediated photochemistry and reduction of Pchlide, both of which are crucial to the existence of life on Earth.
Content may be subject to copyright.
Multiple active site residues are important for photochemical
efciency in the light-activated enzyme protochlorophyllide
oxidoreductase (POR)
Binuraj R.K. Menon, Samantha J.O. Hardman, Nigel S. Scrutton ,DerrenJ.Heyes
a
Centre for Synthetic Biology of Fine and Speciality Chemicals, Manchester Institute of Biotechnology, School of Chemistry, The University of Manchester, Manchester, M1 7DN, UK
abstractarticle info
Article history:
Received 6 April 2016
Received in revised form 17 May 2016
Accepted 30 May 2016
Available online 01 June2016
Protochlorophyllide oxidoreductase (POR) catalyzes the light-driven reduction of protochlorophyllide (Pchlide),
an essential, regulatory step in chlorophyll biosynthesis.The unique requirementof the enzyme for light has pro-
vided the opportunityto investigate how light energy can be harnessed to power biologicalcatalysis and enzyme
dynamics. Excited state interactions between the Pchlide molecule and the protein are known to drive the sub-
sequent reaction chemistry. However, the structural features of POR and active site residues that are important
for photochemistry and catalysis are currently unknown, because there is no crystal structure for POR. Here,
we have used static and time-resolved spectroscopic measurements of a number of active site variants to
study the role of a number of residues, which are located in the proposed NADPH/Pchlide binding site based
on previous homology models, in thereaction mechanism of POR. Our ndings, which are interpreted in the con-
text of a new improved structural model, have identied several residues that are predicted to interact with the
coenzymeor substrate. Severalof the POR variants have a profound effecton the photochemistry, suggesting that
multiple residues are important in stabilizing the excited state required for catalysis. Our work offers insight into
how the POR active site geometry is nely tuned by multiple active site residues to support enzyme-mediated
photochemistry and reduction of Pchlide, both of which are crucial to the existence of life on Earth.
© 2016 The Authors. Published by Elsevier B.V. Thisis an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
Keywords:
Protochlorophyllide oxidoreductase (POR)
Light activation
Photochemistry
Site-directed mutagenesis
Enzyme catalysis
1. Introduction
The reduction of protochlorophyllide (Pchlide) to chlorophyllide,
catalyzed by the enzyme protochlorophyllide oxidoreductase (POR), is
the key light-driven reaction that leads to a profound transformation
in plant development [15]. In non-owering plants, algae and
cyanobacteria there is also a light-independent Pchlide reductase,
consisting of 3 separate subunits, whichcan catalyzethis same penulti-
mate step in the chlorophyll biosynthetic pathway [6,7]. In light-depen-
dent POR the reaction proceeds via a highly endergonic light-driven
hydride transfer from the NADPH coenzyme to the C17 position of the
Pchlide molecule followed by an exergonic thermally activated proton
transfer from a conserved Tyr residue to the C18 position of Pchlide
(Fig. 1A) [812]. Recent advances in our understanding of the POR
reaction mechanism that illustrate POR is an important model for
studying light-activated enzyme dynamics and how light energy can
be harnessed to power biological catalysis [1318].
PORs originating froma variety of photosynthetic organisms, includ-
ing cyanobacteria and higher plants, have been studied in detail using
ultrafast and cryogenic spectroscopic techniques [8,16,1821]. Excited
state interactionsbetween the Pchlide molecule and active site residues
in the enzyme are proposed to result in a reactive charge-separated
state that facilitates the sequential hydride and proton transfer reac-
tions on a microsecond timescale [7,15,18,22]. The light-driven hydride
transfer step is coupled to enzyme motions within the lifetime of the
Pchlide excited state and is followed by the proton transfer step,
which is reliant on solvent dynamics and an extended network of mo-
lecular motions [22]. The catalytic cycle of POR also comprises a series
of conformational changes that are associated with ordered substrate
binding and product release steps to form a reactive active site confor-
mation [7,23,24]. However, despite a detailed molecular understanding
of the catalytic cycle from femtoseconds to seconds the structural fea-
tures of POR, and active site residues that contribute to light activation
and the reaction dynamics, is currently unknown due to the lack of a
crystal structure.
Journal of Photochemistry & Photobiology, B: Biology 161 (2016) 236243
Funding source statement: NSS is an Engineering and Physical Sciences Research
Council (EPSRC) Established Career Fellow (EP/J020192/1). Ultrafast spectroscopy was
performed at the Ultrafast Biophysics Facility, Manchester Institute of Biotechnology, as
funded by BBSRC Alert14 Award BB/M011658/1. The work is a contribution from the
EPSRC/BBSRC Synthetic Biology Research Centre SYNBIOCHEM (BB/M017702/1), which
provided infrastructure funding.
Corresponding authors at: Manchester Institute of Biotechnology, The University of
Manchester, Manchester, M1 7DN, UK.
E-mail addresses: nigel.scrutton@manchester.ac.uk (N.S. Scrutton),
Derren.Heyes@manchester.ac.uk (D.J. Heyes).
http://dx.doi.org/10.1016/j.jphotobiol.2016.05.029
1011-1344/© 2016 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Contents lists available at ScienceDirect
Journal of Photochemistry & Photobiology, B: Biology
journal homepage: www.elsevier.com/locate/jphotobiol
Sequence similarities with other short-chain dehydrogenase/reduc-
tase (SDR) enzymes identies POR as a member of the classical SDR
family of enzymes. The conserved glycine rich-Rossmann dinucleo-
tide-binding domain (GxxxGxG) and the catalytic YxxxK motifs of
POR have suggested a similar catalytic mechanism in POR and other
SDR enzymes [2527]. Consequently, although the overall sequence ho-
mology and pair wise identity between PORand any individual SDR en-
zyme is below the requirement to select an empirical structural
template, different homology models of POR have been proposed by
using closely related SDR enzymes as the structural template [5,27
30]. These homology models were composed of 7 β-sheets surrounded
by 9 α-helices (Fig. 1B). The extensions to some of the central helices
provide the Pchlide binding cleft, where the YxxxK catalytic motif was
included as part of an α-helix that results in residues Tyr193 and
Lys197 interacting with the Pchlide molecule. A unique 33 amino acid
residue insertion in POR was included as a loop region, as no corre-
sponding structural regions were present in the template structures,
and was predicted to beinvolved inestablishing an oligomerisation do-
main for POR inbarley [27,31].Basedonthesepreviousmodelswehave
now selected a number of active site variants to study the role of resi-
dues located in the proposed NADPH and Pchlide bindingsites in the re-
action mechanism of POR (Fig. 2). Detailed binding, multiple turnover
and single turnover laser studies have shown that although many of
the variants have reduced catalytic activity they all retain the ability to
bind substrates, albeit with reduced afnity in some cases. Several of
the active site variants show impaired photochemical behaviour, sug-
gesting that multiple residues are likely to be involved in the formation
of the excited state reactivecharge transfer state that is required prior
to hydride transfer. In the ongoing absence of a crystal structure we
have rationalized these ndings by using multiple SDR enzymes as a
structural template to produce an improved homology model of POR
from Thermosynechocystis elongatus.
2. Materials and Methods
2.1. Homology Modelling
The homology model of POR from Synechocystis sp. [27] was used as
the initial template for the development of a structural model of
Thermosynechococcus elongatus (BP-1) POR, using SWISS-MODEL ho-
mology server [32]. The NADPH coenzyme and Pchlide substrate were
docked in to the model with AutoDock and the ligand-docked structures
were relaxed by means of MD simulations by using the Amber ff94 force
eld implemented inGromacs 4.6.5. Prior to this, the base structure was
corrected and topology les for Gromacs were generated by using the
online tools from MDweb [33]. The quality of developed homology
model was analyzed using QMEAN server [34]. After optimisation the
nal model showed a QMEAN score of 0.50 and Z score of 2.95. In addi-
tion, the SWISS-MODEL homology server was used to determine the 15
SDR enzymes of known structures in the PDB that share the highest se-
quence identity with T. elongatus POR (Tables S1 and S2, Fig. S1). Mus-
tang-MR structural sieving server was used for the multiple structural
alignments of these SDR enzymes with POR and to remove residues
from the alignment that are below a threshold root mean squaredevia-
tion (RMSD) of 1.8 Å (Fig. S2) [35]. A putative uncharacterized SDR pro-
tein (PDB id: 3RD5) that has the highest sequence identity with POR
was used as the reference structure and thesieving procedure removed
68% of the 3RD5 residues at 1.8 Å sieving level using Mustang-MR siev-
ing server. The unsieved structural regions of 3RD5 were mapped on to
the homology model of T. elongatus POR by structural alignment to de-
termine the regions that are conserved between SDR enzymes and
POR and will therefore be modeled most accurately.
2.2. Sample Preparation and Site Directed Mutagenesis
All chemicals were obtained from Sigma-Aldrich. Recombinant POR
from the thermophilic cyanobacteria T. elongatus BP-1 was
overexpressed in Escherichia coli and puried as described previously
[36]. Site-directed mutagenesis of the por gene was performed using
the QuikChange kit (Stratagene) to mutate residues. The primers that
were used (MWG Eurons, Germany) are shown in the supplementary
information (Table S3). The correct mutations were conrmed by DNA
sequencing (MWG Eurons, Germany) and variant protein waspuried
as described previously [36]. The Pchlide pigment was produced and
puried as described previously [37].
2.3. Steady-state Activity and Substrate Binding Measurements
Steady-state activity measurements were carried out as described
previously using a Cary 50 spectrophotometer (Agilent Technologies)
[38]. The binding of NADPH coenzyme was monitored using uores-
cence energy resonance transfer in a Cary Eclipse uorimeter (Agilent
Technologies) [38] and the binding of Pchlide was measured by follow-
ing the red-shift in absorbance at 642 nm, essentially as described [24].
2.4. Laser Flash Photolysis
Absorption transients at 696 nm were measured at 298 K using an
LKS-60 ash photolysis instrument (Applied Photophysics Ltd.) with
the detection system (comprising probe light, rst monochromator,
sample, second monochromator and photomultiplier) at right angles
to the incident laser beam. Dark assembled enzymeNADPHPchlide
ternary complex samples were excited in a cuvette of 1-cm pathlength
as described previously [10] with a 6 ns laser pulse at450 nm (~30 mJ)
Fig. 1. The light-activated reduction of Pchlide catalyzed by POR. A, trans addition of
hydrogen across the C
17
_C
18
carbon double bond of Pchlide to form Chlide in the
chlorophyll biosynthesis pathway is catalyzed by protochlorophyllide oxidoreductase
(POR). B, the three-dimensional homology model of POR from Synechocystis, based on
the crystal structure of 7α-hydroxysteriod dehydrogenase as a structural template [27].
237B.R.K. Menon et al. / Journal of Photochemistry & Photobiology, B: Biology 161 (2016) 236243
using an OPO of a Q-switched Nd-YAG laser (Brilliant B, Quantel). The
reaction samples were prepared by taking 15 μM Pchlide and 250 μM
NADPH in the presence of 60.0 μM enzyme for wild-type POR. Based
on the dissociation constant for Pchlide an equal concentration of terna-
ry complex was used for the variant enzymes. Hence, the reaction sam-
ples (1 ml) were prepared by using 60 μM enzyme concentrations for
R38V, K42A, T147S and S189A T230S and 55 μM for S16C, G19A and
T230S. Higher concentrations of enzymes were used for N39V
(128 μM), N90A (106 μM), Y94F (78 μM), T145 A (160 μM), T230 A
(78 μM), T230F (175 μM), T147F (74 μM), N149 V (100 μM) and
H236A (71 μM) variant enzymes. Rate constants were measured from
the average of at least ve time dependent absorption measurements
by tting to a single exponential function.
2.5. Ultrafast Pump-Probe Transient Absorption Spectroscopy
A Ti:sapphire amplier system (Spectra Physics Solstice Ace) pro-
duced 6 mJ of800 nm pulsesat 1 kHz with 100 fs pulse duration.A por-
tion of the output of the amplier was used to pump a Topas Prime OPA
with associated NirUVis unit which was used to generate the pump
Fig. 2. The sequence alignment and conservation of amino acid among homologous POR enzymes. The protein sequence of T. elongatus POR is aligned with different homologous POR
enzymes. The amino acid residues that were targeted in this study are indicated by arrows. The conserved regions are indicated by black lines. PRALINE; protein multiple sequence
alignment web server tool (by Centre for Integrative Bioinformatics, Vrije Universiteit Amsterdam, Netherlands) was used for the alignment of protein sequences.
238 B.R.K. Menon et al. / Journal of Photochemistry & Photobiology, B: Biology 161 (2016) 236243
beam centred at 450 nm, with FWHM of ca. 10 nm. A broad band ultra-
fast pump-probe transient absorbance spectrometer Helios(Ultrafast
systems LLC) was used to collect data (at random time points) from
~1 ps to 2.6 ns with a time resolution of around 0.2 ps. The probe
beam consisted of a white light continuum generated in a sapphire crys-
tal and absorbance changes were monitoredbetween 475 and 725 nm.
A broad band sub-nanosecond pump-probe transient absorbance spec-
trometer Eos(Ultrafast systems LLC) was used to collect data (at ran-
dom time points) up to 2 μs with a time resolution of around 0.5 ns. A
2 kHz white-light continuum bre laser was used to generate the
probe pulses and the delay between pump and probe was controlled
electronically. For both sets of measurements samples were excited at
450 nm with 0.5 μJ power and a beam diameter of ~200 μm. Samples
(1.5 ml) were owed at a rate of approximately 30 ml/min through a
0.2 mm pathlength quartz cell (at room temperature) to ensure that a
different area of the sample is excited with each pump laser pulse. Sam-
ples were measured upon until theproportion of Chlide in the reaction
mixture became higher than 10%, which resulted in data collection
times of b5 min for the wild-type and in the range of 820 min for the
variants. Samples were prepared in the dark containing 500 μM POR,
200 μM Pchlide and 4 mM NADPH in activity buffer with 10% glycerol,
0.1% 2-mercaptoethanol, and 0.5% Triton X-100.
2.6. Global Analysis
The datasets from the Helios and Eos measurements were merged
by selecting data with the same (small) proportion of product present
and scaling the Eos dataset by a xed factor to match the intensity of
the ground state bleach feature to that in the Helios data. The data
were then analyzed globally using the open-source software Glotaran
[39]. This procedure reduces the matrix of change in absorbance as a
function of time and wavelength, to a model of one or more exponen-
tially decaying time components, as described in the main manuscript
[39], each with a corresponding difference spectrum (speciesassociated
difference spectra (SADS)). Errors quoted with the lifetime values are
the standard errors calculated during the global analysis. The lifetimes
quoted for the conversion between states also include contributions
from the rates of ground state recovery through both radiative and
non-radiative processes. For the analysis, the pre-excitationbackground
was subtracted, and Helios datasets corrected for spectral chirp, and the
datasets were tted to a simple sequential model where one species
converts to another, which then persists for the lifetime of the
experiment.
3. Results
3.1. A New Structural Model for POR
As there is currently no crystal structure available for the POR en-
zyme further insights into therole of putative active site residues in ca-
talysis has been gained by producing a moreaccurate homology model
for the enzyme. This has been achieved by selecting multiple templates
of known PDB structures with a high sequence similarity with POR,
rather than a single template, and by identifying the core regions con-
served across all of the SDR enzymes (Figs. S1 and S2). A total of 15
SDR enzymes were selected as the targets of sequence and structure
alignment with POR, where the identity of individual templates with
POR varied between 19 and 30% (Table S2, Fig. S1)[32]. By using a struc-
tural sieving server, the highly conserved regions were mapped to the
homology model to optimise the POR structure (Fig. 3). The overall
structural model of T. elongatus POR is similar to previous models (Fig.
3AC) [5,2730] with the highly conserved Rossmann-fold region, sur-
rounding β-sheets and α-helices and the catalytic tetrad residues, to-
gether with two insertions (residues 232 to 253 and residues 151 to
186) that are not present in other SDR enzymes and are therefore struc-
turally less well-dened. Consequently, this rened model of POR
provides important insights about the potential role of individual resi-
dues in catalysis. Many of the residues selected for mutagenesis are in
close enough proximity to interact with the NADPH coenzyme. Ser16
and Gly19 are part of the highly conserved GxxxGxG motif, whereas
Arg38, Asn39 and Lys42 are located close to the 2-phosphate of the
adenosine ribose group and Asn90 is located near to the adenine moiety
of NADPH (Figs. 3D, 4A). Other residues, such as Tyr94, Thr145, Thr147,
Asn149, Ser189, Thr230 and His236 are also positioned in theactive site,
close to both Pchlide and NADPH (Figs. 3D, 4A, B). Tyr94, Thr145,
Thr147 and Asn149 are all located in close proximity to the propionate
side-chain of the Pchlide molecule at the C17 position, whereas Thr230
is located at a key position near to the diphosphate chain joining the
adenosine and the nicotinamide ribose groups of NADPH. However, al-
though theSer189 and His236 residues arepositioned relatively nearto
the Pchlide molecule and have previously been implicated in Pchlide
binding from previous homology models [27], the new structural
model shows that this is unlikely as their distance from Pchlide is too
great to allow any direct interactions with the pigment (Fig. 4B).
3.2. Steady-state Characterization of Active Site POR Variants
The new structural model of POR, in addition to previous structural
homology models [5,2730], have highlighted a number of residues
that may be important for catalytic activity. These residues, which can
potentially interact with the NADPH coenzyme [40,41], the Pchlide sub-
strate or are in close proximity to the binding site of both substrates [5,
2730], were altered by site-directed mutagenesis and their catalytic
activity initiallydetermined under steady-state conditions (Table 1). Al-
though many of the variants retained close to wild-type levels of activ-
ity, several showed a signicant reduction in catalytic activity. In
particular, mutations to Asn39, Asn90, Thr145, Thr147 and Asn149
had a major impact on steady-state activity, whereas replacement of
the Thr230 residues with a bulkier Phe residue showed a marked de-
crease in activity, in contrast to the limited effects caused by replace-
ment with Ala or Ser.
In order to provide a more in-depthrationale for this variation in cat-
alytic activities, the ability of all variant enzymes to bind the coenzyme
NADPH and Pchlide was measured (Table 1). The dissociation constant
for NADPH was determined by measuring the FRET signal from Trp res-
idue(s) in the protein tothe bound NADPH coenzyme [38]. Several var-
iants, including S16C, G19A, R38V, K42A, T147F, T230A andT230F, had a
signicantly reduced afnity for NADPH compared to the wild-type en-
zyme, indicating a role for the targeted residues in coenzyme binding.
The binding of Pchlide to the enzymeNADPH complex has been mon-
itored by measuring the red-shift in the absorbance maximum of the
Pchlide molecule upon forming a ternary PORPchlideNADPH ternary
complex (Table 1)[24].Most of the variant POR enzymes exhibited only
minor changes in the Pchlide binding properties. However, exceptions
to this were the N39V, N90A, T145A, N149V and T230F variants,
which showed ~ 510 fold reduction in the afnity of the Pchlide
substrate.
3.3. Laser Excitation Measurements of the Site-directed Mutants
Laser photoexcitation measurements have been used to obtain in-
formation on the excited state and single-turnover kinetics of interme-
diate formation for wild-type POR and the variant enzymes. The
formation of a broad absorbance band at 696 nm represents the hydride
transfer chemistry from NADPH and the disappearance of this 696 nm
band represents the proton transfer step to the C18 position to form
the nal Chlide product [10,11]. Hence, the kinetics and relative yield
of formation/decay of the absorbance band at 696 was measured in
dark-assembled enzyme-substrate ternary complex samples after exci-
tation with a 6 ns laser pulse. For those variants whose reaction could be
accurately measured there were only minor differences in the rates of
hydride and proton transfer (Table 1). More signicantly, many of the
239B.R.K. Menon et al. / Journal of Photochemistry & Photobiology, B: Biology 161 (2016) 236243
mutations resulted in a major loss of the amplitude of the absorbance
signal at 696 nm upon photoexcitation (Table 1), suggesting that the
quantum yield or photochemical efciency of the hydride transfer
step is impaired in those variants (example transients are shown in
Fig. 3). In some cases, notably the N149V, T145A, T147S and T230F var-
iants, the photochemical efciency was so low (Table 1) that it was not
possible to measure accurate rates for the hydride and proton transfer
steps, indicating a potentially important role for these residues in pho-
tochemistry. Indeed, replacement of the Thr230 residue with a bulkier
Phe residue in the T230F variant reduced the photochemical efciency
signicantly compared to the T230A and T230S variants (Fig. 5).
To further explore this potential role in photochemistry the excited
state dynamics were investigated by pump-probe transient absorption
spectroscopy for 5 POR variants that showed an impaired photochemi-
cal efciency and compared to those obtained for Pchlide only and a
wild-type PORPchlideNADPH ternary complex. The time-resolved
absorption difference spectra from these measurements were then
modeled using global analysis to yield species associated difference
spectra (SADS). For clarity only the SADS are shown in the main manu-
script, whereas the raw time-resolved data (Figs. S3S9), and kinetic
traces showing ts at selected probe wavelengths (Fig. S10) can be
found in the supporting information. In the wild-type POR ternary
Fig. 3. Structural models of PORfrom Synechocystis (A) [27],Arabidopis thaliana (B) [30] and from T. elongatus (C) described in the present work are shown for comparison. D, a close-up
view of the active site illustrating the relative positions of the active site residues characterized in this study, which are shown in cyan, and the active site Tyrand Lys residues, which are
shown in red as sticks. The proteinbackbone is representedas a ribbon, NADPH is shown in yellowand Pchlide is shown in green. In Fig.A, B and C, protein backboneis coloured in gray to
indicate regions in the model which are not conserved.
Fig. 4. A close-up view ofthe NADPH (A) and Pchlide (B) binding site,showing potentialactive site aminoacid residues thatcould interactwith the NADH or Pchlidemolecule basedon the
derived structural model of T. elongatus POR.
240 B.R.K. Menon et al. / Journal of Photochemistry & Photobiology, B: Biology 161 (2016) 236243
enzyme-substrate complex the observed transient spectral changes can
be tted to a branched model along catalyticand non-catalyticpath-
ways after the formation of an intramolecular charge transfer state
(S
ICT
) as described previously [18]. In this model the formation of the
hydride transfer intermediate, with the characteristic absorbance band
at 696 nm, is observed with a lifetime of approximately 500 ns (Fig.
Table 1
The steady-state kinetic parameters, coenzyme and substrate binding constants and rates of hydrideand proton transfer for wild-type POR and variant enzymes. All rates and binding
constantswere measured at 25 °C. The ratesof hydride and proton transfer were measured from the average of at least ve timedependent absorption measurements by laser excitation
of similar levels of ternary enzyme-substrate complex and by following absorbance changes at 696 nm. The amplitude of the absorbance change at 696 nm isalsoshown.
Enzyme
k
cat
(s
1
)
K
d
NADPH
(nM)
K
d
Pchlide
(μM)
k
hydride
(×10
6
s
1
)
k
proton
(×10
4
s
1
)ΔmAbs 696 nm
Wild-type 0.17 ± 0.002 21 ± 1 5.6 ± 0.6 2.21 ± 0.06 2.72 ± 0.04 102 ± 1
S16C 0.16 ± 0.001 325 ± 90 3.5 ± 0.5 2.11 ± 0.03 2.68 ± 0.03 61 ± 3
G19A 0.16 ± 0.001 182 ± 31 3.5 ± 0.4 2.08 ± 0.07 2.67 ± 0.02 92 ± 2
R38V 0.16 ± 0.001 411 ± 76 5.4 ± 0.6 2.12 ± 0.03 2.69 ± 0.01 54 ± 4
N39V 0.02 ± 0.001 23 ± 6 28.9 ± 1.4 1.98 ± 0.02 2.71 ± 0.03 10 ± 0.4
K42A 0.16 ± 0.001 230 ± 81 4.5 ± 0.4 2.14 ± 0.09 2.67 ± 0.04 85 ± 3
N90A 0.07 ± 0.002 94 ± 7 21.4 ± 0.6 2.28 ± 0.04 3.17 ± 0.04 31 ± 2
Y94F 0.17 ± 0.003 63 ± 4 11.8 ± 0.7 2.23 ± 0.02 3.11 ± 0.01 80 ± 1
T145A 0.01 ± 0.001 33 ± 2 39.7 ± 2.1 n. d. n. d. 2 ± 0.2
T147S 0.02 ± 0.003 60 ± 3 6.5 ± 0.7 n. d. n. d. 2 ± 0.2
T147F 0.01 ± 0.003 117 ± 3 10.3 ± 0.6 n. d. n. d. 1 ± 0.1
N149V 0.01 ± 0.003 55 ± 2 19.2 ± 1.2 n. d. n. d. 2 ± 0.2
S189A 0.16 ± 0.001 38 ± 16 4.9 ± 0.4 2.15 ± 0.07 2.69 ± 0.01 69 ± 4
T230A 0.16 ± 0.001 181 ± 16 11.9 ± 1.4 1.63 ± 0.07 2.26 ± 0.03 51 ± 3
T230S 0.16 ± 0.001 48 ± 17 4.6 ± 0.5 1.71 ± 0.06 2.38 ± 0.06 60 ± 1
T230F 0.01 ± 0.001 436 ± 87 44.9 ± 3.5 n. d. n. d. 2 ± 0.2
H236A 0.10 ± 0.004 30 ± 11 9.4 ± 0.6 2.18 ± 0.03 2.72 ± 0.01 33 ± 1
Fig. 5. Hydride and proton transfer transients of wild-type and Thr230 variant POR
enzymes. Typical laser transients showing the absorbance change at 696 nm for the
wild-type POR and the Thr230 variants. The formation of the absorbance band at
696 nm represents hydride transfer and the disappearance of the 696 nm band
represents proton transfer [15]. The transients were measured using an equal ternary
complex concentration based on the Pchlide binding constant for each enzyme. The
experimental procedure is described in the materials and method section.
Fig. 6. Species associated difference spectra (SADS) resulting from a global analysis of the
time-resolved visible data for wild type and variant PORPchlideNADPH ternary
complexes after excitation at 450 nm. The data for wild type (A) and N39V (B) were
tted as described in the Supporting Information to a branched model (shown above
the panel), where 60% of the ICT state is converted to the solvated ICT state along the
non-catalyticpathway and 40% is converted to a reactiveICT state along the catalytic
pathway [18]. The data for T145A (C), T147S (D), N149V (E) and T230F (F) were tted
to the sequential model (shown above the panel) as described in Supporting
Information. Kinetic traces showing ts at selected energies are shown in Fig. S9.
241B.R.K. Menon et al. / Journal of Photochemistry & Photobiology, B: Biology 161 (2016) 236243
6A), similar to that obtained in the earlier laser ash photolysis mea-
surements. The excited state dynamics observed for the N39V variant
could also be tted to the same branched model with similar spectral
features and lifetimes (Fig. 6B), suggesting that photochemistry pro-
ceeds in an identical way to the wild type enzyme. In contrast, the hy-
dride transfer intermediate is completely absent in the N149V, T145A,
T147S and T230F variants (Fig. 6C-F), conrming that photochemistry
is impaired in these variants. In all of these cases the SADS required to
model the transient spectral data appear to be very similar to free
Pchlide (Fig. S10) and could be tted to 3 sequentially evolving expo-
nential functions. This represents a simple, linear decay pathway for
the excited state dynamics in these variants, involving the solvation of
the ICT state, followed by decay of the S1/ICT excited state into a long-
lived triplet state on the ns timescale. The triplet state then relaxes
back to the ground state on the μstimescale[18].
4. Discussion
It is now known that light-induced interactions between the protein
and the pigment occur within the lifetime of the Pchlide excited state
and are essential for the subsequent chemical steps, involving sequen-
tial hydride and proton transfer reactions, on themicrosecond timescale
[8,15,18,21,22]. However, the lack of any crystal structure for POR has
made it difcult to understand the structural elements and active site
residues that are required for this unique light-driven catalysis. Conse-
quently, a number of residues located in the proposed substrate binding
sites of POR have now been altered and their role in the reaction mech-
anism of POR investigated by static and time-resolved spectroscopic
measurements. Moreover, the results have been rationalized in terms
of active site structure based on a new structural model of the enzyme.
The majority of SDR enzymes contain a highlyconserved nucleotide-
binding domain, termed the Rossmann fold [4044], which is character-
ized by a TGxxxGxG motif (TGASSGVG in T. elongatus POR). Mutations
to Ser16 and Gly19 in this region were found to result in a lower afnity
for NADPH, although the catalytic activity remained unaffected. Similar-
ly, changes to Arg38 and Lys42 also reduced NADPH afnity and the
close proximity of these residues to the 2-phosphate and hydroxyl
groups of the adenosine ribose of the coenzyme in the structural
model support their important role in dening the coenzyme specicity
in the SDR family [45]. In addition, Thr230 may also be important for the
optimal positioning of the NADPH coenzyme in the active site as it ap-
pears to be located at a key position near to the diphosphate chain join-
ing the adenosine and the nicotinamide ribose groups. Although the
afnity for NADPH was reduced for both the T230 A and T230S variants
the effect was much greater when a bulkier Phe residue was placed at
this position.
In terms of Pchlide binding Asn39, Asn90, Thr145 and Asn149 ap-
pear to be important as mutations to these residues reduce both the cat-
alytic efciency of the enzyme and the afnity of the substrate. In the
majorityof SDR enzymes there is a catalytic tetrad of active site residues
that are essential for activity, consisting of Asn-Ser-Tyr-Lys [44,46].It
has been suggested that the Ser stabilizes the substrate and the Tyr
acts as a proton donor, whereas the Lys lowers the pKa of the Tyr-OH
to promote proton transfer and the Asn residue provides important in-
teractions with theLys to maintain its position inthe active site and pro-
mote the proton relay mechanism [44,46].The importance of the active
site Tyr andLys residues in POR catalysis have already been reported in
previous studies [11]. However, in POR from T. elongatus the Ser residue
from the catalytic tetrad is replaced by a Thr (Thr145), which is located
in close proximity to the catalytic Tyr and Lys and also to the NADPH co-
enzyme in the structural model. Hence, changes to this residue may sig-
nicantly inuence any interactions between the Tyr, Lys, NADPH and
Pchlide substrate. The Asn from the catalytic tetrad is Asn90 in POR,
which is likely to play a similar role to other SDR enzymes by stabilizing
the active site geometry through interactions with the ribose hydroxyl
group of the nicotinamide and the active site Lys [44,46]. However,
the present work has revealed that changes to two other residues,
Ser189 and His236, which had been suggested to be important for
Pchlide binding based on previous homology models [27],onlyhavea
minimal effect on substrate binding and catalytic activity. Although
the new structural model of POR described in the present work is simi-
lar to previous homology models [5,2730], the Pchlide binding sitedif-
fers signicantly and indicates that the proposed role of His236 in
chelating the central Mg ion of Pchlide is unlikely as there is no possibil-
ity for direct interaction with Mg
2+
. In addition,the previous suggestion
that Ser189 may be important in the correct positioning of the Pchlide
towards the pro-S face of NADPH [27] is also unlikely as the S189 A var-
iant retained wild-type levels of substrate binding and catalytic activity.
The crystal structure of the unrelated light-independent Pchlide reduc-
tase shows that thereis no requirement forthe Pchlide substrate to have
adirectMg
2+
coordination with amino acid residues in the active site
[6,7]. Hence, based on the new model, His236 together with other aro-
matic residues such as Phe233 and Phe237, may be involved in Pchlide
binding byproviding a pi stacking interaction with the substrate (Fig. 4).
However, His236 is replaced by serine in some homologous POR en-
zymes, which indicates that this residue is not an absolute necessity.
Moreover, Pro96 and Leu232 are close to the Mg
2+
in the new model
and it is possible that these residues are also involved either in a direct
or indirect coordination with the central metal ion (Fig. 4).
Importantly, the present work highlights how minor changes to the
architecture of the active site can have a profound effect on the efcien-
cy of photochemistry. Previous studies have shown that the photo-
chemistry is driven by excited state interactions between the active
site Tyr and the carboxyl group of the propionate side chain at the C17
position of Pchlide, which pulls electron density away from the C17
C18 double bond to form a reactivecharge transfer state [18].Thiscre-
ates an electron-decient site across the double bond, which triggers
the subsequent transfer of the negatively charged hydride from
NADPH [18]. Several of the active site variants, including T145A,
T147S, T147F, N149V and T230F exhibit impaired photochemical be-
haviour,implying a role for theseresidues inthe formation of the excit-
ed state reactivecharge transfer state. Based on the structural model
the Thr147 and Asn149 residues are in close enough proximity to inter-
act directly with the propionate side-chain at the C17 position of
Pchlide. As this region of the Pchlide molecule is essential for photo-
chemistry [18] it is likely that both of these residues are essential for
the excited state interactions that create the reactivecharge transfer
state. As discussed above, the close proximity of Thr145 to the catalytic
Tyr and Lys and the NADPH coenzyme may mean that any changes to
this residue can signicantly affect any excited state interactions be-
tween the Tyr, Lys, NADPH and Pchlide substrate. Although the
Thr230 is unlikely to directly participate in any excited state interac-
tions it maystill play an important role in maintaining the active site ge-
ometry to allow photochemistry to proceed efciently as replacement
with a bulkier Phe residue abolishes the photochemical step.
Concluding remarks
We have now highlighted a number of key residues in the active site
of POR that are important for coenzyme and substrate binding, as well
as the excited state processes required for catalysis. The role of these
residues have been veried by a new structural model for T. elongatus
POR, which supports all of the ndings from the binding, multiple turn-
over and single turnover laser studies on the active site variants. In the
absence of any crystal structure this work will provide the basis for fu-
ture functional studies of this key light-activated enzyme.
Appendix A. Supplementary data
Supplementary data to this article can be found online at http://dx.
doi.org/10.1016/j.jphotobiol.2016.05.029.
242 B.R.K. Menon et al. / Journal of Photochemistry & Photobiology, B: Biology 161 (2016) 236243
References
[1] R.D. Willows, Biosynthesis of chlorophylls from protoporphyrin IX, Nat. Prod. Rep.
20 (2003) 327341.
[2] D.J. Heyes, C.N. Hunter, Making light work of enzyme catalysis: protochlorophyllide
oxidoreductase, Trends Biochem. Sci. 30 (2005) 642649.
[3] C. Reinbothe, M. El Bakkouri, F. Buhr, N. Muraki, J. Nomata, G. Kurisu, Y. Fujita, S.
Reinbothe, Chlorophyll biosynthesis: spotlight on protochlorophyllide reduction,
Trends Plant Sci. 15 (2010) 614624.
[4] N.S. Scrutton, M.L. Groot, D.J. Heyes, Excited state dynamics and catalytic mecha-
nism of the light-driven enzyme protochlorophyllide oxidoreductase, Phys. Chem.
Chem. Phys. 14 (2012) 88188824.
[5] M. Gabruk, B. Mysliwia-Kurdziel, Light-dependent protochlorophyllide oxidoreduc-
tase: phylogeny, regulation, and catalytic properties, Biochemistry 54 (2015)
52555262.
[6] N. Muraki,J. Nomata, K. Ebata, T. Mizoguchi, T. Shiba, H. Tamiaki, G. Kurisu, Y. Fujita,
X-ray crystal structure of the light-independent protochlorophyllide reductase, Na-
ture 465 (2010) 110114.
[7] M.J. Bröcker, S. Schomburg, D.W. Heinz, D. Jahn, W.D. Schubert, J. Moser, Crystal
structure of the nitrogenase-like dark operative protochlorophyllide oxidoreductase
catalytic complex (ChlN/ChlB)
2
, J. Biol. Chem. 285 (2010) 2733627345.
[8] D.J. Heyes, C. Levy, M.Sakuma, D.L. Robertson, N.S. Scrutton, A twin-track approach
has optimized proton and hydride transfer by dynamically coupled tunneling dur-
ing the evolution of protochlorophyllide oxidoreductase, J. Biol. Chem. 286 (2011)
1184911854.
[9] D.J. Heyes, P. Heathcote, S.E.J. Rigby, M.A. Palacios, R. van Grondelle, C.N. Hunter,The
rst catalytic step of the light-driven enzyme protochlorophyllide oxidoreductase
proceeds via a charge transfer complex, J. Biol. Chem. 281 (2006) 2684726853.
[10] B.R.K. Menon, P.A. Davison, C.N. Hunter, N.S. Scrutton, D.J. Heyes, Mutagenesis alters
the catalytic mechanism of the light-driven enzyme protochlorophyllide oxidore-
ductase, J. Biol. Chem. 285 (2010) 21132119.
[11] B.R.K. Menon, J.P. Waltho, N.S. Scrutton, D.J. Heyes, Cryogenic and laser photoexcita-
tion studies identify multiple roles for active site residues in the light-driven en-
zyme protochlorophyllide oxidoreductase, J. Biol. Chem. 284 (2009) 1816018166.
[12] D.J. Heyes, C.N. Hunter, Site-directed mutagenesis of Tyr-189 and Lys-193 in
NADPH: protochlorophyllide oxidoreductase from Synechocystis, Biochem. Soc.
Trans. 30 (2002) 601604.
[13] B. Dietzek, W. Kiefer, G. Hermann, J. Popp, A.Schmitt, Solvent effects on the excited-
state processes of protochlorophyllide: a femtosecond time-resolved absorption
study, J. Phys. Chem. B 110 (2006) 43994406.
[14] B. Dietzek, W. Kiefer, A. Yartsev, V. Sundstrom, P. Schellenberg, P. Grigaravicius, G.
Hermann, J. Popp, M. Schmitt, The excited-state chemistry of protochlorophyllide
a: a time-resolved uorescence study, ChemPhysChem 7 (2006) 17271733.
[15] D.J. Heyes, M. Sakuma, S.P. de Visser, N.S. Scrutton, Nuclear quantum tunneling in
the light-activated enzyme protochlorophyllide oxidoreductase, J. Biol. Chem. 284
(2009) 37623767.
[16] D.J. Heyes, C.N. Hunter, I.H.M. van Stokkum, R. van Grondelle, M.L. Groot, Ultrafast
enzymatic reaction dynamics in protochlorophyllide oxidoreductase (vol 10, pg
491, 2003), Nat. Struct. Biol. 10 (2003).
[17] D.J. Heyes, A.V. Ruban, H.M. Wilks, C.N. Hunter, Enzymology below200 K: the kinet-
ics and thermodynamics of the photochemistry catalyzed by protochlorophyllide
oxidoreductase, Proc. Natl. Acad. Sci. U. S. A. 99 (2002) 1114511150.
[18] D.J. Heyes, S.J.O.Hardman, D. Mansell,J.M. Gardiner, N.S.Scrutton, Mechanistic reap-
praisal of early stage photochemistry in the light-driven enzyme protochlorophyllide
oxidoreductase, PLoS ONE 7 (2012).
[19] A. Garrone, N. Archipowa, P.F. Zipfel, G. Hermann, B. Dietzek, Plant
protochlorophyllide oxidoreductases A and B: catalytic efciency and initial reac-
tion steps, J. Biol. Chem. 290 (2015) 2853028539.
[20] D.J. Heyes, S.J.O.Hardman, T.M. Hedison, R. Hoeven, G.M. Greetham, M. Towrie, N.S.
Scrutton, Excited-state charge separation in the photochemical mechanism of the
light-driven enzyme protochlorophyllide oxidoreductase, Angew. Chem. Int. Ed.
54 (2015) 15121515.
[21] R. Hoeven,S.J.O. Hardman, D.J. Heyes, N.S. Scrutton, Cross-species analysis of protein
dynamics associated with hydride and proton transfer in the catalytic cycle of the
light-driven enzyme protochlorophyllide oxidoreductase, Biochemistry 55 (2016)
903913.
[22] D.J. Heyes, M. Sakuma, N.S. Scrutton, Solvent-slaved protein motions accompany
proton but not hydride tunneling in light-activated protochlorophyllide oxidore-
ductase, Angew. Chem. Int. Ed. 48 (2009) 38503853.
[23] D.J. Heyes, M. Sakuma, N.S. Scrutton, Laser excitation studies of the product release
steps in the catalytic cycle of the light-driven enzyme, protochlorophyllide oxidore-
ductase, J. Biol. Chem. 282 (2007) 3201532020.
[24] D.J. Heyes, B.R.K. Menon, M. Sakuma, N.S. Scrutton, Rate-limiting conformational
changes control the formation of the ternary enzymesubstrate complex in the
light-driven enzyme protochlorophyllide oxidoreductase (POR), Biochemistry 47
(2008) 1099110998.
[25] M.E. Baker, Protochlorophyllide reductase is homologous to human carbonyl reduc-
tase and pig 20-beta-hydroxysteroid dehydrogenase, Biochem. J. 300 (1994)
605607.
[26] N. Lebedev, M.P. Timko, POR structural domains important for the enzyme activity
in R-capsulatus complementation system, Photosynth. Res. 74 (2002) 153163.
[27] H.E. Townley, R.B. Sessions, A.R. Clarke, T.R. Dafforn, W.T. Grifths,
Protochlorophyllide oxidoreductase: a homology model examined by site-directed
mutagenesis, Proteins 44 (2001) 329335.
[28] C. Dahlin, H. Aronsson, H.M. Wilks, N. Lebedev, C. Sundqvist, M.P. Timko, The role of
protein surface charge in catalytic activity and chloroplast membrane association of
the pea NADPH: protochlorophyllide oxidoreductase (POR) as revealed by alanine
scanning mutagenesis, Plant Mol. Biol. 39 (1999) 309323.
[29] F. Buhr, M. El Bakkouri, O. Valdez, S. Pollmann, N. Lebedev, S. Reinbothe, C.
Reinbothe, Photoprotective role of NADPH: protochlorophyllide oxidoreductase A,
Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 1262912634.
[30] M. Gabruk, J. Grzyb, J. Kruk, B. Mysliwa-Kurdziel, Light-dependent and light-inde-
pendent protochlorophyllide oxidoreductases share similar sequence motifs in
silico studies, Photosynthetica 50 (2012) 529540.
[31] C. Reinbothe, A. Lepinat, M. Deckers, E. Beck, S. Reinbothe, The extra loop
distinguishing POR from the structurally related short-chain alcohol dehydroge-
nases is dispensable for pigment binding but needed for the assembly of light-har-
vesting POR-protochlorophyllide complex, J. Biol. Chem. 278 (2003) 816822.
[32] K. Arnold, L. Bordoli, J. Kopp, T. Schwede, The SWISS-MODEL workspace: a web-
based environment for protein structure homology modelling, Bioinformatics 22
(2006) 195201.
[33] Hospital, A., Andrio, P., Fenollosa, C., Cicin-Sain, D., Orozco, M. & Gelpí, J.L. (2012)
MDWeb and MDMoby: an integrated web-based platform for molecular dynamics
simulation, Bioinformatics 2012, 28(9):12781279.
[34] P. Benkert, S.C.E.Tosatto, D.Schomburg, QMEAN: a comprehensive scoring function
for model quality assessment.", Proteins: Struct., Funct., Bioinf. 71 (1) (2008)
261277.
[35] A.S. Konagurthu, C.F. Reboul, J.W. Schmidberger, J.A. Irving, A.M. Lesk, P.J. Stuckey,
J.C. Whisstock, A.M. Buckle, MUSTANG-MR structural sieving server: applications
in protein structural analysis and crystallography, PLoS ONE 5 (2010).
[36] D.J. Heyes, C.N. Hunter, Identication and characterization of the product release
steps within the catalytic cycle of protochlorophyllide oxidoreductase, Biochemistry
43 (2004) 82658271.
[37] D.J. Heyes, A.V. Ruban, C.N. Hunter, Protochlorophyllide oxidoreductase: spectro-
scopic characterization of the dark reactions, Biochemistry 42 (2003) 523528.
[38] M.J. McFarlane, C.N. Hunter, D.J. Heyes, Kinetic characterisation of the light-driven
protochlorophyllide oxidoreductase (POR) from Thermosynechococcus elongatus,
Photochem. Photobiol. Sci. 4 (2005) 10551059.
[39] J.J. Snellenburg, S.P. Laptenok, R. Seger, K.M. Mullen, I.H.M. Stokkun, Glotaran: a
Java-based graphical user interface for the R package TIMP, J. Stat. Softw. 49
(2012) 122.
[40] J.E. Bray, B.D. Marsden, U. Oppermann, The human short-chain dehydrogenase/re-
ductase (SDR) superfamily: a bioinformatics summary, Chem. Biol. Interact. 178
(2009) 99109.
[41] N. Tanaka, T. Nonaka, M. Nakanishi, Y. Deyashiki, A. Hara, Y.Mitsui, Crystal structure
of the ternary complex of mouse lung carbonyl reductase at 1.8 angstrom resolu-
tion: the structural origin of coenzyme specicity in the short-chain dehydroge-
nase/reductase family, Structures 4 (1996) 3345.
[42] N. Tanaka, T. Nonaka, K.T. Nakamura, A. Hara, SDR: structure, mechanism ofaction,
and substrate recognition, Curr. Org. Chem. 5 (2001) 89111.
[43] U. Oppermann, C. Filling, M. Hult, N. Shafqat, X.Q. Wu, M. Lindh, J. Shafqat, E.
Nordling, Y. Kallberg, B. Persson, H. Jornvall, Short-chain dehydrogenases/reduc-
tases (SDR): the 2002 update, Chem. Biol. Interact. 143 (2003) 247253.
[44] K. Kavanagh, H. Jornvall, B. Persson, U. Oppermann, The SDR superfamily: functional
and structural diversity within a family of metabolic and regulatory enzymes, Cell.
Mol. Life Sci. 65 (2008) 38953906.
[45] Y. Kallberg, B. Persson, Prediction of coenzyme specicity in dehydrogenases/reduc-
tases a hidden Markov model-based method and its application on complete ge-
nomes, FEBS J. 273 (2006) 11771184.
[46] C. Filling, K.D. Berndt, J. Benach, S. Knapp, T. Prozorovski, E. Nordling, R. Ladenstein,
H. Jornvall, U. Oppermann, Critical residues for structure and catalysis in short-chain
dehydrogenases/reductases, J. Biol. Chem. 277 (2002) 2567725684.
243B.R.K. Menon et al. / Journal of Photochemistry & Photobiology, B: Biology 161 (2016) 236243
... The hydrophobic region of Pchlide was thought to interact with hydrophobic residues to form a hydrophobic patch on the surface of the protein [12]. This model (Fig. 1A,B) was shown to be consistent with previous mutagenesis studies [15,16] and highlighted how a number of active site residues that are crucial for POR activity (e.g. Tyr193, Lys197 and Thr145) may interact with the propionic acid group at the C17 position of the Pchlide substrate (< 4Å) [12]. ...
... The differences between the models are likely to have significant implications for the currently proposed hypotheses for the mechanism of catalysis by POR. Previous spectroscopic studies have led to a proposed catalytic mechanism whereby Pchlide photochemistry triggers hydride transfer from NADPH to the C17 position of Pchlide on a timescale of $ 500 ns [6,7,16]. This is followed by proton transfer to the C18 position on the microsecond timescale ($ 50 μs), either directly from the putative active site Tyr193 proton donor or via active site water molecules [5][6][7]15]. ...
... This is followed by proton transfer to the C18 position on the microsecond timescale ($ 50 μs), either directly from the putative active site Tyr193 proton donor or via active site water molecules [5][6][7]15]. The roles of several active site residues in this mechanism have also been supported by previous mutagenesis studies [16]. However, the publication of the new structural model (PDB 7JK9) for the plant POR-Pchlide-NADPH ternary complex has led to an alternative proposed catalytic mechanism [14]. ...
Article
Full-text available
The photoenzyme protochlorophyllide oxidoreductase (POR) is an important enzyme for understanding biological H‐transfer mechanisms. It uses light to catalyse the reduction of protochlorophyllide to chlorophyllide, a key step in chlorophyll biosynthesis. Although a wealth of spectroscopic data have provided crucial mechanistic insight, a structural rationale for POR photocatalysis has proved challenging and remains hotly debated. Recent structural models of the ternary enzyme–substrate complex, derived from crystal and electron microscopy data, show differences in the orientation of the protochlorophyllide substrate and the architecture of the POR active site, with significant implications for the catalytic mechanism. Here, we use a combination of computational and experimental approaches to investigate the compatibility of each structural model with the hypothesised reaction mechanisms and propose an alternative structural model for the cyanobacterial POR ternary complex. We show that a strictly conserved tyrosine, previously proposed to act as the proton donor in POR photocatalysis, is unlikely to be involved in this step of the reaction but is crucial for Pchlide binding. Instead, an active site cysteine is important for both hydride and proton transfer reactions in POR and is proposed to act as the proton donor, either directly or through a water‐mediated network. Moreover, a conserved glutamine is important for Pchlide binding and ensuring efficient photochemistry by tuning its electronic properties, likely by interacting with the central Mg atom of the substrate. This optimal ‘binding pose’ for the POR ternary enzyme–substrate complex illustrates how light energy can be harnessed to facilitate enzyme catalysis by this unique enzyme.
... The hydrophobic region of Pchlide was thought to interact with hydrophobic residues to form a hydrophobic patch on the surface of the protein [12]. This model ( Figure 1A) was shown to be consistent with previous mutagenesis studies [15,16] and highlighted how a number of active site residues that are crucial for POR activity (e.g. Tyr193, Lys197 and Thr145) may interact with the propionic acid group at the C17 position of the Pchlide substrate (<4 Å) Pchlide [12]. ...
... The differences between the models are likely to have significant implications for the current proposed hypotheses for the mechanism of catalysis by POR. Previous spectroscopic studies have led to a proposed catalytic mechanism whereby Pchlide photochemistry triggers hydride transfer from NADPH to the C17 position of Pchlide on a timescale of ~ 500 ns [6,7,16]. This is followed by proton transfer to the C18 position on the microsecond timescale (~ 50 s), either directly from the putative active site Tyr193 proton donor or via active site water molecules [5][6][7]15]. ...
... This is followed by proton transfer to the C18 position on the microsecond timescale (~ 50 s), either directly from the putative active site Tyr193 proton donor or via active site water molecules [5][6][7]15]. The roles of several active site residues in this mechanism have also been supported by previous mutagenesis studies [16]. However, the publication of the new structural model (PDB 7JK9) for the plant POR-Pchlide-NADPH ternary complex has led to an alternative proposed catalytic mechanism [14]. ...
Preprint
The photoenzyme protochlorophyliide oxidoreductase (POR) is an important enzyme for understanding biological H-transfer mechanisms. It uses light to catalyse the reduction of protochlorophyllide (Pchlide) to chlorophyllide, a key step in chlorophyll biosynthesis. Although a wealth of spectroscopic data have provided crucial mechanistic insight about the light-driven reaction chemistry, a structural rationale for POR photocatalysis has proved more challenging and remains hotly debated. Recent structural models of the ternary enzyme-substrate complex, derived from crystal and electron microscopy data, show differences in the orientation of the Pchlide substrate and the architecture of the POR active site that have significant implications for the catalytic mechanism of Pchlide reduction. Here, we have used a combination of computational and experimental approaches to investigate the compatibility of each of these structural models with the hypothesised reaction mechanisms and propose an alternative structural model for the cyanobacterial POR-Pchlide-NADPH ternary complex based on these findings. Through detailed site-directed mutagenesis studies we show that a strictly conserved Tyr residue, which has previously been proposed to act as the proton donor in POR photocatalysis, is not likely to be involved in this step of the reaction but is crucial for Pchlide binding. Instead, an active site Cys residue is important for both hydride and proton transfer reactions in POR and is proposed to act as the proton donor, either directly or through a water-mediated network. Moreover, a conserved Gln residue is found to be important for Pchlide binding and ensuring efficient photochemistry by tuning its electronic properties, likely by interacting with the central Mg atom of the substrate. This optimal ‘binding pose’ for the POR ternary enzyme-substrate complex illustrates how light energy can be harnessed to facilitate enzyme catalysis by this unique enzyme.
... PORB binds NADPH as a hydrogen donor for substrate reduction through a Rossmann-fold domain located at the N terminus and can participate in establishing the enzyme catalytic pocket in concert with the YxxxK motif ( Figure 4B) [19,20]. The Pchlide loop and helix α10 in RsPORB are involved in substrate binding [21], which induces a conformational change in helix α10, resulting in POR oligomerization [22,23]. Phylogenetic analysis with RsPORB and other known PORB proteins from various plant species indicated that RsPORB fell within the well-supported eudicot clade that includes PORBs from Brassica rapa and Arabidopsis, consistent with their close taxonomic relationship in the Brassicaceae (Supplementary Figure S1). ...
... The high conservative characteristics of POR proteins, including NADPH and Pchlide binding sites, assumes that RsPORB has the function of chlorophyll biosynthesis in radish. conformational change in helix α10, resulting in POR oligomerization [22,23]. Phylogenetic analysis with RsPORB and other known PORB proteins from various plant species indicated that RsPORB fell within the well-supported eudicot clade that includes PORBs from Brassica rapa and Arabidopsis, consistent with their close taxonomic relationship in the Brassicaceae (Supplementary Figure S1). ...
Article
Full-text available
Radish (Raphanus sativus) plants exhibit varied root colors due to the accumulation of chlorophylls and anthocyanins compounds that are beneficial for both human health and visual quality. The mechanisms of chlorophyll biosynthesis have been extensively studied in foliar tissues but remain largely unknown in other tissues. In this study, we examined the role of NADPH:protochlorophyllide oxidoreductases (PORs), which are key enzymes in chlorophyll biosynthesis, in radish roots. The transcript level of RsPORB was abundantly expressed in green roots and positively correlated with chlorophyll content in radish roots. Sequences of the RsPORB coding region were identical between white (948) and green (847) radish breeding lines. Additionally, virus-induced gene silencing assay with RsPORB exhibited reduced chlorophyll contents, verifying that RsPORB is a functional enzyme for chlorophyll biosynthesis. Sequence comparison of RsPORB promoters from white and green radishes showed several insertions and deletions (InDels) and single-nucleotide polymorphisms. Promoter activation assays using radish root protoplasts verified that InDels of the RsPORB promoter contribute to its expression level. These results suggested that RsPORB is one of the key genes underlying chlorophyll biosynthesis and green coloration in non-foliar tissues, such as roots.
... Pchlide binds predominately by hydrophobic interactions, with a few polar residues possibly responsible for the substrate specificity of LPOR: for example, Q331 is proximal to Mg 2+ of the tetrapyrrole ring, H319 and Y177 are close to the carbonyl group of Pchlide, and Y276 possibly interacts with the carboxyl group of the propionate moiety (Fig. 3a). In support of our Pchlide binding site, evolutionarily conserved residues that interact with Pchlide in our structure, namely AtPORB T230 and Y306, have previously been shown to decrease Pchlide binding when mutated in cyanobacterial LPORs 6,17,18 . While a binding site for Pchlide was proposed based on homology models and assumed proximity to NADPH 5,17 , our observation that Pchlide is partially embedded within the amphipathic region of the membrane's outer leaflet is unexpected. ...
... Interestingly, similar to S228A in AtPORB, T145A was shown to severely affect the rates of both photostimulated proton and hydride transfer in Thermosynechococcus LPOR 17 . After light-induced catalysis, we speculate that the rearrangement of helix α10 may play a role in the disassembly of the filament. ...
Article
Full-text available
Chlorophyll biosynthesis, crucial to life on Earth, is tightly regulated because its precursors are phototoxic1. In flowering plants, the enzyme light-dependent protochlorophyllide oxidoreductase (LPOR) captures photons to catalyse the penultimate reaction: the reduction of a double bond within protochlorophyllide (Pchlide) to generate chlorophyllide (Chlide)2,3. In darkness, LPOR oligomerizes to facilitate photon energy transfer and catalysis4,5. However, the complete three-dimensional structure of LPOR, the higher-order architecture of LPOR oligomers and the implications of these self-assembled states for catalysis, including how LPOR positions Pchlide and the co-factor NADPH, remain unknown. Here, we report the atomic structure of LPOR assemblies by electron cryo-microscopy. LPOR polymerizes with its substrates into helical filaments around constricted lipid bilayer tubes. Portions of LPOR and Pchlide insert into the outer membrane leaflet, targeting the product, Chlide, to the membrane for the final reaction site of chlorophyll biosynthesis. In addition to its crucial photocatalytic role, we show that in darkness LPOR filaments directly shape membranes into high-curvature tubules with the spectral properties of the prolamellar body, whose light-triggered disassembly provides lipids for thylakoid assembly. Moreover, our structure of the catalytic site challenges previously proposed reaction mechanisms6. Together, our results reveal a new and unexpected synergy between photosynthetic membrane biogenesis and chlorophyll synthesis in plants, orchestrated by LPOR. The light-dependent protochlorophyllide oxidoreductase (LPOR) is involved in chlorophyll synthesis and insertion into the thylakoid membrane. Here, a high-resolution structure of membrane-bound LPOR advances the mechanistic understanding of chlorophyll synthesis and thylakoid biogenesis.
... Information about its structure is limited to relative amounts of secondary structure elements, experimentally determined by circular dichroism, 18 and predictions of secondary and tertiary structures obtained using various bioinformatic tools. [18][19][20][21][22][23] Several in vivo and in vitro washing experiments and mutagenesis experiments have been performed in attempts to elucidate the strength of POR's membrane association. The results show that its association with thylakoid or etioplast inner membranes requires NADPH and ATP. ...
Article
Full-text available
NADPH:protochlorophyllide (Pchlide) oxidoreductase (POR) is a key enzyme of chlorophyll biosynthesis in angiosperms. It is one of few known photoenzymes, which catalyzes the light-activated trans-reduction of the C17-C18 double bond of Pchlide's porphyrin ring. Due to the light requirement, dark-grown angiosperms cannot synthesize chlorophyll. No crystal structure of POR is available, so to improve understanding of the protein's three-dimensional structure, its dimerization, and binding of ligands (both the cofactor NADPH and substrate Pchlide), we computationally investigated the sequence and structural relationships among homologous proteins identified through database searches. The results indicate that α4 and α7 helices of monomers form the interface of POR dimers. On the basis of conserved residues, we predicted 11 functionally important amino acids that play important roles in POR binding to NADPH. Structural comparison of available crystal structures revealed that they participate in formation of binding pockets that accommodate the Pchlide ligand, and that five atoms of the closed tetrapyrrole are involved in non-bonding interactions. However, we detected no clear pattern in the physico-chemical characteristics of the amino acids they interact with. Thus, we hypothesize that interactions of these atoms in the Pchlide porphyrin ring are important to hold the ligand within the POR binding site. Analysis of Pchlide binding in POR by molecular docking and PELE simulations revealed that the orientation of the nicotinamide group is important for Pchlide binding. These findings highlight the complexity of interactions of porphyrin-containing ligands with proteins, and we suggest that fit-inducing processes play important roles in POR-Pchlide interactions. This article is protected by copyright. All rights reserved.
... Based on its amino acid sequence it is considered as a soluble and globular protein with surprisingly high contents of basic and hydrophobic amino acids (Schoefs and Franck, 2003;Masuda and Takamiya, 2004;Heyes and Hunter, 2005;. Circular dichroism studies and bioinformatic tools predicted its secondary and tertiary structure (Schulz et al., 1989;Darrah et al., 1990;Birve et al., 1996;Dahlin et al., 1999;Townley et al., 2001;Buhr et al., 2008;Gabruk et al., 2012Menon et al., 2016;Gholami et al., 2018), but hydropathy plots did not identify potential hydrophobic transmembrane segments in it (Benli et al., 1991;Spano et al., 1992). This is surprising as 98% of LPOR present in etioplasts has been localized to PLB membranes (Ryberg and Sundqvist, 1982a;Ikeuchi and Murakami, 1983), and it represents the major protein of isolated PLB fractions (Ryberg and Sundqvist, 1982a;Selstam and Sandelius, 1984;von Zychlinski et al., 2005;Blomqvist et al., 2008;Kanervo et al., 2008). ...
Article
Full-text available
Chlorophyll (Chl) is essential for photosynthesis and needs to be produced throughout the whole plant life, especially under changing light intensity and stress conditions which may result in the destruction and elimination of these pigments. All steps of the Mg-branch of tetrapyrrole biosynthesis leading to Chl formation are carried out by enzymes associated with plastid membranes. Still the significance of these protein-membrane and protein-lipid interactions in Chl synthesis and chloroplast differentiation are not very well-understood. In this review, we provide an overview on Chl biosynthesis in angiosperms with emphasis on its association with membranes and lipids. Moreover, the last steps of the pathway including the reduction of protochlorophyllide (Pchlide) to chlorophyllide (Chlide), the biosynthesis of the isoprenoid phytyl moiety and the esterification of Chlide are also summarized. The unique biochemical and photophysical properties of the light-dependent NADPH:protochlorophyllide oxidoreductase (LPOR) enzyme catalyzing Pchlide photoreduction and located to peculiar tubuloreticular prolamellar body (PLB) membranes of light-deprived tissues of angiosperms and to envelope membranes, as well as to thylakoids (especially grana margins) are also reviewed. Data about the factors influencing tubuloreticular membrane formation within cells, the spectroscopic properties and the in vitro reconstitution of the native LPOR enzyme complexes are also critically discussed.
Article
Full-text available
The Light-Dependent Protochlorophyllide Oxidoreductase (LPOR) catalyzes a crucial step in chlorophyll biosynthesis: the rare biological photocatalytic reduction of the double C 00000000 00000000 00000000 00000000 11111111 00000000 11111111 00000000 00000000 00000000 C bond in the precursor, protochlorophyllide (Pchlide). Despite its fundamental significance, limited structural insights into the active complex have hindered understanding of its reaction mechanism. Recently, a high-resolution cryo-EM structure of LPOR in its active conformation challenged our view of pigment binding, residue interactions, and the catalytic process. Surprisingly, this structure contrasts markedly with previous assumptions, particularly regarding the orientation of the bound Pchlide. To gain insights into the substrate binding puzzle, we conducted molecular dynamics simulations, quantum-mechanics/molecular-mechanics (QM/MM) calculations, and site-directed mutagenesis. Two Pchlide binding modes were considered, one aligning with historical proposals (mode A) and another consistent with the recent experimental data (mode B). Binding energy calculations revealed that in contrast to the non-specific interactions found for mode A, mode B exhibits distinct stabilizing interactions that support more thermodynamically favorable binding. A comprehensive analysis incorporating QM/MM-based local energy decomposition unraveled a complex interaction network involving Y177, H319, and the C13¹ carboxy group, influencing the pigment's excited state energy and potentially contributing to substrate specificity. Importantly, our results uniformly favor mode B, challenging established interpretations and emphasizing the need for a comprehensive re-evaluation of the LPOR reaction mechanism in a way that incorporates accurate structural information on pigment interactions and substrate-cofactor positioning in the binding pocket. The results shed light on the intricacies of LPOR's catalytic mechanism and provide a solid foundation for further elucidating the secrets of chlorophyll biosynthesis.
Article
Herbicide resistance is a prevalent problem that has posed a foremost challenge to crop production worldwide. Light-dependent enzyme NADPH: protochlorophyllide oxidoreductase (LPOR) in plants is a metabolic target that could satisfy this unmet demand. Herein, for the first time, we embarked on proposing a new mode of action of herbicides by performing structure-based virtual screening targeting multiple LPOR binding sites, with the determination of further bioactivity on the lead series. The feasibility of exploiting high selectivity and safety herbicides targeting LPOR was discussed from the perspective of the origin and phylogeny. Besides, we revealed the structural rearrangement and the selection key for NADPH cofactor binding to LPOR. Based on these, multitarget virtual screening was performed and the result identified compounds 2 affording micromolar inhibition, in which the IC50 reached 4.74 μM. Transcriptome analysis revealed that compound 2 induced more genes related to chlorophyll synthesis in Arabidopsis thaliana, especially the LPOR genes. Additionally, we clarified that these compounds binding to the site enhanced the overall stability and local rigidity of the complex systems from molecular dynamics simulation. This study delivers a guideline on how to assess activity-determining features of inhibitors to LPOR and how to translate this knowledge into the design of novel and effective inhibitors against malignant weed that act by targeting LPOR.
Article
Enzymatic photocatalysis is seldom used in biology. Photocatalysis by light-dependent protochlorophyllide oxidoreductase (LPOR)—one of only a few natural light-dependent enzymes—is an exception, and is responsible for the conversion of protochlorophyllide to chlorophyllide in chlorophyll biosynthesis. Photocatalysis by LPOR not only regulates the biosynthesis of the most abundant pigment on Earth but it is also a ‘master switch’ in photomorphogenesis in early plant development. Following illumination, LPOR promotes chlorophyll production, plastid membranes are transformed and the photosynthetic apparatus is established. Given these remarkable, light-induced pigment and morphological changes, the LPOR-catalysed reaction has been extensively studied from catalytic, physiological and plant development perspectives, highlighting vital, and multiple, cellular roles of this intriguing enzyme. Here, we offer a perspective in which the link between LPOR photocatalysis and plant photomorphogenesis is explored. Notable breakthroughs in LPOR structural biology have uncovered the structural–mechanistic basis of photocatalysis. These studies have clarified how photon absorption by the pigment protochlorophyllide—bound in a ternary LPOR–protochlorophyllide–NADPH complex—triggers photocatalysis and a cascade of complex molecular and cellular events that lead to plant morphological changes. Photocatalysis is therefore the master switch responsible for early-stage plant development and ultimately life on Earth. Light-dependent protochlorophyllide oxidoreductase (LPOR) is a light-activated enzyme that catalyses a vital step in chlorophyll biosynthesis and acts as a key regulator of plant greening. In this Review, the authors summarize recent progress in the functional, chemical and structural understanding of LPOR photocatalysis in plants.
Article
Full-text available
The enzyme protochlorophyllide oxidoreductase (POR, EC 1.3.1.33) has a key role in plant development. It catalyzes one of the later steps in chlorophyll synthesis, the light-induced reduction of protochlorophyllide (PChlide) into chlorophyllide (Chlide) in the presence of NADPH. Two isozymes of plant POR, POR A and POR B from barley, which differ in their function during plant life, are compared with respect to their substrate binding affinity, catalytic efficiency, and catalytic mechanism. POR B as compared with POR A shows an 5-fold higher binding affinity for PChlide and an about 6-fold higher catalytic efficiency measured as kcat/Km. Based on the reaction intermediates, which can be trapped at low temperatures the same reaction mechanism operates in both POR A and POR B. In contrast to results reported for POR enzymes from cyanobacteria, the initial light-driven step, which occurs at temperatures below 180 K already involves the full chemistry of the photoreduction and yields the reaction product, Chlide, in an enzyme-bound form. The subsequent dark reactions, which include cofactor (NADP+) release and cofactor (NADPH) rebinding, show different temperature dependences for POR A and POR B and suggest a higher conformational flexibility of POR B in the surrounding active center. Both the higher substrate binding affinity and well adapted enzyme dynamics are held responsible for the increased catalytic activity of POR B as compared with POR A.
Article
Full-text available
The unique light-driven enzyme protochlorophyllide oxidoreductase (POR) is an important model system for understanding how light energy can be harnessed to power enzyme reactions. The ultrafast photochemical processes, essential for capturing the excitation energy to drive the subsequent hydride- and proton-transfer chemistry, have so far proven difficult to detect. We have used a combination of time-resolved visible and IR spectroscopy, providing complete temporal resolution over the picosecond–microsecond time range, to propose a new mechanism for the photochemistry. Excited-state interactions between active site residues and a carboxyl group on the Pchlide molecule result in a polarized and highly reactive double bond. This so-called “reactive” intramolecular charge-transfer state creates an electron-deficient site across the double bond to trigger the subsequent nucleophilic attack of NADPH, by the negatively charged hydride from nicotinamide adenine dinucleotide phosphate. This work provides the crucial, missing link between excited-state processes and chemistry in POR. Moreover, it provides important insight into how light energy can be harnessed to drive enzyme catalysis with implications for the design of light-activated chemical and biological catalysts.
Article
Full-text available
In the present studies, we have found a fragment of amino acid sequence, called TFT motif, both in light-dependent protochlorophyllide oxidoreductase (LPOR) and in the L subunit of dark-operative (light-independent) protochlorophyllide oxidoreductases (DPOR). Amino acid residues of this motif shared similar physicochemical properties in both types of the enzymes. In the present paper, physicochemical properties of amino acid residues of this common motif, its spatial arrangement and a possible physiological role are being discussed. This is the first report when similarity between LPOR and DPOR, phylogenetically unrelated, but functionally redundant enzymes, is described.
Article
Full-text available
The light-driven enzyme protochlorophyllide oxidoreductase (POR) catalyzes the reduction of protochlorophyllide (Pchlide) to chlorophyllide (Chlide). This reaction is a key step in the biosynthesis of chlorophyll. Ultrafast photochemical processes within the Pchlide molecule are required for catalysis and previous studies have suggested that a short-lived excited-state species, known as I675*, is the first catalytic intermediate in the reaction and is essential for capturing excitation energy to drive subsequent hydride and proton transfers. The chemical nature of the I675* excited state species and its role in catalysis are not known. Here, we report time-resolved pump-probe spectroscopy measurements to study the involvement of the I675* intermediate in POR photochemistry. We show that I675* is not unique to the POR-catalyzed photoreduction of Pchlide as it is also formed in the absence of the POR enzyme. The I675* species is only produced in samples that contain both Pchlide substrate and Chlide product and its formation is dependent on the pump excitation wavelength. The rate of formation and the quantum yield is maximized in 50∶50 mixtures of the two pigments (Pchlide and Chlide) and is caused by direct energy transfer between Pchlide and neighboring Chlide molecules, which is inhibited in the polar solvent methanol. Consequently, we have re-evaluated the mechanism for early stage photochemistry in the light-driven reduction of Pchlide and propose that I675* represents an excited state species formed in Pchlide-Chlide dimers, possibly an excimer. Contrary to previous reports, we conclude that this excited state species has no direct mechanistic relevance to the POR-catalyzed reduction of Pchlide.
Article
Full-text available
In this work the software application called Glotaran is introduced as a Java-based graphical user interface to the R package TIMP, a problem solving environment for fit-ting superposition models to multi-dimensional data. TIMP uses a command-line user interface for the interaction with data, the specification of models and viewing of analysis results. Instead, Glotaran provides a graphical user interface which features interactive and dynamic data inspection, easier – assisted by the user interface – model specification and interactive viewing of results. The interactivity component is especially helpful when working with large, multi-dimensional datasets as often result from time-resolved spec-troscopy measurements, allowing the user to easily pre-select and manipulate data before analysis and to quickly zoom in to regions of interest in the analysis results. Glotaran has been developed on top of the NetBeans rich client platform and communicates with R through the Java-to-R interface Rserve. The background and the functionality of the application are described here. In addition, the design, development and implementation process of Glotaran is documented in a generic way.
Article
Experimental interrogation of the relationship between protein dynamics and enzyme catalysis is challenging. Light-activated protochlorophyllide oxidoreductase (POR) is an excellent model to investigate this relationship because photo-initiation of the reaction cycle enables coordinated turnover in a 'dark-assembled' ternary enzyme-substrate complex. The catalytic cycle involves sequential hydride and proton transfers (from NADPH and an active site tyrosine residue, respectively) to the substrate protochlorophyllide. Studies with a limited cross-species subset of POR enzymes (n = 4) have suggested that protein dynamics associated with hydride and proton transfer are distinct [Heyes et al J. Biol. Chem. 286, 2113 (2010)]. Here, we use steady-state assays and single turnover laser flash spectroscopy to analyze hydride and proton transfer dynamics in an extended series of POR enzymes taken from many species, including cyanobacteria, algae, embryophytes and angiosperms. Hydride/proton transfer in all eukaryotic PORs is faster compared to prokaryotic PORs, suggesting active site architecture has been optimized in eukaryotic PORs following endosymbiosis. Visible pump-probe spectroscopy was also used to demonstrate a common photoexcitation mechanism for representative POR enzymes from different branches of the phylogenetic tree. Dynamics associated with hydride transfer are localized to the active site of all POR enzymes and are conserved. However, dynamics associated with proton transfer are variable. Protein dynamics associated with proton transfer are also coupled to solvent dynamics in cyanobacterial PORs and these networks are likely required to optimize (shorten) the donor-acceptor distance for proton transfer. These extended networks are absent in algal and plant PORs. Our analysis suggests that extended networks of dynamics are disfavoured, possibly through natural selection. Implications for the evolution of POR and more generally for other enzyme catalysts are discussed.
Article
The present review is focused on light-dependent protochlorophyllide oxidoreductase (POR; EC 1.3.1.33). POR catalyses the penultimate reaction of chlorophyll biosynthesis, i.e. the light-triggered reduction of protochlorophyllide to chlorophyllide. In this reaction the chlorine ring of the chlorophyll molecule is formed, which is crucial for photosynthesis. POR is one of the very few enzymes that are driven by light; however, it is unique in the need for its substrate to absorb photons to induce the conformational changes in the enzyme, which are required for its catalytic activation. Moreover, the enzyme is also involved in the negative feedback of the chlorophyll biosynthesis pathway and it controls chlorophyll content due to its light-dependent activity. Even though it has been almost 70 years since the first isolation of active POR complexes, our knowledge of them has markedly advanced in recent years. In the review, we summarise the current state of knowledge about POR including the phylogenetic roots of POR, mechanisms of the regulation of por gene expression, regulation of POR activity, POR import into plastids, the role of POR in PLB formation as well as the molecular mechanism of protochlorophyllide reduction by POR. To our knowledge, no previous review has compiled such a broad set of recent findings on POR.
Article
The unique light-driven enzyme protochlorophyllide oxidoreductase (POR) is an important model system for understanding how light energy can be harnessed to power enzyme reactions. The ultrafast photochemical processes, essential for capturing the excitation energy to drive the subsequent hydride- and proton-transfer chemistry, have so far proven difficult to detect. We have used a combination of time-resolved visible and IR spectroscopy, providing complete temporal resolution over the picosecond–microsecond time range, to propose a new mechanism for the photochemistry. Excited-state interactions between active site residues and a carboxyl group on the Pchlide molecule result in a polarized and highly reactive double bond. This so-called “reactive” intramolecular charge-transfer state creates an electron-deficient site across the double bond to trigger the subsequent nucleophilic attack of NADPH, by the negatively charged hydride from nicotinamide adenine dinucleotide phosphate. This work provides the crucial, missing link between excited-state processes and chemistry in POR. Moreover, it provides important insight into how light energy can be harnessed to drive enzyme catalysis with implications for the design of light-activated chemical and biological catalysts.
Article
Short chain dehydrogenases/reductases (SDR) constitute a large protein family. The SDR family now includes more than 1,000 enzymes from humans, mammals, insects and bacteria, and exhibits a wide variety of substrate specificity for steroids, retinoids, prostaglandins, sugars, alcohols and other small molecules. These enzymes have a residue identity level of 15-30 perent. Much has been done in the last decade to understand the structure function relationships in the SDR enzymes. This review summarizes recent progress of structural and functional studies of the enzymes belonging to the SDR family (X ray crystal structure analyses and site directed mutagenesis studies). Based on these studies, the three dimensional structure, catalytic mechanism, coenzyme specificity, and substrate specificity of the SDR enzymes are discussed.
Article
MDWeb and MDMoby constitute a web-based platform to help access to molecular dynamics (MD) in the standard and high-throughput regime. The platform provides tools to prepare systems from PDB structures mimicking the procedures followed by human experts. It provides inputs and can send simulations for three of the most popular MD packages (Amber, NAMD and Gromacs). Tools for analysis of trajectories, either provided by the user or retrieved from our MoDEL database (http://mmb.pcb.ub.es/MoDEL) are also incorporated. The platform has two ways of access, a set of web-services based on the BioMoby framework (MDMoby), programmatically accessible and a web portal (MDWeb). Availability: http://mmb.irbbarcelona.org/MDWeb; additional information and methodology details can be found at the web site (http://mmb.irbbarcelona.org/MDWeb/help.php) Contact: gelpi{at}ub.edu; modesto.orozco{at}irbbarcelona.org Supplementary information: Supplementary data are available at Bioinformatics online.