fluorescence lifetime snapshots reveal two rapidly reversible … · 2012-10-29 · fluorescence...

6
Fluorescence lifetime snapshots reveal two rapidly reversible mechanisms of photoprotection in live cells of Chlamydomonas reinhardtii Kapil Amarnath a , Julia Zaks b,1 , Samuel D. Park a,1,2 , Krishna K. Niyogi c,d , and Graham R. Fleming a,d,3 a Department of Chemistry, University of California, Berkeley, CA 94720; b Graduate Group in Applied Science and Technology, University of California, Berkeley, CA 94720; c Howard Hughes Medical Institute, Department of Plant and Microbial Biology, University of California, Berkeley, CA 94720; and d Physical Biosciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720 Contributed by Graham R. Fleming, April 4, 2012 (sent for review February 27, 2012) Photosynthetic organisms avoid photodamage to photosystem II (PSII) in variable light conditions via a suite of photoprotective mechanisms called nonphotochemical quenching (NPQ), in which excess absorbed light is dissipated harmlessly. To quantify the contributions of different quenching mechanisms to NPQ, we have devised a technique to measure the changes in chlorophyll fluor- escence lifetime as photosynthetic organisms adapt to varying light conditions. We applied this technique to measure the fluores- cence lifetimes responsible for the predominant, rapidly reversible component of NPQ, qE, in living cells of Chlamydomonas reinhard- tii. Application of high light to dark-adapted cells of C. reinhardtii led to an increase in the amplitudes of 65 ps and 305 ps chlorophyll fluorescence lifetime components that was reversed after the high light was turned off. Removal of the pH gradient across the thylakoid membrane linked the changes in the amplitudes of the two components to qE quenching. The rise times of the amplitudes of the two components were significantly different, suggesting that the changes are due to two different qE mechanisms. We tentatively suggest that the changes in the 65 ps component are due to charge-transfer quenching in the minor light-harvesting complexes and that the changes in the 305 ps component are due to aggregated light-harvesting complex II trimers that have de- tached from PSII. We anticipate that this technique will be useful for resolving the various mechanisms of NPQ and for quantifying the timescales associated with these mechanisms. photosynthesis in vivo spectroscopy time-resolved fluorescence feedback de-excitation pH-dependent regulation T he pigment-protein complexes responsible for light harvesting in photosynthesis are highly dynamic and are able to alter their function in response to their environment (1, 2). In very low light, photosynthetic organisms maximize energy transfer from the chlorophylls in the light-harvesting antenna to those in the special pairs of the reaction centers of photosystems I and II (PSI and PSII), where charge separation occurs (3, 4). However, the capacity for photosynthesis in plants saturates at low light intensities (5). Light that is absorbed by chlorophylls in the light- harvesting antennae but that cannot be used for charge separa- tion presents a significant hazard to PSIIs of oxygen-evolving or- ganisms, because it can lead to the formation of high-energy species that can damage proteins and inhibit photosynthesis (5). To protect itself from photodamage and prevent the loss of useful photosynthesis during its repair (6), PSII performs a col- lection of regulatory mechanisms, called nonphotochemical quenching (NPQ) (1, 2), in which excess absorbed light is dissi- pated harmlessly as heat. In high light, before an NPQ quenching site has turned on, a high population of PSIIs contains closed re- action centers. This closure occurs because electron transfer in the reaction center from the plastoquinone Q A to Q B is slow re- lative to light absorption, transfer, and trapping in PSII (3). PSIIs with closed reaction centers have a fluorescence lifetime greater than 1 ns (710) and are susceptible to damage. NPQ mechan- isms turn on in response to a feedback signal triggered by the high light conditions over a timescale of seconds to tens of minutes. Once an NPQ quenching site has turned on in PSII, the lifetime of the excitation decreases well below 1 ns (710), and PSII is protected. Each mechanism of NPQ has a unique timescale for induction and for the lifetime of PSII once the NPQ quenching site associated with that mechanism has turned on. Measure- ments of NPQ as photosynthetic organisms adapt* to high light are typically done using pulsed amplitude modulated (PAM) chlorophyll fluorescence (11), which is a measurement of the fluorescence yield and thus does not distinguish between differ- ent mechanisms of NPQ. Transient absorption spectroscopy (12) and time-resolved fluorescence (710) have revealed changes in the quenching of chlorophyll excitation, but only by measure- ment before and after light adaptation. Picosecond-resolved spec- troscopic measurements or snapshots of the photosynthetic organism during light adaptation would distinguish between popu- lations of PSIIs undergoing different NPQ quenching processes. The major, rapidly reversible component of NPQ is called qE (1, 2). It is triggered by a pH gradient across the thylakoid membrane. While qE quenching sites are thought to occur in the light-harvesting complexes of PSII in Arabidopsis thaliana, which mechanisms are involved in qE quenching remains highly contro- versial (2). To measure both the lifetimes and the amplitudes associated with different qE mechanisms, we developed an appa- ratus that allowed us to directly measure the onset and relaxation of the qE-associated changes of the excited state lifetime of PSII in live cells of the green alga Chlamydomonas reinhardtii. Results and Discussion PAM Fluorescence Trace of C. reinhardtii. Fig. 1 shows a PAM fluor- escence trace for wild-type C. reinhardtii grown photoautotrophi- cally in high light (400 μmol photons m 2 s 1 ) until the late log phase of growth, and dark adapted for 20 min before measure- ment. Actinic light was applied at T ¼ 0 s, where T is the time axis along which light adaptation occurs. During the first 0.3 s of actinic light illumination (Inset), the fluorescence yield in- creased as the plastoquinone pool became fully reduced, and PSII reaction centers became saturated (13). The induction of NPQ Author contributions: K.A., K.K.N., and G.R.F. designed research; K.A., J.Z., and S.D.P. performed research; K.A. and G.R.F. analyzed data; and K.A., K.K.N., and G.R.F. wrote the paper. The authors declare no conflict of interest. *Strictly speaking, "adaptation" refers to genetic changes in organisms that occur on an evolutionary timescale, and "acclimation" refers to physiological and biochemical changes that occur during the lifetime of an organism. However, in this field, the two terms are used interchangeably. See refs. 1 and 2. 1 J.Z. and S.D.P. contributed equally to this work. 2 Present address: Department of Chemistry, University of Colorado, Boulder, CO 80309. 3 To whom correspondence should be addressed. E-mail: [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/ doi:10.1073/pnas.1205303109/-/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.1205303109 PNAS May 29, 2012 vol. 109 no. 22 84058410 CHEMISTRY PLANT BIOLOGY Downloaded by guest on May 24, 2020

Upload: others

Post on 23-May-2020

9 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Fluorescence lifetime snapshots reveal two rapidly reversible … · 2012-10-29 · Fluorescence lifetime snapshots reveal two rapidly reversible mechanisms of photoprotection in

Fluorescence lifetime snapshots reveal two rapidlyreversible mechanisms of photoprotection inlive cells of Chlamydomonas reinhardtiiKapil Amarnatha, Julia Zaksb,1, Samuel D. Parka,1,2, Krishna K. Niyogic,d, and Graham R. Fleminga,d,3

aDepartment of Chemistry, University of California, Berkeley, CA 94720; bGraduate Group in Applied Science and Technology, University of California,Berkeley, CA 94720; cHoward Hughes Medical Institute, Department of Plant and Microbial Biology, University of California, Berkeley, CA 94720; anddPhysical Biosciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720

Contributed by Graham R. Fleming, April 4, 2012 (sent for review February 27, 2012)

Photosynthetic organisms avoid photodamage to photosystem II(PSII) in variable light conditions via a suite of photoprotectivemechanisms called nonphotochemical quenching (NPQ), in whichexcess absorbed light is dissipated harmlessly. To quantify thecontributions of different quenching mechanisms to NPQ, we havedevised a technique to measure the changes in chlorophyll fluor-escence lifetime as photosynthetic organisms adapt to varyinglight conditions. We applied this technique to measure the fluores-cence lifetimes responsible for the predominant, rapidly reversiblecomponent of NPQ, qE, in living cells of Chlamydomonas reinhard-tii. Application of high light to dark-adapted cells of C. reinhardtiiled to an increase in the amplitudes of 65 ps and 305 ps chlorophyllfluorescence lifetime components that was reversed after thehigh light was turned off. Removal of the pH gradient across thethylakoid membrane linked the changes in the amplitudes of thetwo components to qE quenching. The rise times of the amplitudesof the two components were significantly different, suggestingthat the changes are due to two different qE mechanisms. Wetentatively suggest that the changes in the 65 ps component aredue to charge-transfer quenching in the minor light-harvestingcomplexes and that the changes in the 305 ps component are dueto aggregated light-harvesting complex II trimers that have de-tached from PSII. We anticipate that this technique will be usefulfor resolving the various mechanisms of NPQ and for quantifyingthe timescales associated with these mechanisms.

photosynthesis ∣ in vivo spectroscopy ∣ time-resolved fluorescence ∣feedback de-excitation ∣ pH-dependent regulation

The pigment-protein complexes responsible for light harvestingin photosynthesis are highly dynamic and are able to alter

their function in response to their environment (1, 2). In very lowlight, photosynthetic organisms maximize energy transfer fromthe chlorophylls in the light-harvesting antenna to those in thespecial pairs of the reaction centers of photosystems I and II(PSI and PSII), where charge separation occurs (3, 4). However,the capacity for photosynthesis in plants saturates at low lightintensities (5). Light that is absorbed by chlorophylls in the light-harvesting antennae but that cannot be used for charge separa-tion presents a significant hazard to PSIIs of oxygen-evolving or-ganisms, because it can lead to the formation of high-energyspecies that can damage proteins and inhibit photosynthesis (5).

To protect itself from photodamage and prevent the loss ofuseful photosynthesis during its repair (6), PSII performs a col-lection of regulatory mechanisms, called nonphotochemicalquenching (NPQ) (1, 2), in which excess absorbed light is dissi-pated harmlessly as heat. In high light, before an NPQ quenchingsite has turned on, a high population of PSIIs contains closed re-action centers. This closure occurs because electron transfer inthe reaction center from the plastoquinone QA to QB is slow re-lative to light absorption, transfer, and trapping in PSII (3). PSIIswith closed reaction centers have a fluorescence lifetime greaterthan 1 ns (7–10) and are susceptible to damage. NPQ mechan-

isms turn on in response to a feedback signal triggered by the highlight conditions over a timescale of seconds to tens of minutes.Once an NPQ quenching site has turned on in PSII, the lifetimeof the excitation decreases well below 1 ns (7–10), and PSII isprotected. Each mechanism of NPQ has a unique timescale forinduction and for the lifetime of PSII once the NPQ quenchingsite associated with that mechanism has turned on. Measure-ments of NPQ as photosynthetic organisms adapt* to high lightare typically done using pulsed amplitude modulated (PAM)chlorophyll fluorescence (11), which is a measurement of thefluorescence yield and thus does not distinguish between differ-ent mechanisms of NPQ. Transient absorption spectroscopy (12)and time-resolved fluorescence (7–10) have revealed changesin the quenching of chlorophyll excitation, but only by measure-ment before and after light adaptation. Picosecond-resolved spec-troscopic measurements or snapshots of the photosyntheticorganism during light adaptation would distinguish between popu-lations of PSIIs undergoing different NPQ quenching processes.

The major, rapidly reversible component of NPQ is calledqE (1, 2). It is triggered by a pH gradient across the thylakoidmembrane. While qE quenching sites are thought to occur in thelight-harvesting complexes of PSII in Arabidopsis thaliana, whichmechanisms are involved in qE quenching remains highly contro-versial (2). To measure both the lifetimes and the amplitudesassociated with different qE mechanisms, we developed an appa-ratus that allowed us to directly measure the onset and relaxationof the qE-associated changes of the excited state lifetime of PSIIin live cells of the green alga Chlamydomonas reinhardtii.

Results and DiscussionPAM Fluorescence Trace of C. reinhardtii. Fig. 1 shows a PAM fluor-escence trace for wild-type C. reinhardtii grown photoautotrophi-cally in high light (400 μmol photonsm−2 s−1) until the late logphase of growth, and dark adapted for 20 min before measure-ment. Actinic light was applied at T ¼ 0 s, where T is the timeaxis along which light adaptation occurs. During the first 0.3 sof actinic light illumination (Inset), the fluorescence yield in-creased as the plastoquinone pool became fully reduced, and PSIIreaction centers became saturated (13). The induction of NPQ

Author contributions: K.A., K.K.N., and G.R.F. designed research; K.A., J.Z., and S.D.P.performed research; K.A. and G.R.F. analyzed data; and K.A., K.K.N., and G.R.F. wrotethe paper.

The authors declare no conflict of interest.

*Strictly speaking, "adaptation" refers to genetic changes in organisms that occur on anevolutionary timescale, and "acclimation" refers to physiological and biochemicalchanges that occur during the lifetime of an organism. However, in this field, the twoterms are used interchangeably. See refs. 1 and 2.

1J.Z. and S.D.P. contributed equally to this work.2Present address: Department of Chemistry, University of Colorado, Boulder, CO 80309.3To whom correspondence should be addressed. E-mail: [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1205303109/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1205303109 PNAS ∣ May 29, 2012 ∣ vol. 109 ∣ no. 22 ∣ 8405–8410

CHEM

ISTR

YPL

ANTBIOLO

GY

Dow

nloa

ded

by g

uest

on

May

24,

202

0

Page 2: Fluorescence lifetime snapshots reveal two rapidly reversible … · 2012-10-29 · Fluorescence lifetime snapshots reveal two rapidly reversible mechanisms of photoprotection in

quenching pathways caused the fluorescence yield to rapidly de-crease for 0.3 s < T < 2.5 s, decrease much less for 2.5 s < T <10 s, and plateau by T ¼ 20 s at a yield slightly lower than theyield before actinic light was applied. The actinic light was turnedoff after T ¼ 20 s, and the fluorescence yield dropped below itsvalue before actinic light was turned on. This decrease in fluor-escence yield occurs because oxidation of the plastoquinone poolopened PSII reaction centers while NPQ quenching sites werestill turned on. NPQ turned off rapidly as the algae adapted tothe darkness over the next 60 s, suggesting that the NPQ observedhere is qE quenching.

The PAM measurement allowed us to qualitatively describe thedynamics as algae adapted to high light. However, it is possible thatdifferent qE processes contributed to the changes in fluorescenceyield in this organism and that the amplitudes and fluorescencelifetimes of these processes were averaged out. To determine thelifetimes and amplitudes of qE quenching processes, we measuredfluorescence decays of C. reinhardtii at different points on the Taxis as the algae induced qE in high light for 20 s and as the algaeturned off qE in an ensuing 60 s of darkness.

Measurement of Time-Resolved Fluorescence Decays During LightAdaptation. NPQ is typically measured at different time pointsalong a PAM trace by saturating pulses of light that close all PSIIreaction centers (11). We constructed an apparatus whereby a si-milar strategy could be applied to measuring fluorescence decaysas algae adapted to high light. The apparatus consists of a con-ventional single photon counting (SPC) setup with the addition ofan actinic light source and three shutters in front of the excitation,actinic, and detection beams. The apparatus was built such thatactinic light could be applied to the sample, with short periods inwhich the sample would interact with the laser to measure thetime-resolved fluorescence while the PSII reaction centers re-mained saturated (Fig. 2A). To record an adequate number offluorescence counts for the shortest increments in the adaptationtrajectory, the experiment was also performed using the pulsedlaser as both an actinic and a measuring light source (Fig. 2B).The repetition rate of the laser (75.7 MHz) and the pulse energyat the sample (5 pJ) were chosen to keep PSII reaction centersclosed and maximize fluorescence collection while avoiding sing-let-singlet and singlet-triplet annihilation conditions (SI Text). Ifsinglet-triplet annihilation was a significant factor, the resultsfrom the actinic light treatment and the laser light treatmentwould be different. In fact, measurements on the same sampleusing the two strategies gave very similar results, which validated

the use of the laser as both the actinic light source and the mea-suring light (Fig. S1, Fig. S2).

The apparatus measures fluorescence decays Fðt; TÞ at differ-ent points along the T axis, where t corresponds to the arrivaltime of fluorescence photons after excitation of the sample witha laser pulse. The measurement of Fðt; TÞ using the actinic lightsource began, at T ¼ 0, by opening the shutter in front of theactinic light beam. To make a fluorescence lifetime measurementduring qE induction, when 0 s < T < 20 s, the actinic light shut-ter was closed, shutters in front of a pulsed laser and a detectorwere opened, and a trigger was sent to the SPC board to initiatecollection and binning of fluorescence counts for a time ΔT. AfterΔT, to resume light adaptation, the laser and detector shutterswere closed, and the actinic light shutter was reopened. This strat-egy prevents actinic light from overloading the detector and allowsfor the collection of fluorescence decays during the qE inductionperiod. To measure fluorescence decays during the qE relaxationperiod, the actinic light shutter was closed at T ¼ 20 s. At a timeTrelax later the laser and detector shutters were opened, and theSPC board was triggered to collect and bin counts for ΔT, afterwhich the laser and detector shutters were closed.

Ideally, the collection time, ΔT, would be set such thatFðt; T þ ΔTÞ ≈ Fðt; TÞ. As seen in Fig. 1, Inset, the fluorescenceyield changes every 0.11 s during the first 3 s of the actinic lighttreatment, so ΔT should be less than that time. However, ap-proximately 1;000 fluorescence counts in the maximum bin mustbe obtained during ΔT to allow accurate fitting of the fluores-

Fig. 1. PAM fluorescence trace of wild-type C. reinhardtii. The actinic lightintensity during the light induction period between T ¼ 0 s and T ¼ 20 s was1;000 μmol photonsm−2 s−1. Inset shows the fluorescence yield during thefirst 3 s after the actinic light is applied. The time resolution was 0.11 s.

0 20 20 + TrelaxTime T (s)

A

B

T = 0.08 s

Actinic Light Measurement

Laser Light Measurement

13 ns

Nor

m. F

l. C

ount

s(lo

g sc

ale)

Arrival time t (ns)

Actinic Light DarkDark Dark

T (s) 0 0.1 0.3

Dark Laser Light

0 5

100

10-2

Fig. 2. Schematic of measurement. (A) In the actinic light measurement,the sample remained in the dark until time T ¼ 0 s, when the actinic lighttreatment began. Measurement of the fluorescence lifetime occurred peri-odically by closing the shutter to the actinic light beam and opening thosein front of the pulsed laser and the detector (thin rectangles). The actiniclight shutter was closed at T ¼ 20 s and a measurement was made a timeT relax later. (B) In the laser light treatment, laser light was applied for0 s < T < 20 s. Fluorescence acquisition periods were ΔT ¼ 0.08 s long andspaced 0.2 s apart. The first Fðt; TÞ was collected at T ¼ 0.1 s to avoid lossof counts for the first measurement due to the opening time for the shutter.A measurement was made at a time T relax as in the actinic light measurement.A representative fluorescence lifetime trace corresponding to a collectionperiod at the beginning of the 20 s induction time is shown. The data weresmoothed with a moving average filter with a span of 25 bins for visualclarity.

8406 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1205303109 Amarnath et al.

Dow

nloa

ded

by g

uest

on

May

24,

202

0

Page 3: Fluorescence lifetime snapshots reveal two rapidly reversible … · 2012-10-29 · Fluorescence lifetime snapshots reveal two rapidly reversible mechanisms of photoprotection in

cence decays. A compromise between these two considerationsand the electronic limitations of the photon counting boardallowed us to collect photons for ΔT ¼ 0.08 s every 0.2 s onthe T axis, provided the measurement Fðt; TÞ was repeated onmore than 10 different aliquots from an algal culture to increasethe number of counts in the maximum bin to approximately1;000. This ΔT was too small to use the actinic light source as inthe strategy described above because the shutters had a open/close time of approximately 40 ms. Even if faster shutters wereused, the sample would be exposed to the laser for 40% of thelight adaptation period. For the purposes of measuring light in-duction of qE in C. reinhardtii, we used the laser light as both anactinic light source and as a measuring light.

At Trelax, the time for which the algae are in darkness afterthe high light treatment, we wanted to measure a fluorescencedecay with the PSII reaction centers closed and without any qEinduced by the laser during the measurement. However, as seenin Fig. 1, Inset, it takes approximately 0.3 s of exposure of dark-adapted cells to high light before the reaction centers are closed.qE quenching turns on immediately after saturation of the reac-tion centers. Because these changes may occur at different ratesdepending on Trelax, we averaged the fluorescence curves col-lected between 0.2 and 1.0 s within ΔT. Selection of this intervalresulted in the observation of full reversal of changes in the fluor-escence lifetime caused by the 20 s laser light treatment. Only onemeasurement atTrelax was done per sample to avoid the influenceof the laser on any measurements at later Trelax times.

Fig. 3 shows the results of the experiment described in Fig. 2B.The qualitative dynamics of the fluorescence decays curvesmatched those seen in the PAM fluorescence trace in Fig. 1. Thefluorescence decay time increased in the first 0.3 s due to the clos-ing of all reaction centers in the excitation volume. The fluores-cence lifetime decreased substantially in the next 2 s, decreasedless from 2.5 to 10.1 s, and barely decreased from 10.1 to 19.5 s.Laser illumination was stopped after 20 s, and the lifetime re-turned to its initial value 60 s after the light treatment.

Amplitudes and Lifetimes from the Time-Resolved Fluorescence De-cays. To quantify the changes seen in Fig. 3 over T, fluorescencecounts were summed over all collected wavelengths for eachFðt; TÞ, and the resulting fluorescence decays were normalizedand globally fit using the method of least squares to the equation

Fðt; TÞ ¼ ∑i

AiðTÞe−tτi ; [1]

where the ith component has an amplitude AiðTÞ, such that∑iAiðTÞ ¼ 1. Three lifetime components were required to givea reasonable fit, as judged by the χ 2 value being between 0.8and 1.2 and uncorrelated residuals for each Fðt; TÞ. Each lifetimeand its corresponding amplitude do not necessarily correspondto one physical process and its absorption cross-section in thethylakoid membrane. Rather, each lifetime and its amplitude arepossibly the result of a sum of the amplitudes of several processesin the membrane with approximately the same lifetime. To calcu-

late the uncertainty in the fitted parameters, we performed boot-strapping (SI Text) on the fits, which was repeated 500 timesto obtain 68% confidence intervals for each of the fitted para-meters.

The lifetimes from the global fit were 64 (63, 65) ps, 305 ps(304, 320), and 1,000 (990, 1,074) ps. The 68% confidence intervalsindicated in parenthesis were taken as an average over several timepoints. For simplicity, we will refer to these components as the65 ps, 305 ps, and 1 ns components with amplitudes A65 psðTÞ,A305 psðTÞ, and A1 nsðTÞ, respectively.

Because variable fluorescence is associated only with PSIIin the thylakoid membrane (3) we associated any changes inthe amplitudes of these three components with respect to T withchanges in excitation trapping in PSII. A65 psðTÞ and A305 psðTÞare shown in Fig. 4 A and B, respectively. A65 psðTÞ þA305 psðTÞand A1 nsðTÞ are shown in Fig. 4C. The 68% confidence intervalsof the amplitudes at selected points are indicated in this figure byerror bars.

A65 psðTÞ (Fig. 4A) decreased slightly for the first 0.3 s of lightillumination (Inset), then rose from approximately 0.61 atT ¼ 0.3 s to approximately 0.68 when T ¼ 1 s and remainedconstant within error for the remainder of the light adaptationperiod. Based on this plot, no definite conclusions on thedynamics of this component could be made for 20 s < T < 80 sbecause of the uncertainty in the amplitude. A305 psðTÞ (Fig. 4B)also decreased for the first 0.3 s of illumination (Inset) to approxi-mately 0.16 but subsequently increased for the remainder ofthe saturating light treatment, reaching approximately 0.25 atT ¼ 20 s. Forty-five seconds after the light treatment ended,the amplitude decreased back to the value at T ¼ 0.3 s andappeared to decrease to an even lower value by 60 s in darkness.Finally, A1 nsðTÞ increased in the first 0.3 s of light illumination,decreased substantially from 0.3 s < T < 3 s, and plateaued by20 s in actinic light (Fig. 4C). It increased over 60 s in the darknessback to A1 nsðT ¼ 0.3 sÞ.

We ascribe the changes in amplitudes to changes in the quan-tities of PSIIs with such lifetimes. Coarse-grained models ofenergy transfer and trapping in PSII (14) as well as measurementsof the fluorescence lifetime (7–10) suggest that PSIIs with a sitethat can trap excitation energy, such as a qE site and/or an openreaction center, have a lifetime of less than 700 ps. PSIIs withclosed reaction centers and no access to an energy trapping sitehave a lifetime of 1 ns or greater. Therefore, we associate thedecrease in A65 psðTÞ and A305 psðTÞ in the first approximately0.3 s of illumination with a decrease in the number of PSIIs withaccess to an open reaction center as the plastoquinone poolbecomes fully reduced. The subsequent increase in A65 psðTÞandA305 psðTÞ with 0.3 s < T < 20 s is an increase in the numberof PSIIs with closed reaction centers and access to a qE quench-ing site. The decrease in A305 psðTÞ and, to a lesser extentA65 psðTÞ, from 20 s < T < 80 s we ascribe to the decrease inthe number of PSIIs with closed reaction centers and access toa qE quenching site as qE sites turn off in the darkness. Thechanges in A1 nsðTÞ are changes in the number of PSIIs withaccess to neither a qE site nor an open reaction center. Because,

100

Nor

m. F

l. C

ount

s

t (ns) t (ns) t (ns) t (ns) t (ns)

T=0.1-0.3 s T=0.3-2.5 s T=2.5-10.1 s T=10.1-19.9 s Trelax=1-60 s

0 5 0 5 0 5 0 50 5

10-1

10-2

Fig. 3. Normalized fluorescence decays from emission at 680 nm from the measurement described in Fig. 1B. The data were smoothed with a moving averagefilter with a span of 20 bins for visual clarity. The direction of the black arrow indicates increasing T . Blue represents early T and red represents late T within thetimespan specified in each panel.

Amarnath et al. PNAS ∣ May 29, 2012 ∣ vol. 109 ∣ no. 22 ∣ 8407

CHEM

ISTR

YPL

ANTBIOLO

GY

Dow

nloa

ded

by g

uest

on

May

24,

202

0

Page 4: Fluorescence lifetime snapshots reveal two rapidly reversible … · 2012-10-29 · Fluorescence lifetime snapshots reveal two rapidly reversible mechanisms of photoprotection in

for T > 0.3 s, the changes in A305 psðTÞ are clearly reversible andthose for the A65 psðTÞ may be reversible, we are observing oneand possibly two lifetimes associated with qE switching on and offin live cells of C. reinhardtii.

To test if the increases in both A65 psðTÞ and A305 psðTÞ arelinked to a transmembrane pH gradient and thus to qE, we mea-sured time-resolved fluorescence decays after 150 s of an actiniclight treatment and after adding 100 μM of the ionophore niger-icin to the sample after the light treatment (Fig. S3). Nigericinremoves the pH gradient across the thylakoid membrane andis typically used to determine whether changes in fluorescencelifetime are due to qE (7). The fluorescence decays from the twomeasurements were globally fit to three exponential decays, andthe results are shown in Table 1. The amplitudes of both shortcomponents (in this fit, 70 and 330 ps) decrease upon the additionof nigericin, suggesting that the changes in both A65 psðTÞ andA305 psðTÞ in Fig. 4 are due to qE. We associate the remainingamplitude of the 70 ps component with PSI quenching (9, 15,16), and we ascribe the remaining amplitude of the 330 ps com-ponent partially to PSI quenching and partly to detached, phos-phorylated light-harvesting complex II (LHCII) trimers (17).

Although A65 psðTÞ was not unambiguously reversible(Fig. 4A), the fact that this component was linked to the pHgradient and that other measurements indicate that this compo-nent’s amplitude changes are reversible (Fig. S2A) lead us tobelieve that this component is reversible and attributable to qE.

We fit A65 psðTÞ and A305 psðTÞ from T ¼ 0.3 s, when theplastoquinone pool was fully reduced, to the end of the lighttreatment at T ¼ 20 s to a simple rise curve with the form

AiðTÞ ¼ a�1 − exp

�−T − 0.3τrise

��þ c: [2]

The results of the fitting are shown in Table 2. The 95% con-fidence intervals of the rise time, τrise, for the two componentsdo not overlap. These results give evidence that the changes inA65 psðTÞ and A305 psðTÞ in Fig. 4 are due to two different qEtriggering processes.

Interpretation and Prospects. To interpret the origin of the two qEcomponents observed, we turn to the current hypotheses for themolecular mechanisms underlying qE, which have been put forthprimarily for plants. There is a consensus that the protonation ofthe PsbS protein as a result of a pH decrease in the thylakoidlumen is thought to be the trigger for the qE process and thatthe xanthophylls zeaxanthin and lutein are necessary for quench-ing in vivo (1, 2). However, hypotheses differ on the role ofxanthophylls and the location of the qE quenching site. In a seriesof papers, Avenson et al. (18) and Ahn et al. (19) have proposedthat charge-transfer quenching of chlorophyll excitations byxanthophylls in the minor complexes of PSII (CP24, CP26, andCP29) is an important contribution to qE in A. thaliana. In con-trast to this model, Ruban et al. (2) propose that low lumen pHleads to the aggregation of LHCII, the major light-harvestingcomplex of PSII. The aggregated LHCII trimers undergo a con-formational change (20) and quench chlorophyll excitationsby energy transfer to the S1 state of lutein and subsequent relaxa-tion (21). Holzwarth and coworkers have proposed that quench-ing occurs in both detached LHCII trimers and the minorcomplexes of PSII (9). C. reinhardtii lacks PsbS but Bonente et al.suggested that the LHCSR3 protein plays a similar triggering role(22). In addition, unlike PsbS, LHCSR3 binds pigments, andfluorescence lifetime measurements indicated that the proteincould perform pH-dependent quenching (22).

Changes in fluorescence lifetimes in response to high lightin C. reinhardtii have been reported by Holub et al. for large

0 20 40 60 80

0.8

0.9

A65

ps(T

) +

A30

5 ps

(T)

T (s)

0.1

0.2

A1

ns(T

)

0 20 40 60 80

0.55

0.6

0.65

0.7

T (s)

A65

ps(T

)

0 20 40 60 800.1

0.15

0.2

0.25

0.3

T (s)

A30

5 ps

(T)

0 1 2 3T (s)

0 1 2 3T (s)

A

B

C

Light Dark

Light Dark

Light Dark

Fig. 4. Changes in the amplitudes of the fluorescence lifetime componentsover T. (A) Amplitude of the 65 ps component [A65 psðTÞ, green circles], (B)amplitude of the 305 ps component [A30 5psðTÞ, violet circles], and (C) Ampli-tude of the 1 ns component [A1 nsðTÞ, blue circles] and the sum of the 65 psand 305 ps components [A65 psðTÞ þ A305 psðTÞ, red circles]. The error bars in-dicate 1 standard deviation in the uncertainty of that parameter and areshown for T ¼ 0.3 s, 1.9 s, 7.9 s, 17.7 s, and 50 s. Insets in A and B showthe A65 psðTÞ and A305 psðTÞ for 0.3 s < T < 3 s. The lines in Insets are shownto indicate the changes in the amplitude.

Table 1 Amplitudes and lifetimes for wild-type cells adaptedto high light (600 μmolphotonsm−2 s−1) for 150 s (lightadapted) and treated with nigericin after the high lighttreatment (nigericin)

Amplitudes

Condition 70 ps 330 ps 1.2 ns hτi (ps)Light adapted 0.51 0.43 0.06 250Nigericin 0.35 0.30 0.35 544

The fluorescence was collected at 680 nm and 68% confidenceintervals of the fits as calculated by bootstrapping were within�0.01 of the amplitudes shown.

Table 2 Fits for the increase inA65 psðTÞ andA305 psðTÞ for 0.3 s < T <20 s (Fig. 4 A and B)

Component τrise (s) a c

65 ps 0.68 (0.45, 1.47) 0.07 (0.05, 0.10) 0.60 (0.58, 0.62)310 ps 3.18 (2.27, 5.26) 0.07 (0.06, 0.09) 0.18 (0.16, 0.20)

The data were fit to Eq. 2. The 95% confidence bounds are shown inparenthesis.

8408 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1205303109 Amarnath et al.

Dow

nloa

ded

by g

uest

on

May

24,

202

0

Page 5: Fluorescence lifetime snapshots reveal two rapidly reversible … · 2012-10-29 · Fluorescence lifetime snapshots reveal two rapidly reversible mechanisms of photoprotection in

T values (23). However, there have been no measurements on thequenching mechanisms associated with qE in C. reinhardtii. Therewere, however, recent fluorescence lifetime imaging measure-ments (17) of a second component of NPQ, called state transitions,which involves the detachment of phosphorylated LHCII trimersfrom the PSII supercomplex and subsequent reattachment of someof the LHCII to PSI (24). Iwai et al. (17) observed a 250 ps com-ponent for detached, phosphorylated, and aggregated LHCIIs inC. reinhardtii and suggested that this lifetime could be the same fordetached, aggregated LHCIIs due to qE.

Turning to the data in Fig. 4, Table 1, and Table 2, A65 psðTÞincreases much more rapidly under saturating light than doesA305 psðTÞ after the reduction of the plastoquinone pool, reachingits maximum value at T ≈ 1.3 s. Kinetic modeling of the transientabsorption data showing formation of a cation radical in thyla-koids from A. thaliana suggested that the time for chlorophyllexcitations to be quenched in the minor complexes is approxi-mately 30 ps (25). A possible interpretation of the changes inA65 psðTÞ for T > 0.3 s is that they are due to changes in the num-ber of PSII supercomplexes that can perform the minor complex/charge-transfer mechanism. Because of the slower rise time ofA305 psðTÞ following reduction of the plastoquinone pool, we sug-gest that the changes in A305 psðTÞ for T > 0.3 s correspond tochanges in the number of aggregated LHCIIs that are detachedfrom PSII supercomplexes and that this lifetime may be related tothe 250 ps lifetime reported by Iwai et al. The protonation ofLHCSR3 may cause it to become a quencher along with the min-or complexes as well as a trigger for LHCII aggregation.

These tentative interpretations of the two components canbe tested by applying this technique to different mutants ofC. reinhardtii (26–28). Our interpretation for the two short life-times may also be tested by numerical modeling of energy trans-fer and trapping in PSII and LHCII aggregates (29, 30). The risetimes of A65 psðTÞ and A305 psðTÞ may be representative of thetime for the pH to drop in the lumen, the time for a rearrange-ment to happen in the thylakoid membrane (31, 32), or somecombination of the two.

More generally, the technique described here can easily beapplied to studying other photosynthetic organisms. The additionof the fluorescence lifetime time axis t to measurements of theadaptation of photosynthetic organisms to varied light conditionsshould help enable the elucidation of the molecular mechanismsof NPQ.

Concluding RemarksWe have developed an apparatus that can measure time-resolvedfluorescence snapshots of photosynthetic organisms as they adaptto varying light intensities. Our data on wild-type C. reinhardtiisuggest that there may be two mechanisms involved in the rapidlyreversible component of NPQ known as qE. We speculate thatone mechanism involves quenching in detached LHCII com-plexes whereas a second occurs in the minor complexes and/orthe LHCSR3 protein. These ideas can be tested by applicationof the method developed here to a wide variety of photosyntheticmutants both in C. reinhardtii and in other model organisms, byextension of the snapshot technique to time-resolved absorptionspectroscopy, and by numerical modeling of energy transfer andtrapping in PSII and the dynamics of the thylakoid membrane.

Materials and MethodsGrowth and Sample Preparation of C. reinhardtii. Cells of the 4A− strain weregrown in low light (40 μmol photonsm−2 s−1) with a reduced carbon source[tris-acetate phosphate (TAP) medium] for 1 to 2 d until exponential growthwas reached. Approximately 1 × 105 cells from this culture were used to in-oculate 50 mL of minimal [high salt (HS)] media. The HS culture was grown inhigh light (400–440 μmol photonsm−2 s−1) in air, with ambient CO2 as theonly carbon source. Cultures that were in the late-logarithmic growth phase,between 3 and 5 × 106 cellsmL−1, were used for experiments. Before perform-

ing fluorescence measurements on the samples, the cells were dark incubatedfor 30–50 min. The sample was prepared in a 1 cm pathlength cuvette. Thecuvette contained algae diluted to a concentration of 0.5 OD at 410 nm with500 μL of 0.1% agarose to suspend the algae and water to 1.5 mL.

PAM Fluorescence Measurements. The maximum efficiency of PSII (Fv∕Fm) wasmeasured as described in Peers et al. (26) and was approximately 0.6 for allmeasurements. The measurement shown in Fig. 1 was done as described inthe text.

Fluorescence Lifetime Apparatus. Time-resolved fluorescence measurementswere acquired using a time-correlated single photon counting (TCSPC) appa-ratus. A commercial mode-locked oscillator (Mira 900F; Coherent) pumped bya diode-pumped, frequency-doubled Nd:YVO4 laser (Verdi V-10; Coherent)generated approximately 150 fs pulses with a repetition rate of 76 MHzandwas tuned to 820 nmwith an FWHMof 12 nm. This output was frequencydoubled to 410 nm using a 1-mm-thick beta barium borate crystal. The pulseenergy at the sample was 10 pJ. Fluorescence emission was sent through apolarizer set at the magic angle. The fluorescence was then sent through aspectrograph (Newport; 77400-M) and detected with a microchannel platephotomultiplier tube (MC-PMT) (Becker-Hickl; PML-16C). The MC-PMT coulddiscriminate the fluorescence photons into 16 channels, separated 12 nmapart from each other. Each channel contained the fluorescence decay forphotons at that wavelength, �1 nm. The MC-PMT was controlled using theDCC-100 detector control (Becker-Hickl), with the gain set to 90%. Thephotons were collected into 3.6 ps wide bins using a Becker-Hickl SPC-630counting board. The FWHM of the instrument response function was 150 ps.

Shutter Apparatus. The sample can interact with the laser and/or a white ac-tinic light source (Schott KL1500 LCD), both of whose access to the sample isgated by a controllable shutter. The path of fluorescence photons to the de-tector is also gated by a shutter. The shutters were constructed using a rotarysolenoid (LEDEX; part no. 810-282-330) which was connected to a 12-V powersupply (Circuit Specialists, Inc.; 3645A). To control the shutters, a 5-V squarewave was sent from the computer through a data acquisition card (NationalInstruments; PCI-6229) to three relays (Potter & Brumfield; JWS-117-1). A 5-Vsquare wave from the computer triggers the TCPSC board to start a measure-ment and collect photons (33).

Measurement of qE Induction and Relaxation. As depicted in Fig. 2, the laserwas used to induce changes in the sample. The laser illuminated the samplefor 20 s, and time-resolved fluorescence decays were collected for 80ms every200 ms. After 20 s of laser exposure, the laser shutter was closed for a relaxa-tion time. The measurement sequence was repeated for 12 aliquots from thesame culture, and the fluorescence decay at each collection time was aver-aged over all aliquots. After the relaxation time ended, the laser shutter wasopened for 1.6 s, duringwhich eight fluorescence decays were collected every200 ms. The relaxation times used were 1, 2, 5, 7, 10, 14, 18, 22, 26, 30, 45, and60 s. Of the eight fluorescence decays collected in the 1.6 s of the relaxationmeasurement, the second to the fifth fluorescence traces were summed togive the relaxation trace. Because more counts were collected in the mea-surement of relaxation than in measurement of the induction, each relaxa-tion lifetime came from the average of two aliquots. Fluorescence decayswere generated by adding counts at each collection time over all aliquotsand over channels 6 (644 nm) to 12 (716 nm). Time resolution was reducedto 12.2 ps per bin by summing counts every four bins of the raw data.Measurement of changes in the lifetime due to a white actinic light sourceis described in SI Text. The measurement of the fluorescence lifetime of light-adapted and nigericin-treated algae is described in Fig. S3. The data wereanalyzed as described in the main text. A more detailed description of thedata analysis is in SI Text.

ACKNOWLEDGMENTS. K.A. would like to thank Tae Kyu Ahn, Graham Peers,and Thuy Truong for helpful discussions. This work was supported by theDirector, Office of Science, Office of Basic Energy Sciences, of the US Depart-ment of Energy under Contract DE-AC02-05CH11231 and the Division of Che-mical Sciences, Geosciences, and Biosciences, Office of Basic Energy Sciencesof the US Department of Energy through Grants DE-AC03-76SF000098 andFWP 449B. K.A. was supported by a National Science Foundation GraduateResearch Fellowship and by the University of California, Berkeley ChemicalBiology Graduate Program Training Grant 1 T32 GMO66698. J.Z. was partiallysupported by a Chancellor’s Fellowship from University of California,Berkeley.

Amarnath et al. PNAS ∣ May 29, 2012 ∣ vol. 109 ∣ no. 22 ∣ 8409

CHEM

ISTR

YPL

ANTBIOLO

GY

Dow

nloa

ded

by g

uest

on

May

24,

202

0

Page 6: Fluorescence lifetime snapshots reveal two rapidly reversible … · 2012-10-29 · Fluorescence lifetime snapshots reveal two rapidly reversible mechanisms of photoprotection in

1. de Bianchi S, Ballottari M, Dall’Osto L, Bassi R (2010) Regulation of plant light harvest-ing by thermal dissipation of excess energy. Biochem Soc Trans 38:651–660.

2. Ruban AV, Johnson MP, Duffy CDP (2012) The photoprotective molecular switch in thephotosystem II antenna. Biochim Biophys Acta 1817:167–181.

3. Blankenship R (2002) Molecular Mechanisms of Photosynthesis (Blackwell Science,Oxford, Malden, MA).

4. van Amerongen H, Valkunas L, van Grondelle R (2002) Photosynthetic Excitons (WorldScientific, River Edge, NJ).

5. Li Z, Wakao S, Fischer BB, Niyogi KK (2009) Sensing and responding to excess light.Annu Rev Plant Biol 60:239–260.

6. Raven JA (2011) The cost of photoinhibition. Physiol Plant 142:87–104.7. Gilmore AM, Hazlett TL, Govindjee (1995) Xanthophyll cycle-dependent quenching of

photosystem-II chlorophyll-a fluorescence: Formation of a quenching complex with ashort fluorescence lifetime. Proc Natl Acad Sci USA 92:2273–2277.

8. Miloslavina Y, et al. (2009) Ultrafast fluorescence study on the location andmechanismof nonphotochemical quenching in diatoms. Biochim Biophys Acta 1787:1189–1197.

9. Holzwarth AR, Miloslavina Y, Nilkens M (2009) Identification of two quenching sitesactive in the regulation of photosynthetic light-harvesting studied by time-resolvedfluorescence. Chem Phys Lett 483:262–267.

10. Johnson MP, Ruban AV (2009) Photoprotective energy dissipation in higher plantsinvolves alteration of the excited state energy of the emitting chlorophylls in the lightharvesting antenna II (LHCII). J Biol Chem 284:23592–23601.

11. Baker NR (2008) Chlorophyll fluorescence: A probe of photosynthesis in vivo.Annu RevPlant Biol 59:89–113.

12. Holt NE, et al. (2005) Carotenoid cation formation and the regulation of photo-synthetic light harvesting. Science 307:433–436.

13. Schansker G, Tóth SZ, Kovács L, Holzwarth AR, Garab G (2011) Evidence for a fluor-escence yield change driven by a light-induced conformational change withinphotosystem II during the fast chlorophyll a fluorescence rise. Biochem Biophys Acta1807:1032–1043.

14. Valkunas L, Trinkunas G, Chmeliov J, Ruban AV (2009) Modeling of exciton quenchingin photosystem II. Phys Chem Chem Phys 11:7576–7584.

15. Owens TG, Webb SP, Mets L, Alberte RS, Fleming GR (1989) Antenna structure andexcitation dynamics in photosystem I. II. Studies with mutants of Chlamydomonasreinhardtii lacking photosystem II. Biophys J 56:95–106.

16. Ihalainen JA, et al. (2005) Kinetics of excitation trapping in intact photosystem Iof Chlamydomonas reinhardtii and Arabidopsis thaliana. Biochim Biophys Acta1706:267–275.

17. Iwai M, Yokono M, Inada N, Minagawa J (2010) Live-cell imaging of photosystem IIantenna dissociation during state transitions. Proc Natl Acad Sci USA 107:2337–2342.

18. Avenson TJ, et al. (2008) Zeaxanthin radical cation formation in minor light-harvestingcomplexes of higher plant antenna. J Biol Chem 283:3550–3558.

19. Ahn TK, et al. (2008) Architecture of a charge-transfer state regulating light harvestingin a plant antenna protein. Science 320:794–797.

20. Pascal AA, et al. (2005) Molecular basis of photoprotection and control of photo-synthetic light-harvesting. Nature 436:134–137.

21. Ruban AV, et al. (2007) Identification of a mechanism of photoprotective energydissipation in higher plants. Nature 450:575–578.

22. Bonente G, et al. (2011) Analysis of LhcSR3, a protein essential for feedback de-excita-tion in the green alga Chlamydomonas reinhardtii. PLoS Biol 9:e1000577.

23. Holub O, et al. (2007) Fluorescence lifetime imaging microscopy of Chlamydomonasreinhardtii: Nonphotochemical quenching mutants and the effect of photosyntheticinhibitors on the slow chlorophyll fluorescence transient. J Microsc 226:90–120.

24. Lemeille S, Rochaix JD (2010) State transitions at the crossroad of thylakoid signalingpathways. Photosynth Res 106:33–46.

25. Cheng YC, et al. (2008) Kinetic modeling of charge-transfer quenching in the CP29minor complex. J Phys Chem B 112:13418–13423.

26. Peers G, et al. (2009) An ancient light-harvesting protein is critical for the regulation ofalgal photosynthesis. Nature 462:518–521.

27. Niyogi KK, Bjorkman O, Grossman AR (1997) Chlamydomonas xanthophyll cyclemutants identified by video imaging of chlorophyll fluorescence quenching. Plant Cell9:1369–1380.

28. Niyogi KK, Bjorkman O, Grossman AR (1997) The roles of specific xanthophylls inphotoprotection. Proc Natl Acad Sci USA 94:14162–14167.

29. Valkunas L, et al. (2011) Excitation migration, quenching, and regulation of photosyn-thetic light harvesting in photosystem II. J Phys Chem B 115:9252–9260.

30. Caffarri S, Broess K, Croce R, van Amerongen H (2011) Excitation energy transferand trapping in higher plant photosystem II complexes with different antenna sizes.Biophys J 100:2094–2103.

31. Betterle N, et al. (2009) Light-induced dissociation of an antenna hetero-oligomer isneeded for nonphotochemical quenching induction. J Biol Chem 284:15255–15266.

32. Johnson MP, et al. (2011) Photoprotective energy dissipation involves the reorganiza-tion of photosystem II light-harvesting complexes in the grana membranes of spinachchloroplasts. Plant Cell 23:1468–1479.

33. Becker W The bh TCSPC Handbook. 4th Ed, http://www.becker-hickl.de/handbook.htm.

8410 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1205303109 Amarnath et al.

Dow

nloa

ded

by g

uest

on

May

24,

202

0