Skip to main content
PLOS One logoLink to PLOS One
. 2021 Dec 9;16(12):e0256410. doi: 10.1371/journal.pone.0256410

Irradiance and nutrient-dependent effects on photosynthetic electron transport in Arctic phytoplankton: A comparison of two chlorophyll fluorescence-based approaches to derive primary photochemistry

Yayla Sezginer 1,*, David J Suggett 2, Robert W Izett 1, Philippe D Tortell 1,3
Editor: Matheus C Carvalho4
PMCID: PMC8659313  PMID: 34882695

Abstract

We employed Fast Repetition Rate fluorometry for high-resolution mapping of marine phytoplankton photophysiology and primary photochemistry in the Lancaster Sound and Barrow Strait regions of the Canadian Arctic Archipelago in the summer of 2019. Continuous ship-board analysis of chlorophyll a variable fluorescence demonstrated relatively low photochemical efficiency over most of the cruise-track, with the exception of localized regions within Barrow Strait, where there was increased vertical mixing and proximity to land-based nutrient sources. Along the full transect, we observed strong non-photochemical quenching of chlorophyll fluorescence, with relaxation times longer than the 5-minute period used for dark acclimation. Such long-term quenching effects complicate continuous underway acquisition of fluorescence amplitude-based estimates of photosynthetic electron transport rates, which rely on dark acclimation of samples. As an alternative, we employed a new algorithm to derive electron transport rates based on analysis of fluorescence relaxation kinetics, which does not require dark acclimation. Direct comparison of kinetics- and amplitude-based electron transport rate measurements demonstrated that kinetic-based estimates were, on average, 2-fold higher than amplitude-based values. The magnitude of decoupling between the two electron transport rate estimates increased in association with photophysiological diagnostics of nutrient stress. Discrepancies between electron transport rate estimates likely resulted from the use of different photophysiological parameters to derive the kinetics- and amplitude-based algorithms, and choice of numerical model used to fit variable fluorescence curves and analyze fluorescence kinetics under actinic light. Our results highlight environmental and methodological influences on fluorescence-based photochemistry estimates, and prompt discussion of best-practices for future underway fluorescence-based efforts to monitor phytoplankton photosynthesis.

Introduction

Phytoplankton productivity in polar marine waters is constrained by nutrient and light availability, which fluctuate dramatically across seasonal cycles and shorter time and space scales [1]. In late summer, when sea ice cover is at a minimum and the mixed layer is shallow and highly stratified, phytoplankton are exposed to high solar irradiance and low nutrient concentrations [2]. Under these conditions, the growth and photosynthetic efficiency of Arctic phytoplankton becomes nitrogen-limited [35]. However, localized regions of elevated productivity can persist where various processes transport nutrients into the mixed layer, including upwelling, tidal mixing, and freshwater input from rivers and glaciers [68]. Modelling studies suggest that Barrow Strait in the Canadian Arctic Archipelago is one such productivity hotspot, with strong tidal currents and shallow sills driving vertical mixing in a region where Pacific and Atlantic-derived water masses converge [9, 10]. Additionally, Barrow Strait receives glacial and land-derived nutrients from the Cornwallis and Devon Island rivers [11, 12]. Rapid climate change in the Arctic is expected to have complex effects on these nutrient delivery mechanisms through the intensification of coastal erosion [13], increasing river inputs [14] and reduced vertical mixing due to intensifying stratification [15, 16]. At present, it is unclear how phytoplankton productivity will respond to these anticipated perturbations.

Assessing phytoplankton productivity in physically-dynamic marine waters requires high spatial resolution measurements that cannot be obtained from traditional discrete bottle incubation methods, such as 14C uptake experiments. For this reason, oceanographic field studies have increasingly employed continuous sampling of surface water properties using variable chlorophyll a (Chla) fluorescence from Fast Repetition Rate Fluorometry (FRRf) and other related methods to rapidly and autonomously assess phytoplankton photophysiology and primary photochemistry as a proxy for primary productivity (e.g. [1721]). Such variable fluorescence techniques rely on the inverse relationship between Chla fluorescence and photochemistry. These processes, along with heat dissipation, comprise the three energy dissipation pathways for light energy absorbed within Photosystem II (PSII) [22]. Fast Repetition Rate Fluorometry operates by supplying rapid excitation light pulses to progressively saturate the photosynthetic pathway and simultaneously induce a measurable Chla fluorescence response—often referred to as a fluorescence transient [23]. Analysis of Chla fluorescence transients provides information on the photochemical efficiency and functional absorption cross section of PSII, as well as estimates of the turnover rate of photosynthetic electron transport chain molecules [24, 25]. Fast Repetition Rate fluorescence transients can be obtained nearly instantaneously, offering opportunities for very high temporal and spatial resolution measurements.

Primary photochemistry is typically estimated from variable Chla fluorescence measurements by calculating the rate of photosynthetic electron transport out of PSII (ETRPSII). There are several algorithms that may be applied to derive ETRPSII [17, 2630]. Each algorithm relies on the same principles of light harvesting and photosynthetic electron transport, but arrive at ETRPSII estimates using slightly different, but theoretically equivalent, combinations of photophysiological metrics. As a result, different algorithms confer different field-sampling advantages and challenges (see [31]). The so-called ‘amplitude’ based approach (abbreviated ETRa; sometimes also referred to as the sigma-algorithm), calculates ETRPSII as the product of photosynthetically active radiation (PAR), the functional absorption cross section of PSII in the dark-acclimated state (σPSII), and the photochemical efficiency of PSII normalized by the dark-acclimated maximum photochemical efficiency of PSII [27]. This approach reduces uncertainty in ETRPSII estimates by using σPSII measurements made in the dark-acclimated state, which are subject to less noise than σPSII’ measurements made in the light [31]. As a result, ETRa is a favorable approach in low biomass waters where low signal-noise measurements represent a considerable challenge.

To achieve dark acclimation for ETRa measurements, phytoplankton samples are kept in darkness or very low light to relax non-photochemical quenching processes (NPQ), which upregulate heat dissipation of absorbed light energy, thereby reducing photochemistry and fluorescence yields [32]. Optimal dark acclimation times vary between phytoplankton species and depend to a large extent on the environmental history of samples [33], making it challenging to design widely applicable field protocols for high resolution data acquisition. In practice, applied dark-acclimation periods range from 5–30 minutes (e.g. [34, 35]). Many FRRf field deployments have focused on discrete sample analysis, applying extended dark acclimation periods to ensure samples reach a short-term steady state condition [25, 3638]. Such discrete sample analysis enables standardized measurements and characterization of light-dependent physiological responses, at the cost of significantly reduced spatial and temporal measurement resolution. In contrast, continuous underway flow-through data acquisition yields high-resolution, real-time measurements of phytoplankton photophysiology, but creates uncertainty in the light exposure history of phytoplankton transiting through a ship’s seawater supply lines. As a result, samples analyzed in continuous mode are neither fully representative of in-situ photophysiology or fully dark-acclimated states, and thus incompatible with the ETRa algorithm, which requires both dark- and light-acclimated measurements.

Recently, a new fluorescence approach has been developed to derive ETRPSII based on the turnover rate of the primary electron acceptor molecule (Qa) within the photosynthetic electron transport chain [30]. In this ‘kinetic’ approach, Qa turnover rates are derived from analysis of fluorescence relaxation time-constants, resulting in the term ETRk. The derivation of ETRk does not depend on dark-acclimated measurements, and can thus significantly increase the frequency of ship-board ETRPSII measurements. The kinetic fluorescence approach for ETRk was originally developed and implemented in mini-Fluorescence Induction-Relaxation (mini-FIRe) instruments [30], which use similar data acquisition protocols, but a different numerical approach for fluorescence relaxation analysis than FRRf. Gorbunov et al. [30] observed greater coherence between growth rates of laboratory cultures and FIRe-derived ETRk compared to alternative ETRPSII estimates, suggesting a strong potential for ETRk to quantify in-situ primary photochemistry. To our knowledge, there have been very few direct comparisons of FRRf-derived ETRk and ETRa estimates for natural phytoplankton assemblages.

In this article, we examine the relationship between ETRk and ETRa estimates across a range of hydrographic regimes in the Canadian Arctic Ocean. We employed a hybrid approach to data collection along a ship-track through Lancaster Sound and Barrow Strait, combining semi-continuous flow-through measurements with light response curves on static samples. Continuous sampling enabled us to obtain high-resolution measurements and examine the effects of light history and nutrient status on the physiology of Arctic phytoplankton assemblages. Data from rapid light-response curves allowed direct comparison of FRRf-derived ETRk and ETRa estimates. Our results demonstrate residual light-dependent NPQ effects on dark (low light) sample measurements, and a decoupling of ETRk and ETRa under conditions of phytoplankton photophysiological stress. We relate the spatial patterns in our observations to regional and fine-scale patterns in hydrography and nutrient supply in the eastern Canadian Arctic Archipelago, and discuss the potential effects of different data analysis approaches on ETRPSII-based primary photochemistry estimates. Results from our work will inform future ship-based deployments of FRRf and related techniques to understand spatial patterns in phytoplankton productivity.

Materials and methods

Underway sampling

Arctic Ocean samples and hydrographic data were collected aboard the CCGS Amundsen from August 10–15, 2019, within the eastern region of Lancaster Sound and Barrow Strait, in the waters surrounding Devon and Cornwallis Islands. Access to these sampling regions was granted by the Nunavut Research Institute Scientific Research License (0501119R-M). All FRRf measurements were obtained with a LIFT (Light Induced Fluorescence Transient)-FRR fluorometer (LIFT-FRRf; Soliense Inc.). Water samples were delivered to the LIFT-FRRf sampling cuvette using the ship’s seawater line as a primary supply (BC-4C-MD pump, March MFG Inc, nominal flow rate ∼20 L min-1), combined with two secondary peristaltic pumps. The first pump (Masterflex L/S, model 7518–10), was used to create a continuous sampling loop (∼200 mL min-1) that was connected via t-fitting to a custom-built peristaltic pump actuated by the LIFT-FRRf software. The FRRf-actuated pump enabled precise synchronization of the sample handling and fluorescence measurements, allowing us to employ a semi-continuous sampling strategy of alternating fluorescence transients and light response curve measurements (details below).

In parallel with FRRf measurements, in-situ Chla fluorescence, surface water salinity and temperature were measured using a flow-through Seabird thermosalinograph system (SBE 38), equipped with a WETStar fluorometer (WET Labs). The underway Chla fluorescence signal was calibrated against discrete samples collected from surface Niskin bottles to approximate the along-track Chla biomass (mg m-3). Surface PAR measurements were obtained from an QCR-22000 Biospherical Instruments probe mounted above the ship’s super-structure. Biological oxygen saturation, ΔO2/Ar, was measured as a metric of net community production using a Hiden Analytical quadrupole membrane inlet mass spectrometer (MIMS; HAL20) following the approaches outlined by Tortell et al. [39, 40]. These gas measurements were made continuously on seawater obtained from the same underway lines that supplied the FRRf system. Briefly, seawater was circulated at a constant flow rate past the mass spectrometer’s inlet cuvette consisting of a 0.18 mm thick silicone membrane. Measurements of the mass-to-charge ratios at 32 (O2) and 40 (Ar) atomic mass units were obtained at approximately 20 s. intervals. Air standards, consisting of filtered seawater (<0.2 μm) incubated at ambient sea surface temperature and gently bubbled using an aquarium air pump, were run periodically by automatically switching the inflow water source every 45–90 minutes. For both seawater and air standard measurements, the inflowing water passed through a 6-m heat exchange coil immersed in a constant-temperature 4°C water bath before passing the MIMS inlet. The seawater and air standard O2/Ar ratios ([O2/Ar]sw and [O2/Ar]std, respectively) were used to derive underway ΔO2/Ar by linearly interpolating between air standard measurements.

In addition to continuous surface water sampling, water column hydrographic profiles were examined using CTD casts to a depth of 200m at 16 stations (Fig 1). Mixed layer depths were derived from a density difference criterion of 0.125 kg m-3 from surface water values. The CTD and underway surface salinity, temperature, Chla, and PAR data were provided by the Amundsen Science group of Université Laval [4244].

Fig 1. Study region map.

Fig 1

The inset globe shows the geographic region of interest outlined in blue. The detailed map shows the numbered CTD stations sampled in August 2019, overlayed on bathymetric contours. We use the 200m bathymetric contour line around 92°W to separate Lancaster Sound and Barrow Strait. The ship track is shown in red. The complete list of geographic coordinates sampled is available in the Polar Data Catalogue (doi: 10.5884/12715). The map was produced in Matlab (R2020a) using the publicly available m_map package [41].

LIFT-FRRf sampling protocols & parameter retrieval

Chlorophyll a fluorescence transients were obtained using a Single Turnover (ST) flash protocol and fit to the biophysical model of Kolber et al. [23] to derive photosynthetic parameters under dark acclimation and under actinic light, where the latter is denoted with the’ notation. A summary of parameters measured and definitions is given in Table 1. Specifically, we derived estimates of the functional absorption cross section of PSII (σPSII, σPSII’), the minimum and maximum fluorescence when the reaction center pool is fully open and closed (Fo, F’ and Fm, Fm’, respectively), and the variable fluorescence (Fv = [FmFo]; Fq=[FmF]). Single Turnover excitation flashlets were delivered simultaneously by 445, 470, 505, 535, and 590 nm LED lamps. Each curve fit and acquisition was based on fluorescence yields averaged from five sequential ST flashes, each with a total duration of 0.2 s. Each 5 flash sequence was separated by a 1s interval. Curve fitting and fluorescence averaging was completed within LIFT software (Soliense Inc.).

Table 1. Commonly referred to photophysiological terms and abbreviations.

Term Definition Units
Chla Chlorophyll a fluorescence Relative units
Fo, F Minimum Chla measured in dark and light acclimated state, respectively. Relative units
Fm, Fm Maximum Chla measured in dark and light acclimated state, respectively. Relative units
σPSII, σPSII Functional absorption cross-section for PSII in the dark and light acclimated state, respectively Å2PSII−1
Fv/Fm Photochemical yield in the dark acclimated state; (FmFo)/Fm Dimensionless
Fq/Fm150 Photochemical yield measured under 150 μmol quanta m−2 s−1 Dimensionless
Fq/FmEmax Photochemical yield measured under super saturating irradiance Dimensionless
NPQ Non-photochemical quenching; (FmFm)/Fm Dimensionless
τ Qa Photosynthetic electron turnover rate measured at saturating irradiance s−1
p Probability of energy transfer between RCIIs Unitless
CQa(t) Fraction of initially available RCIIs closed by excitation flashlets Unitless
PAR Photosynthetically active radiation μmol quanta m−2 s−1
E(t,λ) Transient PAR provided by excitation flashlets μmol quanta m−2 s−1
ETRa Amplitude-based electron transport rate e- s-1 RCII-1
ETRk Kinetic-based electron transport rate e- s-1 RCII-1

Each seawater sample was initially held in the cuvette under low light for five minutes to allow NPQ relaxation. During this low-light acclimation period, five LEDs at peak wavelength excitation of 445, 470, 505, 535, and 590 nm each supplied 1 μmol quanta m-2 s-1 of actinic light, providing a total irradiance of 5 μmol quanta m-2 s-1. Samples were acclimated to low light rather than total darkness to avoid RCII closure and fluorescence quenching associated with back flow of electrons from the PQ pool to Qa (e.g. [45]). After this short acclimation period, 40 acquisitions (each consisting of 5 averaged fluorescence transients) were collected. Irradiance was then increased to 150 μmol quanta m-2 s-1 (30 μmol quanta m-2 s-1 per LED) and samples were held for an additional five minutes to acclimate to the higher irradiance level before collecting another 40 acquisitions to measure photophysiology close to the saturation irradiance for primary photochemistry, previously found to range from 96–213 μmol quanta m-2 s-1 for surface assemblages in the Canadian Arctic during summer [46]. The LIFT-FRRf cuvette was then flushed for 60 s with seawater from the ship seawater supply, displacing at least five full cuvette volumes before the pump was turned off to isolate the next sample for measurements.

Semi-continuous sampling was interrupted every 90 mins to perform light response curves on static samples. For each light response curve, the actinic irradiance supplied by each LED lamp increased incrementally from 0 to 350 μmol quanta m-2 s-1 for all 5 LEDs to create light steps of 0, 15, 35, 75, 110, 150, 200, 300, 400, 550, 850, 1250, and 1750 μmol quanta m-2 s-1 total actinic light. A total of 25 acquisitions (of 5 averaged fluorescence transients each), were obtained at each light step with a pause of 30s at each new light level to provide some acclimation time. Fluorescence amplitudes generally stabilized by the third round of data acquisition, and we thus excluded the first two flash sequences at each light level from data analysis. Light response curves were used to calculate ETRPSII as a function of increasing light intensity (see below for calculation details). To produce final photosynthesis-irradiance curves, ETRPSII was plotted against irradiance and fit to the hyperbolic function described by Webb et al. [47], using a least squares non-linear regression to derive the maximum light-saturated photosynthetic rate (ETRmax), the saturating light intensity (Ek) and the light utilization efficiency (α).

Daily blank corrections were performed by analyzing seawater gently passed through a 0.2 μm syringe filter (e.g. [25]). Soliense software was programmed to subtract the resulting fluorescence intensity from all measurements. Prior to the field deployment, the LIFT-FRRf lamps used to produce actinic light and the probing flashes were calibrated using a Walz spherical submersible micro quantum sensor (Walz, US-SQS-L).

Electron Transport Rates (ETRPSII)

The amplitude and kinetics-based ETRPSII algorithms (ETRa and ETRk, both e- s-1 RCII-1) were applied to fluorescence data collected during light response curves to determine ETRPSII at increasing actinic light intensities. ETRa was applied to all 85 light response curves collected along the ship-track to produce photosynthesis-irradiance curves following Webb et al. [47]. At each light level, ETRa (e- RCII-1 s-1) was determined as the quantity of incident light absorbed by PSII directed towards photochemistry [27, 28];

ETRa=PAR×σPSII×(Fq/Fm)/(Fv/Fm)×6.022×103, (1)

Here, PAR (μmol quanta m-2 s-1) is the total actinic light provided by the FRRf LEDs, σPSII2 PSII-1) is the PSII functional absorption cross section measured in dark acclimated samples, and Fq/Fm divided by Fv/Fm (dimensionless) is the PSII photochemical efficiency measured under actinic light normalized by the dark-measured maximum photochemical efficiency. The constant 6.022*10-3 converts σPSII units from Å2 PSII-1 to m2 PSII-1 and PAR from μmol quanta to quanta.

For ETRk calculations, fluorescence data from a subsample of 25 light response-curves with optimal signal-to-noise ratios were re-analyzed using a 3-component multi-exponential model to describe Qa reoxidation kinetics [23]. This numerical procedure was applied as a fitting option in the FRRf Soliense Software.

F(t)=Fo+(FmFo)CQa(t)1p1CQa(t)p, (2)
CQat=E(t,λ)×σPSII×1CQa(t)1CQa(t)pCQa(t)i=13αi/τi, (3)

Here, F(t) represents the measured fluorescence signal, CQa(t) (dimensionless) is the fraction of available PSII reaction centers closed by excitation flashlets, with CQa(t) equal to 1 when the photochemistry pathway is fully closed. The term p (dimensionless) is the connectivity factor, which describes the likelihood of energy transfer between RCIIs. αi and τi refer to the amplitude and time constant of the ith component of Qa reoxidation, respectively. The value CQat is determined by the balance between primary photochemistry induced by excitation flashlets (E; μmol quanta m-2 s-1) and electron transfer from Qa to secondary electron acceptor, Qb, mediated by three kinetic components. Fluorescence transients were fit to retrieve photophysiological parameters by integrating Eq 3 over the length of the ST flash sequence and then iteratively fitting Eq 2. to the fluorescence data. Importantly, we note Eq 3 differs slightly from that applied by FIRe-based data analysis, in which there is an added term to describe RCII closure induced by actinic irradiances (See ‘Computational Considerations’ in the Discussion).

The rate of Qa reoxidation, τQa, was calculated by averaging the time constants of the two primary components of electron transfer from Qa to secondary electron acceptor Qb following the approach of Gorbunov and Falkowski [30]. The kinetic-based ETRPSII was then calculated as,

ETRk=1τPAR×Fq/FmPARmax×Fq/FmEmax, (4)

Here, PARmax is a super-saturating light level chosen as a value three–fold higher than the light saturation parameter, Ek, derived from photosynthesis irradiance curves. Fq/FmEmax is the PSII photochemical efficiency measured under PARmax. The photosynthetic turnover rate, 1τ (s-1), is taken as 1τQa at saturating irradiance where primary photochemistry is at a maximum [30].

Non-Photochemical Quenching (NPQ)

We quantified NPQ as a measure of the relative increase in heat dissipation of absorbed energy by PSII between samples exposed to low light and 150 μmol quanta m-2 s-1, following Bilger and Bjorkman [48] as:

NPQSV=(FmFm)/Fm, (5)

This derivation assumes full relaxation of all NPQ processes in the dark-acclimated state. However, as discussed in the Results section, this assumption did not hold in our samples. For this reason, we used an additional measure of long-term NPQ processes, based on the ratio of photochemical efficiency measured under low and high light (i.e. (Fq/Fm)/(Fv/Fm)).

Statistical analysis

Pair-wise statistical relationships between measured variables were determined using Spearman Rank correlation tests. Samples from two regions of our transect (Lancaster Sound and Barrow Strait) were compared using Kruskal-Wallis test. Lilliefors test rejected the null hypothesis that underway data were normally distributed, so we report median rather than mean values of all photophysiological variables. Deviation from the median was determined as the median absolute deviation. All analyses were completed using Matlab (Mathworks, R2020a).

Results

Hydrographic properties

Sea surface temperature (SST) and salinity varied significantly across our study region, reflecting the influence of different water masses and freshwater inputs. Sea surface temperature ranged from 0.6 to 15.5°C, and salinity ranged from 22.8 to 31.6 PSU. Salinity and SST strongly covaried, and exhibited a number of sharp transitions across prominent hydrographic fronts associated with freshwater input (Fig 2; S1 Table). Surface layer (∼7m) phytoplankton biomass, approximated by in-situ Chla fluorescence measurements (non-FRRf), was low throughout the entire transect, with a mean Chla fluorescence equivalent to 0.15 ± 0.04 mg m-3 (n = 7200). These in situ Chla measurements showed significant diel periodicity, likely reflecting daytime fluorescence quenching effects. Both SST and salinity exhibited weak correlations with Chla (S1 Table). Mixed layer depths recorded at the 16 profiling stations were shallow, with a mean of 9.2 ± 4.5 m. Mixed layer nitrate and nitrite concentrations at profiling stations were frequently below the detection limit (S2 Table, mean 0.03 ± 0.049 μM, n = 16), indicative of post-bloom, nutrient-limited summer conditions. The mixed layer Chla concentration averaged across all 16 profiling stations was 0.40 ± 0.18 mg m-3 (n = 16). There was no statistically significant relationship between station Chla and mixed layer nitrate and nitrite concentrations (ρ = 0.09, p = 0.75, n = 16). Biological oxygen saturation, derived from ΔO2/Ar measurements, was greater than zero across the entire cruise track, implying net autotrophic conditions. Small-scale features in ΔO2/Ar distributions were observed across hydrographic frontal regions with only weak correlation to salinity or temperature measured along the cruise track (S1 Table).

Fig 2. Summary of oceanographic conditions.

Fig 2

Hydrographic conditions, phytoplankton biomass and biological oxygen saturation measured in surface waters (∼ 7 m depth) along the cruise track. Shaded area on Aug. 13–14 marks measurements made in Barrow Strait.

Photophysiological measurements

Continuous flow-through measurements

The photophysiological parameters Fv/Fm and σPSII fluctuated over diel cycles throughout our sampling period (Fig 3a and 3b), showing daily maxima during the night, and decreasing during daylight hours. Both of these variables exhibited a significant negative correlation with actinic surface PAR intensity averaged over the 5 min window prior to sample measurements (Table 2; Fig 4a and 4b). This result is consistent with previously noted in-situ photophysiological diurnal patterns of daytime fluorescence quenching [49, 50]. We note, however, that these Fv/Fm and σPSII measurements were collected after 5 minutes of low light exposure, indicating the persistence of longer-lived light-dependent quenching effects after this acclimation period. We thus conclude that dark-acclimation (i.e. NPQ relaxation) was not achieved in our measurement protocol, and that Fv/Fm and σPSII are thus representative of an intermediate state between in-situ and dark-acclimated values.

Fig 3. Summary of photophysiological conditions.

Fig 3

Semi-continuous FRRf measurements of photophysiology (black dots) are superimposed over the in-situ surface PAR (blue line). A loess smoothing function was applied to photo-physiological measurements (red line). Shading denotes the Barrow Strait portion of the transect.

Table 2. Correlation of underway photophysiological variables and surface PAR.
Photophysiological variables Surface PAR
Lancaster Sound
n = 283
Barrow Strait
n = 198
Total
n = 481
Fv/Fm ρ = -0.72** ρ = 0.27** ρ = -0.72**
σ PSII ρ = -0.79** ρ = -0.60** ρ = -0.50**
NPQ ρ = -0.66** ρ = -0.65** ρ = -0.65**
Fq/Fm150:Fv/Fm ρ = 0.81** ρ = 0.66** ρ = 0.68**
Fq/Fm150 ρ = 0.21** ρ = 0.67** ρ = -0.01

Spearman rank correlations between underway photophysiological measurements and recent surface PAR exposure, analyzed by region and for the total dataset.

** is used to indicate p values ≤ 0.01.

Fig 4. Recent light history effects on photophysiology measured under low light.

Fig 4

FRRf-derived photophysiological parameters measured under low light are plotted against in-situ surface PAR at the time of sample acquisition. Panels (a) and (b) show Fv/Fm and σPSII measurements made after 5 min of low light exposure. Panels (c) and (d) show NPQ and residual quenching measured after low light treatment. Lines of best fit are shown for Barrow Strait (solid line) and Lancaster Sound (dashed line). Full regression analyses results are reported in S2 Table. Error bars show standard error, but are often concealed by size of data symbols.

Despite signs of strong light-dependent quenching, NPQ values, calculated as NPQSV = (Fm − Fm’)/Fm, were negatively correlated to surface PAR (Fig 4c). This surprising result can be explained by the derivation used here for NPQ, which measures the fractional change in NPQ between light and dark (low light) measurements, rather than total NPQ. As a result, this approach does not account for any residual quenching present in samples after five minutes of low light acclimation. To address this limitation, we used the ratio (Fq/Fm150)/(Fv/Fm) to estimate the extent of residual quenching present in samples after five minutes of NPQ relaxation. As expected, this derived variable was well correlated to surface PAR (Fig 4d, Table 2), reflecting the effects of residual quenching and, potentially, short-term photoinhibition.

To further examine low light-acclimated Fv/Fm and σPSII values, we isolated night-time Fv/Fm and σPSII measurements collected under relatively low ambient surface PAR. Due to the long summer daylight hours in the Arctic, only 1% of data points represents true night when surface PAR = 0. We thus chose light levels ≤ 100 μmol quanta m-2 s-1 to represent night-time conditions. By comparison, midday surface PAR ranged from 400–1000 μmol quanta m-2 s-1. The median night-time Fv/Fm for the entire transect was 0.36 ± 0.03 (n = 214), a value similar to the global average of 0.35 ± 0.11 [51], but lower than the 0.55 median value previously recorded for late-summer assemblages in the Canadian Arctic [52]. The lowest Fv/Fm values were recorded at the beginning of the ship track (August 10–12) within Lancaster Sound (0.32 ± 0.03, n = 120), while Fv/Fm increased significantly in Barrow Strait (0.40 ± 0.02, n = 94), indicating greater photosynthetic potential in this region (Fig 3b). In contrast, night-time σPSII did not significantly vary between Lancaster Sound (250 ± 17.2) and Barrow Strait (241.1 ± 15.3) (Fig 3a). Note that these absolute σPSII values are somewhat lower than those reported in previous studies, likely reflecting our use of simultaneous excitation flashlets centered around 445, 470, 505, 535, and 590 nm. Relative to blue light, not all of these wavelengths are efficiently absorbed by phytoplankton, resulting in an apparent decrease σPSII [53]. As discussed below, σPSII values are also subject to physiological, taxonomic and environmental effects [54].

As observed for Fv/Fm, the photosynthetic efficiency measured under 150 μmol quanta m-2 s-1 actinic light (Fq/Fm150) displayed strong regional differences, increasing from 0.21 ± 0.01 (n = 283) in Lancaster Sound to 0.30 ± 0.02 (n = 198) in Barrow Strait. However, unlike the low-light measurements, Fq/Fm150 did not exhibit a diel signature and was remarkably consistent across samples measured within Lancaster Sound, where Fq/Fm150 was weakly correlated to surface PAR (Table 2; Figs 3e and 5). Whereas, Fq/Fm150 in Barrow Strait was strongly positively correlated to surface PAR (Table 2). Such regional differences in the Fq/Fm150, surface PAR relationship imply differential capacities of Lancaster Sound and Barrow Strait phytoplankton to photoacclimate to higher irradiances.

Fig 5. Recent light history effects on photophysiology measured under low light.

Fig 5

The PSII photochemical efficiency measured under 150 μmol quanta m-2 s-1 in relation to natural surface irradiance at the time of sample acquisition. Lines of best fit are shown for Barrow Strait (solid line) and Lancaster Sound (dashed line). Full regression analyses results are reported in S2 Table. Error bars show standard error, but are often concealed by size of data symbols.

Photosynthesis-irradiance curves and ETRPSII comparisons

We used light response curves to compare FRRf-based ETRa and ETRk estimates. In this approach, ETRa (Eq 1) was plotted against actinic irradiance (Fig 6) to derive the maximum rate of charge separation at RCII (ETRmax), the light-dependent increase in the charge separation rates (α), and the saturating light intensity (Ek). Fit parameters from these curves varied considerably, with mean values of ETRmax, α, and Ek of 460 ± 345 e- s-1 RCII-1, 1.37 ± 0.57 e- RCII-1 quanta-1 m-2, and 358.0 ± 199.0 μmol quanta m-2 s-1, respectively (n = 85).

Fig 6. Photosynthesis-irradiance curves derived by ETRa and ETRk.

Fig 6

Consolidated mean ETRa (red) and ETRk (black) estimates at each light step of the 25 reprocessed P-I curves. Error bars represent the standard error of all individual measurements from all curves at each light step. Curves were produced using the photosynthesis-irradiance function described by Webb et al. (1974).

A subsample of 25 light response curves was re-analyzed using the ‘kinetic’ approach (ETRk; Eq 5), and compared with ETRa values. This comparison revealed a strong correlation between the ETRPSII values (ρ = 0.81, p < 0.001; Fig 7). However, the kinetics-based algorithm produced consistently higher results than ETRa, with values 1.96 ± 1.2 times greater, on average, than ETRa (Fig 6).

Fig 7. ETRk plotted against ETRa derived from photosynthesis-irradiance curves and colored by Fv/Fm.

Fig 7

Each point represents the mean ETR value at a given light intensity within a Photosynthesis-Irradiance curve, and error bars are the standard error. The dashed line indicates a 1:1 relationship.

Along our cruise track, we observed a distinct spatial pattern in the ETRk:ETRa ratio. The highest values (3.26 ± 0.95, n = 14) occurred during the early part of our survey (August 10–13), with significantly lower ETRk:ETRa (1.42 ± 0.16, n = 11) observed near the end of our cruise, particularly in Barrow Strait where Fv/Fm was at its maximum (Fig 8). More generally, ETRk:ETRa displayed a strong negative correlation with Fv/Fm values (ρ = -0.60, p < 0.01, n = 25). As discussed below, this result suggests ETRk:ETRa decoupling is strongly driven by photophysiological and environmental variability.

Fig 8. Parameters contributing to the ETRk:ETRa ratio measured in Lancaster Sound (LS) versus Barrow Strait (BS).

Fig 8

Horizontal lines within each boxplot represent the median. The upper and lower edge of each box demarks upper and lower quartiles, respectively, while whiskers extend over the entire data range, excluding outliers. Outliers, determined as data points falling over 1.5x the interquartile distance away from box edges, appear as unfilled circles. p values in each subplot are results from 2-group Kruskal-Wallis tests. All data shown here was collected during the 25 reprocessed P-I curves.

Discussion

The primary focus of our work was to quantify phytoplankton photophysiology and photochemical yields along our ship-track using active Chla fluorescence methods. With this in mind, we applied a sampling and analysis strategy to support both amplitude-based and kinetic analysis of FRRf data to derive ETR estimates. In the following, we first discuss the spatial patterns in FRRf-based measurements in relation to nutrient concentrations and light histories of phytoplankton assemblages encountered along our sampling transect. We then examine potential factors leading to the uncoupling of ETRPSII estimates, including environmental and physiological variability, and potential influences of different data analysis methods. We conclude by discussing the implications of our results for future ship-board FRRf deployments.

Spatial variation of photophysiology

Across our survey region, we observed notable spatial patterns in FRRf data, with low Fv/Fm and Fq/Fm150 values in Lancaster Sound, and significantly higher values in Barrow Strait. Low Fv/Fm values are typical in the late-summer Arctic, and have been shown to increase in response to nitrate (but not phosphate) enrichments [5, 55]. This result, coupled with the nutrient depletion observed at profiling stations along our cruise track (Table 2), suggests that the low Fv/Fm values we observed likely reflect nitrogen deficiency. Iron (Fe)-limitation has also been shown to exert a strong negative effect on Fv/Fm, coincident with increases in σPSII values. These Fe-dependent effects result from a combination of physiological responses [36, 56, 57] and taxonomic shifts towards smaller cells [54]. In contrast, the low Fv/Fm values recorded in the Lancaster Sound were not associated with high σPSII, and correlation analysis of night-time σPSII and Fv/Fm values revealed a weak positive relationship between σPSII and Fv/Fm (ρ = 0.17, p < 0.05, n = 214). Based on these results, and the proximity of our sampling to land-based Fe sources [58], we ruled out iron limitation as a likely cause for the low photo-efficiencies observed from August 10–13 in Lancaster Sound.

Notwithstanding the low background nutrient concentrations measured at profiling stations, chemical analyses from the Canadian Arctic Archipelago Rivers Program and Canadian Arctic GEOTRACES program have revealed elevated nitrate and nitrite concentrations within several rivers that discharge into Barrow Strait (Fig 9; [11, 59]). In this region, we observed low surface water salinity and elevated Fv/Fm, suggesting a link between river input and increased photosynthetic efficiency, which we ascribe to nutrient inputs. Additionally, the greatest mixed layer depths were found at the two CTD profiling stations situated between Cornwallis Island and Devon Island (Table 2), indicative of enhanced mixing associated with strong tidal currents, in agreement with model predictions of elevated mixing in Barrow Strait [10]. These observations suggest that spatial differences observed in FRRf-derived photophysiology may reflect elevated nutrient availability in Barrow Sound, resulting from a combination of river inputs and mixing effects.

Fig 9. Spatial distribution of riverine nutrient inputs and photochemical efficiency (Fv/Fm) along the ship track.

Fig 9

River contributions of nitrate and nitrite are indicated by the size of grey bubbles. The largest inputs of nitrate and nitrite in the region are concentrated in the strait between Cornwallis and Devon Islands, coincident with observations of raised Fv/Fm values. Larger circles are used to denote night-time measurements of Fv/Fm, whereas smaller triangles denote day-time measurements.

Several mechanisms can explain the apparent effects of increased nitrogen availability on phytoplankton photophysiology. First, increased nitrogen availability enables protein synthesis needed to repair inactive reaction centers [60, 61]. Indeed, phytoplankton in Barrow Strait appeared to have an increased ability to acclimate to higher light levels, as evidenced by the strong positive relationship between Fq/Fm150 and surface PAR (ρ = 0.67, p < 0.01, n = 94; Fig 5) observed in this region. By comparison, Fq/Fm150 exhibited only a weak relationship to surface PAR in Lancaster Sound where Fq/Fm150 was significantly reduced and relatively constant (Fig 5). This result suggests that, phytoplankton in Lancaster Sound had a reduced capacity to maintain protein repair rates needed to prevent lasting photodamage at higher light levels.

In addition to direct physiological effects, localized nutrient loading may indirectly affect FRRf signatures by stimulating a shift from small to larger phytoplankton species, for instance from nano-flagellates to diatoms [55]. Unfortunately, we lack information on phytoplankton assemblage composition, and thus cannot directly examine any potential taxonomic effects on photophysiological signatures. However, such a shift from small to large cells would be expected to drive a decrease in σPSII, concurrent with increases in Fv/Fm [54], which we did not observe. Previous pigment analyses conducted in the Canadian Arctic Archipelago found that Lancaster Sound and Barrow Strait were both dominated by diatom species, followed by dinoflagellates, in summer [56]. We thus infer that the spatially-divergent Fv/Fm values, coupled with persistently low σPSII, primarily reflect photophysiological nutrient effects in relatively large cells.

Light-dependent effects and residual NPQ

We observed strong residual NPQ effects after five minutes of low light acclimation (Figs 3 and 4). Notably, the extent of these quenching effects was a predictable function of the short-term light history experienced by in situ phytoplankton assemblages (Fig 3d). Previous studies examining the drivers of NPQ variability [6264] suggest that the magnitude of NPQ effects at a given light level is tied to a number of environmental factors (e.g. temperature and CO2 concentrations), phytoplankton taxonomy and physiological status. Given these sources of variability, NPQ relaxation times needed for robust Fv/Fm measurements are expected to differ significantly across ocean regimes. In cold waters, such as those encountered along our ship track, NPQ relaxation is slower [65], and this may have contributed to the longer-lived quenching observed in our low-light samples. As the spatial and temporal resolutions of Fv/Fm and ETRa measurements are constrained by such acclimation periods, it is recommended that future FRRf field deployments conduct experiments using natural assemblages to determine the regional minimum relaxation period necessary to achieve steady-state dark-acclimation. This acclimation step can then be incorporated into underway FRRf protocols, resulting in more robust ETRa estimates, albeit with reduced measurement frequency. Such routine determinations of the minimum NPQ relaxation time requirements have not been commonly carried out for marine phytoplankton [25]. As a result, there is little systematic knowledge of the global variability of NPQ relaxation times. Adopting such pre-study tests (or within protocol) as standard practice would improve current understanding of environmental and taxonomic controls on NPQ relaxation kinetics. Moreover, it may be necessary to adjust the length of the dark-acclimation period to reflect changing conditions over the duration of a cruise. We thus recommend that future work incorporate semi-regular assessments of dark-acclimation times into field-sampling protocols.

As an initial step towards resolving the potential influence of dark acclimation times on retrieved photo-physiological parameters, we conducted preliminary experiments on natural Arctic phytoplankton assemblages to determine optimal dark acclimation times, whereby σPSII and Fv/Fm recovery rates were determined for surface water samples retained in darkness over 30 minutes. These preliminary experiments were carried out on the most recent CCGS Amundsen expedition (Sept. 03–09, 2021) as the ship transited through Lancaster Sound.

Our analysis revealed that phytoplankton reached maximum, stable, levels of light utilization efficiency (as judged by σPSII and Fv/Fm) within 9.65 ± 1.53 min (n = 34). In our present study, we used a constant 5 minute low light acclimation period. Increasing this to 10 minutes would have increased our total measurement interval from 12 to 17 minutes, resulting in a ∼ 30% reduction in sample size. Ultimately, the appropriate low light/dark acclimation time depends on the research question of interest as sampling protocols with shorter acclimation periods provide higher resolution datasets more reflective of in-situ photophysiological conditions, whereas longer acclimation periods reduce NPQ effects on σPSII and Fv/Fm values, and thus simplify data interpretation.

Decoupling of ETRa and ETRk

Across our study region, ETRk significantly exceeded ETRa (Fig 6), with the magnitude of ETRk and ETRa decoupling varying strongly in response to phytoplankton photophysiological conditions (Fig 7). The ratio between ETRk and ETRa depends on a number of variables:

ETRk:ETRa=Fv/FmσPSII×τ×Fq/FmEmax×Emax×6.022×103, (6)

Importantly, all of the terms defining ETRk:ETRa are potentially responsive to shifts in nutrient abundances and phytoplankton taxonomic composition. Nutrient enrichment experiments have demonstrated increasing τQa with nutrient deficiency and elevated actinic irradiance [30, 55]. Moreover, σPSII can also change with nutrient availability, but the observed percent change in σPSII following short-term nitrate enrichment is small compared with that of Fv/Fm [5, 55]. Baseline fluorescence, may also influence the terms used to define ETRk:ETRa. This phenomenon represents a non-variable contribution to the Chl fluorescence signal, which is understood to reflect the presence of energetically decoupled light harvesting complexes under nutrient limitation or photoinhibitory stress [31]. High baseline fluorescence decreases the amplitude of fluorescence transients, making Fv/Fm a useful gauge of phytoplankton physiological stress [5, 66, 67]. Analysis of our data revealed a strong correlation between Fv/Fm and Fq/FmEmax, (ρ = 0.93, p < 0.001, n = 25), suggesting Fq/FmEmax is similarly affected by baseline fluorescence and reflective of physiological status. However, in one sample, we recorded the highest Fv/Fm but the lowest Fq/FmEmax. The sample also showed the highest ETRk:ETRa value (Fig 7). After removing this one sample, the correlation between Fv/Fm and ETRk:ETRa strengthened from ρ = -0.60 before removal to ρ = -0.86 after removal (p < 0.01 for both, n = 25 and 24, respectively). We conclude that differential environmental and taxonomic sensitivities of the variables used to derive ETRa and ETRk can lead to discrepancies between these two productivity photosynthesis metrics.

Further investigation of ETRk and ETRa divergence is critical to inform our understanding of electron requirements for carbon assimilation and biomass production, particularly under nutrient-limiting conditions [28]. For instance, Schuback et al. [36] found that iron-limited phytoplankton assemblages exhibited elevated ETRa and greater decoupling between ETRa and C-assimilation rates as compared to iron-enriched assemblages. This result was attributed to the higher σPSII values in iron limited samples. Since ETRk does not directly include σPSII, we speculate that C-assimilation will show less decoupling with this photosynthesis metric under low iron conditions. This hypothesis remains to be tested in future studies.

Computational considerations

Beyond the physiological and taxonomic effects described above, the computational procedures used to analyze Chla fluorescence relaxation kinetics and derive turnover rates of electron transport molecules may also have a direct effect on the observed relationship between ETRk and ETRa. We derived τQa parameters used to calculate ETRk using the FRRf fluorescence transient fitting approach, as outlined in ‘Electron Transport Rates, ETRPSII’ Materials and Methods section. This approach relies on numerically fitting the rate of change in the redox state of Qa, driven by electron fluxes in and out of PSII reaction centers.

CQat=eineout, (7)

Here ein is equivalent to the rate of primary photochemistry induced by excitation flashlets, and eout is controlled by Qa reoxidation. Within FRRf Soliense software, ein is formulated as,

ein=E(t,λ)×σPSII×(1C(t)1C(t)p), (8)

By comparison, FIRe-based analysis of fluorescence transients deviates from FRRf by including an additional term to describe reaction center closure by background actinic light (PAR) [30]. As a result, Eq 9 is modified as:

ein=E(t,λ)×σPSII×(1C(t)1C(t)p)+PAR(λ)×σPSII×(1C(t)1C(t)p), (9)

The FRRf based analysis does not include a PAR term, as it is presumed that constant background light influences the baseline of the fluorescence signal, but does not contribute to dynamic changes in fluorescence measured over the course of an ST flash [Z. Kolber, pers. comm.]. In this interpretation, C(t) represents the fraction of initially available reaction centers closed by excitation pulses, such that C(t = 0) always equals 0.

Gorbunov and Falkowski [30] conducted a primary analysis of differences in τQA values retrieved from FRRf and FIRe fluorescence relaxation analyses. Their results showed that when the effect of background light was explicitly included in numerical formulations, FIRe-derived τQA values displayed strong actinic light dependencies, increasing with irradiance until plateauing around saturating irradiances (Ek). By contrast, their FRRf-derived τQA values varied little with actinic irradiance and resulted in a markedly shorter photosynthetic turnover time. Our own FRRf-derived τQA values displayed a weak relationship with applied actinic irradiances (ρ = 0.17, p < 0.01, n = 203). It is possible that applying the FIRe model fit to our own data may have also yielded slower photosynthetic turnover times, and therefore lower ETRk estimates, but it is unclear to what extent the alternative model may have affected our ETRk results and the observed decoupling between ETRk and ETRa.

Going forward, it will be important to separate physiological drivers of ETRk and ETRa decoupling from offsets resulting from the use of different mathematical numerical approaches to data analysis. Discrepancies between ETRk and ETRa that cannot be explained by different derivations of τQa must be attributable to differences between the two ETRPSII algorithms themselves. This raises the important question of which approach is most accurate, as neither has been established as a “gold standard”. Addressing this issue will require parallel independent measurements of PSII activity, such as gross oxygen evolution measurements from 18O experiments, as preformed previously for ETRa (e.g. [54]) but not ETRk. Such fluorescence-independent validations of ETRa and ETRk are critically lacking, and will help elucidate the taxonomic and environmental influences on FRRf-based productivity photochemistry measurements [28]. This, in turn, will be of significant practical utility to FRRf users seeking to derive ship-based primary productivity estimates.

Conclusions and future recommendations

Fast Repetition Rate fluorometry offers a means to rapidly assess both physiological status and photosynthetic electron transport rates of phytoplankton. The aim of this study was to evaluate an autonomous protocol for high resolution FRRf measurements of phytoplankton physiology, and to compare two alternative models for deriving primary productivity photochemistry estimates from FRRf data. Our results demonstrate significant residual NPQ effects after five minutes of low light acclimation, suggesting the need for extended low light acclimation periods, which would significantly decrease measurement frequency. Our findings also illustrated localized regions of elevated Fv/Fm, likely linked to local nitrogen loading by freshwater inputs and tidal mixing. Although the amplitude-based and kinetic-based derivations of photosynthetic electron transport rates were well correlated, absolute agreement between estimates appeared to be affected by phytoplankton photophysiology, with the two models diverging under nutrient-limited conditions.

As a first step, resolving discrepancies between ETRa and ETRk will require consensus regarding the analysis of raw fluorescence transient data. This aim is fundamental for consistent data reporting among the growing community of FRRf and FIRe users [31]. Second, ETRa and ETRk should be validated against fluorescence-independent measures of productivity. Considering that measurement of ETRk does not require a dark acclimation step, short (∼ 5 min.) acclimation steps should be sufficient to accurately derive this term. Based on these findings, the ETRk model, if validated against independent productivity metrics, may be advantageous for high resolution evaluations of in-situ photosynthetic rates under ambient light conditions.

Supporting information

S1 Table. Correlation of underway hydrographic variables.

Results of Spearman Rank correlation analyses between each underway hydrographic variable are displayed. ** is used to indicate p values < 0.001. In all instances n = 7200.

(PDF)

S2 Table. CTD profiling stations.

Sampling station locations are indicated as LS for Lancaster Sound or BS for Barrow Strait. Station mixed layer depth (MLD), mean mixed layer nitrate and nitrite concentration, and Chl a concentration within the mixed layer are shown.

(PDF)

S3 Table. Surface PAR and underway photophysiological variable regression analyses.

All p values were < 0.001. Standard error is reported for the regression intercept and coefficient, respectively.

(PDF)

Acknowledgments

CTD station, surface PAR, and TSG data presented herein were collected by the Canadian research icebreaker CCGS Amundsen and made available by the Amundsen Science program, which is supported through Université Laval by the Canada Foundation for Innovation. We wish to also thank Maxim Gorbunov and Nina Schuback for valuable comments on this manuscript.

Data Availability

Salinity and temperature data collected by Amundsen Science are available in the Polar Data Catalogue (doi: 10.5884/12715). Surface photosynthetically active radiation data are available in the Polar Data Catalogue (doi: 10.5884/12518), alongside additional meteorological data provided by the Amundsen Science group of U. Laval. CTD station conductivity, temperature, and salinity data provided by the Amundsen Science group of U. Laval are available in the Polar Data Catalogue (doi: 10.5884/12713). Underway oxygen saturation and ΔO2/Ar data are available in the Polar Data Catalogue (doi: 10.5884/13242). All Fast Repetition Rate Fluorometry measurements of phytoplankton photophysiology are available in the Polar Data Catalogue (doi: 10.5884/13254). The Canadian Cryospheric Information Network reference number for this dataset is 13254, and will be trackable using this number once published. A doi will be provided as soon as available. River data collected by the Canadian GEOTRACES program is available from the PANGEA database (https://doi.org/10.1594/PANGAEA.908497).

Funding Statement

This work was supported by ArcticNet, a Network of Centres of Excellence of Canada and by the Natural Sciences and Engineering Research Council of Canada (NSERC) grants awarded to PT (https://www.nserc-crsng.gc.ca). The funders had no role in study design, data collection, decision to publish, or preparation of the manuscript.

References

  • 1. Tremblay JÉ, Anderson LG, Matrai P, Coupel P, Bélanger S, Michel C, et al. Global and regional drivers of nutrient supply, primary production and CO2 drawdown in the changing Arctic Ocean. Prog Oceanogr. 2015;139:171–96. doi: 10.1016/j.pocean.2015.08.009 [DOI] [Google Scholar]
  • 2. Tremblay JÉ, Gagnon J. The effects of irradiance and nutrient supply on the productivity of Arctic waters: a perspective on climate change. In: Nihoul JCJ, Kostianoy AG, editors. Influence of Climate Change on the Changing Arctic and Sub-Arctic Conditions. Dordrecht: Springer Netherlands; 2009. p. 73–93. [Google Scholar]
  • 3. Tremblay JÉ, Michel C, Hobson KA, Gosselin M, Price NM. Bloom dynamics in early opening waters of the Arctic Ocean. Limnol Oceanogr. 2006;51(2):900–12. doi: 10.4319/lo.2006.51.2.0900 [DOI] [Google Scholar]
  • 4. Lewis KM, Arntsen AE, Coupel P, Joy-Warren H, Lowry KE, Matsuoka A, et al. Photoacclimation of Arctic Ocean phytoplankton to shifting light and nutrient limitation. Limnol Oceanogr. 2019;64(1):284–301. doi: 10.1002/lno.11039 [DOI] [Google Scholar]
  • 5. Ko E, Gorbunov MY, Jung J, Joo HM, Lee Y, Cho KH, et al. Effects of Nitrogen Limitation on Phytoplankton Physiology in the Western Arctic Ocean in Summer. J Geophys Res Ocean. 2020;125(11):1–19. doi: 10.1029/2020JC016501 [DOI] [Google Scholar]
  • 6. Randelhoff A, Holding J, Janout M, Sejr MK, Babin M, Tremblay JÉ, et al. Pan-Arctic Ocean Primary Production Constrained by Turbulent Nitrate Fluxes. Front Mar Sci. 2020;7(March):1–15.32802822 [Google Scholar]
  • 7. Terhaar J, Lauerwald R, Regnier P, Gruber N, Bopp L. Around one third of current Arctic Ocean primary production sustained by rivers and coastal erosion. Nat Commun. 2021;12(1):1–10. doi: 10.1038/s41467-020-20470-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Ardyna M, Gosselin M, Michel C, Poulin M, Tremblay JÉ. Environmental forcing of phytoplankton community structure and function in the Canadian High arctic: Contrasting oligotrophic and eutrophic regions. Mar Ecol Prog Ser. 2011;442:37–57. doi: 10.3354/meps09378 [DOI] [Google Scholar]
  • 9. Hannah CG, Dupont F, Dunphy MD. Polynyas and tidal currents in the Canadian Arctic Archipelago. Arctic. 2009;62(1):83–95. doi: 10.14430/arctic115 [DOI] [Google Scholar]
  • 10. Hughes KG, Klymak JM, Williams WJ, Melling H. Tidally Modulated Internal Hydraulic Flow and Energetics in the Central Canadian Arctic Archipelago. J Geophys Res Ocean. 2018;123(8):5210–29. doi: 10.1029/2018JC013770 [DOI] [Google Scholar]
  • 11.Brown KA, Williams WJ, Carmack EC, Fiske G, Francois R, McLennan D, et al. Canadian Arctic Archipelago Rivers Project Geochemical Data 2014-2016 [Internet]. PANGAEA; 2019. Available from: 10.1594/PANGAEA.908497 [DOI]
  • 12. Young KL, Lafrenière MJ, Lamoureux SF, Abnizova A, Miller EA. Recent multi-year streamflow regimes and water budgets of hillslope catchments in the canadian high arctic: Evaluation and comparison to other small Arctic watershed studies. Hydrol Res. 2015;46(4):533–50. doi: 10.2166/nh.2014.004 [DOI] [Google Scholar]
  • 13. Fritz M, Vonk JE, Lantuit H. Collapsing Arctic coastlines. Nat Clim Chang. 2017;7(1):6–7. doi: 10.1038/nclimate3188 [DOI] [Google Scholar]
  • 14. Peterson BJ, Holmes RM, McClelland JW, Vörösmarty CJ, Lammers RB, Shiklomanov AI, et al. Increasing river discharge to the Arctic Ocean. Science (80-). 2002;298(5601):2171–3. doi: 10.1126/science.1077445 [DOI] [PubMed] [Google Scholar]
  • 15. Nummelin A, Ilicak M, Li C, Smedsrud LH. Consequences of future increased Arctic runoff on Arctic Ocean stratification, circulation, and sea ice cover. J Geophys Res Ocean. 2016;121(1):617–37. doi: 10.1002/2015JC011156 [DOI] [Google Scholar]
  • 16. Coupel P, Ruiz-Pino D, Sicre MA, Chen JF, Lee SH, Schiffrine N, et al. The impact of freshening on phytoplankton production in the Pacific Arctic Ocean. Prog Oceanogr. 2015;131:113–25. doi: 10.1016/j.pocean.2014.12.003 [DOI] [Google Scholar]
  • 17. Kolber Z, Falkowski PG. Use of active fluorescence to estimate phytoplankton photosynthesis in situ. Limnol Oceanogr. 1993;38(8):1646–65. doi: 10.4319/lo.1993.38.8.1646 [DOI] [Google Scholar]
  • 18. Behrenfeld MJ, Kolber ZS. Widespread iron limitation of phytoplankton in the south pacific ocean. Science (80-). 1999;283(5403):840–3. doi: 10.1126/science.283.5403.840 [DOI] [PubMed] [Google Scholar]
  • 19. Moore CM, Suggett D, Holligan PM, Sharples J, Abraham ER, Lucas MI, et al. Physical controls on phytoplankton physiology and production at a shelf sea front: A fast repetition-rate fluorometer based field study. Mar Ecol Prog Ser. 2003;259:29–45. doi: 10.3354/meps259029 [DOI] [Google Scholar]
  • 20. Smyth TJ, Pemberton KL, Aiken J, Geider RJ. A methodology to determine primary production and phytoplankton photosynthetic parameters from Fast Repetition Rate Fluorometry. J Plankton Res [Internet]. 2004. Nov 1;26(11):1337–50. Available from: 10.1093/plankt/fbh124 [DOI] [Google Scholar]
  • 21. Fujiki T, Hosaka T, Kimoto H, Ishimaru T, Saino T. In situ observation of phytoplankton productivity by an underwater profiling buoy system: Use of fast repetition rate fluorometry. Mar Ecol Prog Ser. 2008;353:81–8. doi: 10.3354/meps07151 [DOI] [Google Scholar]
  • 22. Kromkamp JC, Forster RM. The use of variable fluorescence measurements in aquatic ecosystems: Differences between multiple and single turnover measuring protocols and suggested terminology. Eur J Phycol. 2003;38(2):103–12. doi: 10.1080/0967026031000094094 [DOI] [Google Scholar]
  • 23. Kolber ZS, Prášil O, Falkowski PG. Measurements of variable chlorophyll fluorescence using fast repetition rate techniques: Defining methodology and experimental protocols. Biochim Biophys Acta—Bioenerg. 1998;1367(1–3):88–106. doi: 10.1016/S0005-2728(98)00135-2 [DOI] [PubMed] [Google Scholar]
  • 24. Gorbunov MY, Kolber ZS, Falkowski PG. Measuring photosynthetic parameters in individual algal cells by Fast Repetition Rate fluorometry. Photosynth Res. 1999;62(2):141–53. doi: 10.1023/A:1006360005033 [DOI] [Google Scholar]
  • 25. Moore CM, Suggett DJ, Hickman AE, Kim YN, Tweddle JF, Sharples J, et al. Phytoplankton photoacclimation and photoadaptation in response to environmental gradients in a shelf sea. Limnol Oceanogr. 2006;51(2):936–49. doi: 10.4319/lo.2006.51.2.0936 [DOI] [Google Scholar]
  • 26. Oxborough K, Moore CM, Suggett DJ, Lawson T, Chan HG, Geider RJ. Direct estimation of functional PSII reaction center concentration and PSII electron flux on a volume basis: A new approach to the analysis of Fast Repetition Rate fluorometry (FRRf) data. Limnol Oceanogr Methods. 2012;10(MARCH):142–54. doi: 10.4319/lom.2012.10.142 [DOI] [Google Scholar]
  • 27. Suggett D, Právsil O, Borowitzka M. Chlorophyll a fluorescence in Aquatic Sciences: Methods and Applications. Dev Appl Phycol. 2011; [Google Scholar]
  • 28. Hughes DJ, Varkey D, Doblin MA, Ingleton T, Mcinnes A, Ralph PJ, et al. Impact of nitrogen availability upon the electron requirement for carbon fixation in Australian coastal phytoplankton communities. Limnol Oceanogr. 2018;63(5):1891–910. doi: 10.1002/lno.10814 [DOI] [Google Scholar]
  • 29. Boatman TG, Geider RJ, Oxborough K. Improving the accuracy of single turnover active fluorometry (STAF) for the estimation of phytoplankton primary productivity (PhytoPP). Front Mar Sci. 2019;6(JUN):1–16. [Google Scholar]
  • 30. Gorbunov MY, Falkowski PG. Using chlorophyll fluorescence kinetics to determine photosynthesis in aquatic ecosystems. Limnol Oceanogr. 2020;1–13. [Google Scholar]
  • 31. Schuback N, Tortell PD, Berman-Frank I, Campbell DA, Ciotti A, Courtecuisse E, et al. Single-Turnover Variable Chlorophyll Fluorescence as a Tool for Assessing Phytoplankton Photosynthesis and Primary Productivity: Opportunities, Caveats and Recommendations. Front Mar Sci [Internet]. 2021;8:895. Available from: https://www.frontiersin.org/article/10.3389/fmars.2021.690607 [Google Scholar]
  • 32. Kolber Z, Falkowski PG. Variations in Chlorophyll fluorescence yields in phytoplankton in the World Oceans. Aust J Plant Physiol. 1995;22:341–55. [Google Scholar]
  • 33. Bates SS. Sample preconditioning for measurement of fluorescence induction of chlorophyll a in marine phytoplankton. J Plankton Res. 1985;7(5):703–14. doi: 10.1093/plankt/7.5.703 [DOI] [Google Scholar]
  • 34. Walter B, Peters J, van Beusekom JEE. The effect of constant darkness and short light periods on the survival and physiological fitness of two phytoplankton species and their growth potential after re-illumination. Aquat Ecol. 2017;51(4):591–603. doi: 10.1007/s10452-017-9638-z [DOI] [Google Scholar]
  • 35. Alderkamp AC, De Baar HJW, Visser RJW, Arrigo KR. Can photoinhibition control phytoplankton abundance in deeply mixed water columns of the Southern Ocean? Limnol Oceanogr. 2010;55(3):1248–64. doi: 10.4319/lo.2010.55.3.1248 [DOI] [Google Scholar]
  • 36. Schuback N, Schallenberg C, Duckham C, Maldonado MT, Tortell PD. Interacting Effects of Light and Iron Availability on the Coupling of Photosynthetic Electron Transport and CO 2 -Assimilation in Marine Phytoplankton. PLoS One [Internet]. 2015;1–30. Available from: 10.1371/journal.pone.0133235 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Wei Y, Zhao X, Sun J, Liu H. Fast Repetition Rate Fluorometry (FRRF) Derived Phytoplankton Primary Productivity in the Bay of Bengal. Front Mar Sci. 2019;10(May):1–13. doi: 10.3389/fmicb.2019.01164 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Zhu Y, Ishizaka J, Tripathy SC, Wang S, Sukigara C, Goes J, et al. Relationship between light, community composition and the electron requirement for carbon fixation in natural phytoplankton. Mar Ecol Prog Ser. 2017;580:83–100. doi: 10.3354/meps12310 [DOI] [Google Scholar]
  • 39. Tortell PD. Dissolved gas measurements in oceanic waters made by membrane inlet mass spectrometry. Limnol Oceanogr Methods. 2005;3(1):24–37. doi: 10.4319/lom.2005.3.24 [DOI] [Google Scholar]
  • 40. Tortell PD, Guéguen C, Long MC, Payne CD, Lee P, DiTullio GR. Spatial variability and temporal dynamics of surface water pCO2, δO2/Ar and dimethylsulfide in the Ross Sea, Antarctica. Deep Res Part I Oceanogr Res Pap. 2011;58(3):241–59. doi: 10.1016/j.dsr.2010.12.006 [DOI] [Google Scholar]
  • 41.Pawlowicz, R., 2020. M_Map: A mapping package for MATLAB version 1.4m, Matlab R2020a, available online at www.eoas.ubc.ca/rich/map.html.
  • 42.Amundsen Science Data Collection. 2019. CTD-Rosette data collected by the CCGS Amundsen in the Canadian Arctic. Arcticnet Inc., Québec, Canada. Processed data. Version 1. Archived at www.polardata.ca, Canadian Cryospheric Information Network (CCIN), Waterloo, Canada. Accessed 2021.
  • 43.Amundsen Science Data Collection. 2019. TSG data collected by the CCGS Amundsen in the Canadian Arctic. Arcticnet Inc., Québec, Canada. Processed data. Version 1. Archived at www.polardata.ca, Canadian Cryospheric Information Network (CCIN), Waterloo, Canada. Accessed on 2021.
  • 44.Amundsen Science Data Collection. 2019. AVOS Meteorological Data collected by the CCGS Amundsen in the Canadian Arctic. Arcticnet Inc., Québec, Canada, Processed data. Version 1. Archived at www.polardata.ca, Canadian Cryospheric Information Network (CCIN), Waterloo, Canada. Accessed on 2021.
  • 45. Suggett DJ, Goyen S, Evenhuis C, Szabó M, Pettay DT, Warner ME, et al. Functional diversity of photobiological traits within the genus Symbiodinium appears to be governed by the interaction of cell size with cladal designation. New Phytol. 2015;208(2):370–81. doi: 10.1111/nph.13483 [DOI] [PubMed] [Google Scholar]
  • 46. Schuback N, Hoppe CJM, Tremblay JÉ, Maldonado MT, Tortell PD. Primary productivity and the coupling of photosynthetic electron transport and carbon fixation in the Arctic Ocean. Limnol Oceanogr. 2017;62(3):898–921. doi: 10.1002/lno.10475 [DOI] [Google Scholar]
  • 47. Webb WL, Newton M, Starr D. Carbon dioxide exchange of Alnus rubra—A mathematical model. Oecologia. 1974;17(4):281–91. doi: 10.1007/BF00345747 [DOI] [PubMed] [Google Scholar]
  • 48. Bilger W, Björkman O. Role of the xanthophyll cycle in photoprotection elucidated by measurements of light-induced absorbance changes, fluorescence and photosynthesis in leaves of Hedera canariensis. Photosynth Res. 1990;25(3):173–85. doi: 10.1007/BF00033159 [DOI] [PubMed] [Google Scholar]
  • 49. Schuback N, Tortell PD. Diurnal regulation of photosynthetic light absorption, electron transport and carbon fixation in two contrasting oceanic environments. Biogeosciences Discuss. 2019;1–33. [Google Scholar]
  • 50. Xie Y, Laws EA, Yang L, Huang B. Diel patterns of variable fluorescence and carbon fixation of picocyanobacteria prochlorococcus-dominated phytoplankton in the South China Sea basin. Front Microbiol. 2018;9(AUG). doi: 10.3389/fmicb.2018.01589 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Falkowski PG, Lin H, Gorbunov MY. What limits photosynthetic energy conversion efficiency in nature? Lessons from the oceans. Philos Trans R Soc B Biol Sci. 2017;372 (1730). doi: 10.1098/rstb.2016.0376 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52. Martin J, JÉ T, Gagnon J, Tremblay G, Lapoussière A, Jose C, et al. Prevalence, structure and properties of subsurface chlorophyll maxima in Canadian Arctic waters. Mar Ecol Prog Ser [Internet]. 2010;412:69–84. Available from: https://www.int-res.com/abstracts/meps/v412/p69-84/ [Google Scholar]
  • 53. Schreiber U, Klughammer C, Kolbowski J. Assessment of wavelength-dependent parameters of photosynthetic electron transport with a new type of multi-color PAM chlorophyll fluorometer. Photosynth Res. 2012;113(1–3):127–44. doi: 10.1007/s11120-012-9758-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Suggett DJ, Moore CM, Hickman AE, Geider RJ. Interpretation of fast repetition rate (FRR) fluorescence: Signatures of phytoplankton community structure versus physiological state. Mar Ecol Prog Ser. 2009;376:1–19. doi: 10.3354/meps07830 [DOI] [Google Scholar]
  • 55. Mills MM, Brown ZW, Laney SR, Ortega-Retuerta E, Lowry KE, van Dijken GL, et al. Nitrogen limitation of the summer phytoplankton and heterotrophic prokaryote communities in the Chukchi Sea. Front Mar Sci. 2018;5(OCT):1–22.29552559 [Google Scholar]
  • 56. Viljoen JJ, Philibert R, Van Horsten N, Mtshali T, Roychoudhury AN, Thomalla S, et al. Phytoplankton response in growth, photophysiology and community structure to iron and light in the Polar Frontal Zone and Antarctic waters. Deep Res Part I [Internet]. 2018;141(August):118–29. Available from: 10.1016/j.dsr.2018.09.006 [DOI] [Google Scholar]
  • 57. Roncel M, González-Rodríguez AA, Naranjo B, Bernal-Bayard P, Lindahl M, Hervás M, et al. Iron deficiency induces a partial inhibition of the photosynthetic electron transport and a high sensitivity to light in the diatom Phaeodactylum tricornutum. Front Plant Sci. 2016;7(AUG2016):1–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58. Colombo M, Jackson SL, Cullen JT, Orians KJ. Dissolved iron and manganese in the Canadian Arctic Ocean: On the biogeochemical processes controlling their distributions. Geochim Cosmochim Acta. 2020;277:150–74. doi: 10.1016/j.gca.2020.03.012 [DOI] [Google Scholar]
  • 59. Brown KA, Williams WJ, Carmack EC, Fiske G, François R, McLennan D, et al. Geochemistry of Small Canadian Arctic Rivers with Diverse Geological and Hydrological Settings. J Geophys Res Biogeosciences [Internet]. 2020;125(1):e2019JG005414. Available from: https://agupubs.onlinelibrary.wiley.com/doi/abs/10.1029/2019JG005414 [Google Scholar]
  • 60. Murata N, Takahashi S, Nishiyama Y, Allakhverdiev SI. Photoinhibition of photosystem II under environmental stress. Biochim Biophys Acta. 2007. Jun;1767(6):414–21. doi: 10.1016/j.bbabio.2006.11.019 [DOI] [PubMed] [Google Scholar]
  • 61. Li Z, Lan T, Zhang J, Gao K, Beardall J, Wu Y. Nitrogen Limitation Decreases the Repair Capacity and Enhances Photoinhibition of Photosystem II in a Diatom. Photochem Photobiol. 2021;(March). [DOI] [PubMed] [Google Scholar]
  • 62. Algarra P, Rudiger W. Acclimation processes in the light harvesting complex of the red alga Porphyridium purpureum (Bory) Drew et Ross, according to irradiance and nutrient availability. Plant Cell Environ. 1993;16(2):149–59. doi: 10.1111/j.1365-3040.1993.tb00856.x [DOI] [Google Scholar]
  • 63. Salvucci ME, Crafts-Brandner SJ. Relationship between the heat tolerance of photosynthesis and the thermal stability of rubisco activase in plants from contrasting thermal environments. Plant Physiol. 2004;134(4):1460–70. doi: 10.1104/pp.103.038323 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64. Trimborn S, Thoms S, Petrou K, Kranz SA, Rost B. Photophysiological responses of Southern Ocean phytoplankton to changes in CO2 concentrations: Short-term versus acclimation effects. J Exp Mar Bio Ecol [Internet]. 2014;451(2014):44–54. Available from: 10.1016/j.jembe.2013.11.001 [DOI] [Google Scholar]
  • 65. Sherman J, Gorbunov MY, Schofield O, Falkowski PG. Photosynthetic energy conversion efficiency in the West Antarctic Peninsula. Limnol Oceanogr. 2020;65(12):2912–25. doi: 10.1002/lno.11562 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66. Parkhill JP, Maillet G, Cullen JJ. Fluorescence-based maximal quantum yield for PSII as a diagnostic of nutrient stress. J Phycol. 2001;37(4):517–29. doi: 10.1046/j.1529-8817.2001.037004517.x [DOI] [Google Scholar]
  • 67. Saeck EA, O’Brien KR, Burford MA. Nitrogen response of natural phytoplankton communities: a new indicator based on photosynthetic efficiency Fv/Fm. Mar Ecol Prog Ser. 2016; doi: 10.3354/meps11729 [DOI] [Google Scholar]

Decision Letter 0

Matheus C Carvalho

14 Sep 2021

PONE-D-21-24983Chlorophyll fluorescence-based estimates of photosynthetic electron transport in Arctic phytoplankton assemblagesPLOS ONE

Dear Dr. Sezginer,

Thank you for submitting your manuscript to PLOS ONE. After careful consideration, we feel that it has merit but does not fully meet PLOS ONE’s publication criteria as it currently stands. Therefore, we invite you to submit a revised version of the manuscript that addresses the points raised during the review process.

Please submit your revised manuscript by Oct 29 2021 11:59PM. If you will need more time than this to complete your revisions, please reply to this message or contact the journal office at plosone@plos.org. When you're ready to submit your revision, log on to https://www.editorialmanager.com/pone/ and select the 'Submissions Needing Revision' folder to locate your manuscript file.

Please include the following items when submitting your revised manuscript:

  • A rebuttal letter that responds to each point raised by the academic editor and reviewer(s). You should upload this letter as a separate file labeled 'Response to Reviewers'.

  • A marked-up copy of your manuscript that highlights changes made to the original version. You should upload this as a separate file labeled 'Revised Manuscript with Track Changes'.

  • An unmarked version of your revised paper without tracked changes. You should upload this as a separate file labeled 'Manuscript'.

If you would like to make changes to your financial disclosure, please include your updated statement in your cover letter. Guidelines for resubmitting your figure files are available below the reviewer comments at the end of this letter.

If applicable, we recommend that you deposit your laboratory protocols in protocols.io to enhance the reproducibility of your results. Protocols.io assigns your protocol its own identifier (DOI) so that it can be cited independently in the future. For instructions see: https://journals.plos.org/plosone/s/submission-guidelines#loc-laboratory-protocols. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols.

We look forward to receiving your revised manuscript.

Kind regards,

Matheus C. Carvalho

Academic Editor

PLOS ONE

Journal Requirements:

When submitting your revision, we need you to address these additional requirements.

1. Please ensure that your manuscript meets PLOS ONE's style requirements, including those for file naming. The PLOS ONE style templates can be found at 

https://journals.plos.org/plosone/s/file?id=wjVg/PLOSOne_formatting_sample_main_body.pdf and 

https://journals.plos.org/plosone/s/file?id=ba62/PLOSOne_formatting_sample_title_authors_affiliations.pdf

2. In your Methods section, please provide additional information regarding the permits you obtained for the work. Please ensure you have included the full name of the authority that approved the field site access and, if no permits were required, a brief statement explaining why.

3. In your Methods section, please provide additional location information, including geographic coordinates for the data set if available. 

4. We note that you have stated that you will provide repository information for your data at acceptance. Should your manuscript be accepted for publication, we will hold it until you provide the relevant accession numbers or DOIs necessary to access your data. If you wish to make changes to your Data Availability statement, please describe these changes in your cover letter and we will update your Data Availability statement to reflect the information you provide.

5. Please update your submission to use the PLOS LaTeX template. The template and more information on our requirements for LaTeX submissions can be found at http://journals.plos.org/plosone/s/latex.

6. We note that Figures 1 and 9 in your submission contain [map/satellite] images which may be copyrighted. All PLOS content is published under the Creative Commons Attribution License (CC BY 4.0), which means that the manuscript, images, and Supporting Information files will be freely available online, and any third party is permitted to access, download, copy, distribute, and use these materials in any way, even commercially, with proper attribution. For these reasons, we cannot publish previously copyrighted maps or satellite images created using proprietary data, such as Google software (Google Maps, Street View, and Earth). For more information, see our copyright guidelines: http://journals.plos.org/plosone/s/licenses-and-copyright.

We require you to either (1) present written permission from the copyright holder to publish these figures specifically under the CC BY 4.0 license, or (2) remove the figures from your submission:

a) You may seek permission from the original copyright holder of Figures 1 and 9 to publish the content specifically under the CC BY 4.0 license.  

We recommend that you contact the original copyright holder with the Content Permission Form (http://journals.plos.org/plosone/s/file?id=7c09/content-permission-form.pdf) and the following text:

“I request permission for the open-access journal PLOS ONE to publish XXX under the Creative Commons Attribution License (CCAL) CC BY 4.0 (http://creativecommons.org/licenses/by/4.0/). Please be aware that this license allows unrestricted use and distribution, even commercially, by third parties. Please reply and provide explicit written permission to publish XXX under a CC BY license and complete the attached form.”

Please upload the completed Content Permission Form or other proof of granted permissions as an "Other" file with your submission.

In the figure caption of the copyrighted figure, please include the following text: “Reprinted from [ref] under a CC BY license, with permission from [name of publisher], original copyright [original copyright year].”

b) If you are unable to obtain permission from the original copyright holder to publish these figures under the CC BY 4.0 license or if the copyright holder’s requirements are incompatible with the CC BY 4.0 license, please either i) remove the figure or ii) supply a replacement figure that complies with the CC BY 4.0 license. Please check copyright information on all replacement figures and update the figure caption with source information. If applicable, please specify in the figure caption text when a figure is similar but not identical to the original image and is therefore for illustrative purposes only.

The following resources for replacing copyrighted map figures may be helpful:

USGS National Map Viewer (public domain): http://viewer.nationalmap.gov/viewer/

The Gateway to Astronaut Photography of Earth (public domain): http://eol.jsc.nasa.gov/sseop/clickmap/

Maps at the CIA (public domain): https://www.cia.gov/library/publications/the-world-factbook/index.html and https://www.cia.gov/library/publications/cia-maps-publications/index.html

NASA Earth Observatory (public domain): http://earthobservatory.nasa.gov/

Landsat: http://landsat.visibleearth.nasa.gov/

USGS EROS (Earth Resources Observatory and Science (EROS) Center) (public domain): http://eros.usgs.gov/#

Natural Earth (public domain): http://www.naturalearthdata.com/

[Note: HTML markup is below. Please do not edit.]

Reviewers' comments:

Reviewer's Responses to Questions

Comments to the Author

1. Is the manuscript technically sound, and do the data support the conclusions?

The manuscript must describe a technically sound piece of scientific research with data that supports the conclusions. Experiments must have been conducted rigorously, with appropriate controls, replication, and sample sizes. The conclusions must be drawn appropriately based on the data presented.

Reviewer #1: Partly

Reviewer #2: Yes

Reviewer #3: Yes

**********

2. Has the statistical analysis been performed appropriately and rigorously?

Reviewer #1: Yes

Reviewer #2: Yes

Reviewer #3: Yes

**********

3. Have the authors made all data underlying the findings in their manuscript fully available?

The PLOS Data policy requires authors to make all data underlying the findings described in their manuscript fully available without restriction, with rare exception (please refer to the Data Availability Statement in the manuscript PDF file). The data should be provided as part of the manuscript or its supporting information, or deposited to a public repository. For example, in addition to summary statistics, the data points behind means, medians and variance measures should be available. If there are restrictions on publicly sharing data—e.g. participant privacy or use of data from a third party—those must be specified.

Reviewer #1: Yes

Reviewer #2: Yes

Reviewer #3: Yes

**********

4. Is the manuscript presented in an intelligible fashion and written in standard English?

PLOS ONE does not copyedit accepted manuscripts, so the language in submitted articles must be clear, correct, and unambiguous. Any typographical or grammatical errors should be corrected at revision, so please note any specific errors here.

Reviewer #1: Yes

Reviewer #2: Yes

Reviewer #3: Yes

**********

5. Review Comments to the Author

Please use the space provided to explain your answers to the questions above. You may also include additional comments for the author, including concerns about dual publication, research ethics, or publication ethics. (Please upload your review as an attachment if it exceeds 20,000 characters)

Reviewer #1: Reviewer comments on manuscript entitled “Chlorophyll fluorescence-based estimates of photosynthetic electron transport in Arctic phytoplankton assemblages” authored by Yayla Sezginer, David J. Suggett, Robert W. Izett1c, and Philippe D. Tortell1 for publication in PLOS ONE.

FRRf measurements was carried out in Arctic sea, to clarify how the productivity of phytoplankton community respond to glacial and land-derived nutrients. Authors suggested a new ETRk model based on the relaxation kinetics of PSII fluorescence based on a recent study (Gorbunov & Falkowski, 2021), and compared to ETRa, which is derived by a commonly-used biophysical model. Results show that ETRa and ETRk were similar in a nutrient-rich region but clearly differed in an oligotrophic region. The sampling and FRRf measurement were carried out at a high frequency, and analysis was performed appropriately. The results are valuable data for the Arctic Ocean. I have one caveat before publication, however, and it needs careful consideration.

The major point is that there is no support of apparent (measured) O2 evolution and C assimilation rate for the two ETRs in the study area. The amplitude based ETRa have been developed as biophysical models to derive gross primary production (GPP) because FRRf can reduces only QA but not PQ pool and thus examine the electron flow rate for oxygen evolution. However, ETRa-based estimated GPP does not always equal to apparent GPP (Regaudie-de-Gioux et al., 2014). Many previous studies have investigated the factor affecting the relationships between modeled ETRa and measured O2 evolution rate (Suggett et al., 2001; Robinson et al., 2009; Deblois, Marchand & Juneau, 2013) and C assimilation rate in natural communities (Lawrenz et al., 2013; Schuback et al., 2017; Zhu et al., 2017; Hughes et al., 2018b; Ryan-Keogh et al., 2018; Kazama et al., 2021). For oxygen-based GPP, it is well confirmed for cultivated species (Suggett et al., 2009) but species composition can affect conversion factor in natural communities (Deblois et al., 2013). For, C-based GPP, although there are many studies but no global model is derived to convert ETRa to GPP (Hughes et al., 2018a). On the other hand, authors’ models is not supported by oxygen production rate or C assimilation of algae yet. The kinetic based ETRk by FIRe system (FIRe-ETRk) is a novel method to improve the errors from the parameters (Gorbunov & Falkowski, 2021). The relationships between the FIRe-ETRk and the growth rate (net primary production, NPP) of two model species, Thalassiosira pseudonana, and Dunaliella tertiolecta were examined, but not in natural community yet. Also the relationships between FIRe-ETRk and GPP is not examined yet. Therefore, the use of authors’ kinetic model in natural algal communities without reference measured productivity (O2 or C) must be considered carefully, even if it is likely analogous to the FIRe-ETRk.

Because authors’ models is not supported by oxygen production rate of algae yet, the absolute value of ETRk cannot compare to ETRa. For example, when these two ETR values are used as estimators of GPP and NPP, respectively, in the study area, the results are still questionable. Because net productivity does not include respiration, ETRk must be lower than ETRa on the same phytoplankton assemblage. However, present study clearly showed ETRk > ETRa in Lancaster Sound (Fig. 8). This paradoxical results may be due to underestimation of ETRa, and/or overestimation of ETRk. The former is likely due to inadequate dark acclimation time (5 min, P11 L211). For rapid light curve, there is no consensus rule but typically 10~20 min of dark adaptation period is used (Schuback et al., 2021).The latter is likely due to the underestimation of turnover rate, as authors pointed out. If authors primarily focus to the comparison of relative oxygen productivity between Lancaster Sound and Barrow Strait, use Fv/Fm and ETRa but not ETRk.

Minor points

P2 L25 fast repetition rate fluorometry

P5 L93 photosynthetically active radiation

P8 L157 Provide company name and city for every instruments of the manuscript.

P8 L178 Unify the unit of whole manuscript (s, min).

P10 L207 Provide the version and company of the software.

P10 L209 Table 1 What is the ChlF?

Ų PSII−1

Definition of Fq’/Fm’ (max) should not be “under 150 μmol quanta m-2 s-1”.

P11 L217 Provide the duration of each acquisitions and total time per sample.

P11 L218 Use mol instead of E.

P11 L226 Provide the duration of each light step. For rapid light curve, less than 30 s is recommended (Perkins et al., 2010).

P13 L254-258 Ų PSII−1

P13 L262 Provide the reference for 3 component multi-exponential model.

Equation (2) What are the F(t) and CQA(t)?

P13 L273-276 QA, QB

P16 L319 Provide version, company and city of the software.

P17 L349 Table 2 Correct notation Chl a.

P17 L354 Table 3 Does the zero mean true zero or lower than detection limit?

(mg m−3)

P19 L387 Spearman’s rank correlation. Use ρ (rho) or “rs” instead of R. See Schober et al. 2018. Correlation Coefficients: Appropriate Use and Interpretation, Anesthesia & Analgesia: Volume 126, 5, 1763-1768. doi: 10.1213/ANE.0000000000002864

P20 L403 Use 25th and 75th percentile, or range with median value, instead of SD.

P20 L410-415 Integrate this part into discussion.

P20 L417 Explanation is needed why 150 μmol quanta m-2 s-1 is used.

P21 L423-426 Integrate this part into discussion.

P22 L460 Fig. 8

P23 L485 Spatial variation of photophysiology?

P24 L512 Provide the reference of “model predictions”.

P24 L517 Correct notation Fv/Fm as in Table 1.

P28 L598-625 These paragraphs should be included in Methods.

P32 L683- Follow the style in journal guidelines.

Deblois C.P., Marchand A. & Juneau P. (2013). Comparison of photoacclimation in twelve freshwater photoautotrophs (chlorophyte, bacillaryophyte, cryptophyte and cyanophyte) isolated from a natural community. PLOS ONE 8, e57139. https://doi.org/10.1371/journal.pone.0057139

Gorbunov M.Y. & Falkowski P.G. (2021). Using chlorophyll fluorescence kinetics to determine photosynthesis in aquatic ecosystems. Limnology and Oceanography 66, 1–13. https://doi.org/10.1002/lno.11581

Hughes D., Campbell D., Doblin M.A., Kromkamp J., Lawrenz E., Moore C.M., et al. (2018a). Roadmaps and detours: active chlorophyll-a assessments of primary productivity across marine and freshwater systems. Environmental Science & Technology 52, 12039–12054. https://doi.org/10.1021/acs.est.8b03488

Hughes D.J., Varkey D., Doblin M.A., Ingleton T., Mcinnes A., Ralph P.J., et al. (2018b). Impact of nitrogen availability upon the electron requirement for carbon fixation in Australian coastal phytoplankton communities. Limnology and Oceanography 63, 1891–1910. https://doi.org/10.1002/lno.10814

Kazama T., Hayakawa K., Kuwahara V.S., Shimotori K., Imai A. & Komatsu K. (2021). Development of photosynthetic carbon fixation model using multi-excitation wavelength fast repetition rate fluorometry in Lake Biwa. PLOS ONE 16, e0238013. https://doi.org/10.1371/journal.pone.0238013

Lawrenz E., Silsbe G., Capuzzo E., Ylöstalo P., Forster R.M., Simis S.G.H., et al. (2013). Predicting the electron requirement for carbon fixation in seas and oceans. PLoS ONE 8, e58137. https://doi.org/10.1371/journal.pone.0058137

Perkins R.G., Kromkamp J.C., Serôdio J., Lavaud J., Jesus B., Mouget J.L., et al. (2010). The Application of Variable Chlorophyll Fluorescence to Microphytobenthic Biofilms. In: Chlorophyll a Fluorescence in Aquatic Sciences: Methods and Applications. Developments in Applied Phycology, (Eds D.J. Suggett, O. Prášil & M.A. Borowitzka), pp. 237–275. Springer Netherlands, Dordrecht.

Regaudie-de-Gioux A., Lasternas S., Agustí S. & Duarte C.M. (2014). Comparing marine primary production estimates through different methods and development of conversion equations. Frontiers in Marine Science 1, 19. https://doi.org/10.3389/fmars.2014.00019

Robinson C., Tilstone G.H., Rees A.P., Smyth T.J., Fishwick J.R., Tarran G.A., et al. (2009). Comparison of in vitro and in situ plankton production determinations. Aquatic Microbial Ecology 54, 13–34

Ryan-Keogh T.J., Thomalla S.J., Little H. & Melanson J.-R. (2018). Seasonal regulation of the coupling between photosynthetic electron transport and carbon fixation in the Southern Ocean. Limnology and Oceanography 63, 1856–1876. https://doi.org/10.1002/lno.10812

Schuback N., Hoppe C.J.M., Tremblay J.-É., Maldonado M.T. & Tortell P.D. (2017). Primary productivity and the coupling of photosynthetic electron transport and carbon fixation in the Arctic Ocean. Limnology and Oceanography 62, 898–921. https://doi.org/10.1002/lno.10475

Schuback N., Tortell P.D., Berman-Frank I., Campbell D.A., Ciotti A., Courtecuisse E., et al. (2021). Single-turnover variable chlorophyll fluorescence as a tool for assessing phytoplankton photosynthesis and primary productivity: opportunities, caveats and recommendations. Frontiers in Marine Science 0. https://doi.org/10.3389/fmars.2021.690607

Suggett D.J., Kraay G., Holligan P., Davey M., Aiken J. & Geider R. (2001). Assessment of photosynthesis in a spring cyanobacterial bloom by use of a fast repetition rate fluorometer. Limnology and Oceanography 46, 802–810. https://doi.org/10.4319/lo.2001.46.4.0802

Suggett D.J., MacIntyre H.L., Kana T.M. & Geider R.J. (2009). Comparing electron transport with gas exchange: parameterising exchange rates between alternative photosynthetic currencies for eukaryotic phytoplankton. Aquatic Microbial Ecology 56, 147–162. https://doi.org/10.3354/ame01303

Zhu Y., Ishizaka J., Tripathy S., Wang S., Sukigara C., Goes J., et al. (2017). Relationship between light, community composition and the electron requirement for carbon fixation in natural phytoplankton. Marine Ecology Progress Series 580, 83–100. https://doi.org/10.3354/meps12310

Reviewer #2: Review of PLOS ONE

Chlorophyll fluorescence-based estimates of photosynthetic electron transport in Arctic

phytoplankton assemblages Aug 29 2021

General comments

The analysis is, to my knowledge sound and robust, but I have some issues with the narrative. I think this stems to some extent to the introduction not really clearly presenting the authors’ hypotheses. I don’t think all science needs to be hypothesis-based but here we can identify several research questions but I don’t feel I got a good sense of what was anticipated and why. The end of the introduction would be where I would expect to find this, rather than the recap of key methods/results/discussion in this iteration. I would also recommend highlighting the biggest finding in the title instead of a general description of what was done. The title had me looking forwards to the “assemblages” part but since it wasn’t possible to address community composition in any way with this dataset I wonder if there isn’t a more just way to phrase this.

Another question that I would like to raise relates to the relationship between photochemical efficiency, environmental stress and species composition. In my experience, using lab cultures, fluorescence can vary a lot between taxonomic groups despite ressource replete and exponential growth conditions (notably I’d consider FvFm values to be higher for diatoms and greens and ower for (pico)cyanobacteria and heterotrophs). I guess I really feel like knowledge of species composition at lower resolution is necessary at this point to distinguish whether we are measuring stress or community changes as a result of changes in the environment (including stress!). Although I don’t think there is a whole lot of literature discussing this, I would be happy to see some support to whatever position is taken.

Finally I would have been interested in knowing how high resolution the data needs to be to have adequate dark acclimation. How many fewer samples would be analyzed with a longer dark acclimation? (10, 20, 30+ min) and how would this affect the results (maybe this adds too much to the text but comparing the statistical power of the tests used and by subsampling the current set based on different scenarioswould be one way to elegantly present this...).

If PLOS ONE does supplementary material: Do the hydrographic variables for each station need to be presented in the main text? This data is great to make available but I might stick to presenting the data directly relevant to the analyses (graphically to highlight differences between LS and BS) and leave the rest out of the main text.

Detailed comments:

Make sure the tables and figures are presented in order of appearance (I think Fig 8 is mentionned very early on). Spell check the tables.

line 318: Is it necessary to say data wasn’t corrected for autocorrelation? Otherwise specify why this wasn’t necessary.

Figures are a bit fuzzy...can this be improved?

Fig1: Should have insert of greater geographical scope to better situate sites

Fig2 and 3: grey highlight should be under curve

Fig 3: match axis colors to the colors of the curve? Improve axis clarity (put long labels on 2 lines?)

Fig4 and 5: If possible add confidence interval? Why weren’t the rank correlations calculated for BS and LS separately if we are interested in comparing these two locales

Fig 7: Fv/Fm seems related to the relationship between ETRk and ETRa...except for the steepest slope of points which has very high Fv/Fm values. This is presented in the results but isn’t discussed as far as I can tell. I would be curious to know more about why this site is different.

Fig8: hard to interpret. Could sites close by be averaged to better distinguish between day/night sites? Which could then be kept of similar size?

Reviewer #3: This is an interesting study. The article is well written and comprehensive in its interpretation of the data. With some minor revisions, detailed below, I believe it will be suitable for publication in PlosOne.

One concern I have with the paper is the use of the term “primary productivity” throughout the manuscript when speaking of the FRRf measurements. Generally (though not exclusively), this term is used to refer to either oxygen production or carbon fixation. While electron transport rates are correlated with these, they are not synonymous with them. As the authors are very well aware given their previous work, the difficulty involved in converting measurements of electron transport rates to carbon fixation has been the major stumbling block to utilizing shipboard FRRf measurements to estimate productivity in oceanographic studies. Thus, the use of the term primary production to refer to electron transport measurements derived from either of the two ETR algorithms is misleading. This is particularly so given that no effort is made in the study to directly measure photosynthetic oxygen production or carbon fixation or to determine which of the two algorithms provides a better estimate of either (as the authors point out, this is important for future work). As such, I believe it would be better to use terms like “electron transport rates” or “primary photochemistry” when referring to the measurements derived from the FRRf.

Other than that, there several minor issues within the paper that should be corrected.

Line 261: Change “ETRk” to “ETRk”.

Line 309: There are several instances in the paper where Fv/Fm is not italicized. Please make sure that italics are consistently used.

Line 394: Italicize “Fv/Fm”.

Line 460: Change “Fig8” to “Fig 8”.

Line 460: Italicize “Fv/Fm”.

Line 487: This paragraph could be cleaned up slightly. Specifically, the sentence in line 500, “we infer that nitrogen, rather than iron deficiency was the most likely cause of low photo-efficiencies”, is something of a repeat of the sentence at line 493, “suggest that the low Fv/Fm values we observed likely reflect nitrogen deficiency.”

Line 584: Change “F’q/F’m” to “Fq’/Fm’ ”.

Line 591: Change “nutrient limiting” to “nutrient-limiting”.

Line 657: Change “NQP” to “NPQ”.

Line 660: Change “F’q/F’m” to “Fq’/Fm’ ”.

Figure 3: The y-axis label for Figure E (Fq’/Fm’ 150) is somewhat confusing to read. The 150 is in line with the label for Figure D and at first it was unclear which figure it belonged to. Adjusting the figure so that the labels are not in line with or so close to each other would make it easier to read.

Figure 4: correct uE to �E in x-axis labels.

**********

6. PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #1: Yes: Takehiro Kazama

Reviewer #2: No

Reviewer #3: No

[NOTE: If reviewer comments were submitted as an attachment file, they will be attached to this email and accessible via the submission site. Please log into your account, locate the manuscript record, and check for the action link "View Attachments". If this link does not appear, there are no attachment files.]

While revising your submission, please upload your figure files to the Preflight Analysis and Conversion Engine (PACE) digital diagnostic tool, https://pacev2.apexcovantage.com/. PACE helps ensure that figures meet PLOS requirements. To use PACE, you must first register as a user. Registration is free. Then, login and navigate to the UPLOAD tab, where you will find detailed instructions on how to use the tool. If you encounter any issues or have any questions when using PACE, please email PLOS at figures@plos.org. Please note that Supporting Information files do not need this step.

Attachment

Submitted filename: PONE-D-21-24983_Review.docx

PLoS One. 2021 Dec 9;16(12):e0256410. doi: 10.1371/journal.pone.0256410.r002

Author response to Decision Letter 0


17 Oct 2021

Dear Dr. Carvalho,

We wish to thank you for your consideration of this manuscript, the careful attention to detail apparent in comments from yourself and the Reviewers, and for the opportunity to revise. All requested changes (below) ultimately proved to be relatively minor, not changing the interpretation of the data or conclusion drawn. We provide our detailed point by point response as follows, providing further explanation where necessary. References to page and line numbers below correspond to the edited document containing tracked changes. We are confident that our responses address the insightful concerns raised, and have further improved our manuscript.

Yours Sincerely,

Yayla Sezginer

Responses to the Editor

1. Please ensure that your manuscript meets PLOS ONE's style requirements, including those for file naming. The PLOS ONE style templates can be found at https://journals.plos.org/plosone/s/file id=wjVg/PLOSOne_formatting_sample_main_body.pdf and https://journals.plos.org/plosone/s/file id=ba62/PLOSOne_formatting_sample_title_authors_affiliations.pdf

We have amended the manuscript to meet these style and file naming conventions.

2. In your Methods section, please provide additional information regarding the permits you obtained for the work. Please ensure you have included the full name of the authority that approved the field site access and, if no permits were required, a brief statement explaining why.

Amundsen Science and the CCGS Amundsen were granted the following permits for our expedition:

-Nunavut Research Institute Scientific Research License # 0501119R-M

-Department of Fisheries and Oceans License to Fish for Scientific Purposes in the waters of Nunavut # S-19/20-1016-NU

-Parks Canada – National Parks Auyuittuq, Sirmilik and Quttirnipaaq # ANP-2019-32477 Canadian Wildlife Service – Access to Migratory Birds sanctuary on Bylot Island # MM-NR-2019-NU-014

-Canadian Wildlife Service – Access to National Wildlife Areas at Coburg Island, Akpait and Qaqulluit # NF-NR-2019-NU-008

-Vessel Clearance to conduct scientific work in Greenland waters. Danish Ministry of Foreign Affairs file # 2019-2308

-Permission for State Flight over Greenland File # 19/01017

-Government of Greenland Survey License # G19-036

Of these various permits, only the Nunavut Research Institute Scientific Research License #0501119R-M was required for the operations discussed in this manuscript. This license is now cited on page 7, line 180 in the Methods.

3. In your Methods section, please provide additional location information, including geographic coordinates for the data set if available.

The complete dataset of geographic coordinates where underway sampling took place is included in the salinity and temperature dataset available in the Polar Data Catalogue (doi: 10.5884/12715). A reference to the geographic coordinates in now provided in the study region map figure legend (Fig 1).

4. We note that you have stated that you will provide repository information for your data at acceptance. Should your manuscript be accepted for publication, we will hold it until you provide the relevant accession numbers or DOIs necessary to access your data. If you wish to make changes to your Data Availability statement, please describe these changes in your cover letter and we will update your Data Availability statement to reflect the information you provide

Salinity and temperature data collected by Amundsen Science are available in the Polar Data Catalogue (doi: 10.5884/12715). Surface photosynthetically active radiation data are available in the Polar Data Catalogue (doi: 10.5884/12518), alongside additional meteorological data provided by the Amundsen Science group of U. Laval. CTD station conductivity, temperature, and salinity data provided by the Amundsen Science group of U. Laval are available in the Polar Data Catalogue (doi: 10.5884/12713). Underway oxygen saturation and ΔO2/Ar data are available in the Polar Data Catalogue (doi: 10.5884/13242). All Fast Repetition Rate Fluorometry measurements of phytoplankton photophysiology have been uploaded to the Polar Data Catalogue and are currently awaiting approval. The Canadian Cryospheric Information Network reference number for this dataset is 13254, and will be trackable using this number once published. A doi will be provided as soon as available.

5. Please update your submission to use the PLOS LaTeX template. The template and more information on our requirements for LaTeX submissions can be found at http://journals.plos.org/plosone/s/latex.

We have revised the manuscript format using the PLOS LaTeX template. The reformatted manuscript and latex source code have both been uploaded under the file names Manuscript.pdf and ManuscriptSource.tex, respectively.

6. We note that Figures 1 and 9 in your submission contain [map/satellite] images which may be copyrighted. All PLOS content is published under the Creative Commons Attribution License (CC BY 4.0), which means that the manuscript, images, and Supporting Information files will be freely available online, and any third party is permitted to access, download, copy, distribute, and use these materials in any way, even commercially, with proper attribution. For these reasons, we cannot publish previously copyrighted maps or satellite images created using proprietary data, such as Google software (Google Maps, Street View, and Earth). For more information, see our copyright guidelines: http://journals.plos.org/plosone/s/licenses-and-copyright.

Both map figures were produced in Matlab R2020a using the freely available m_map package available online at www.eoas.ubc.ca/~rich/map.html. The package includes a coastline, global elevation database, and river database, and provides access to publicly available bathymetry data. The data are not proprietary. A citation for m_map has been added to the map figure legend and the references section (Fig 1).

Responses to Reviewer 1:

“The major point is that there is no support of apparent (measured) O2 evolution and C assimilation rate for the two ETRs in the study area. The amplitude based ETRa have been developed as biophysical models to derive gross primary production (GPP) because FRRf can reduces only QA but not PQ pool and thus examine the electron flow rate for oxygen evolution. However, ETRa-based estimated GPP does not always equal to apparent GPP (Regaudie-de- Gioux et al., 2014). Many previous studies have investigated the factor affecting the relationships between modeled ETRa and measured O2 evolution rate (Suggett et al., 2001; Robinson et al., 2009; Deblois, Marchand & Juneau, 2013) and C assimilation rate in natural communities (Lawrenz et al., 2013; Schuback et al., 2017; Zhu et al., 2017; Hughes et al., 2018b; Ryan-Keogh et al., 2018; Kazama et al., 2021). For oxygen-based GPP, it is well confirmed for cultivated species (Suggett et al., 2009) but species composition can affect conversion factor in natural communities (Deblois et al., 2013). For, C-based GPP, although there are many studies but no global model is derived to convert ETRa to GPP (Hughes et al., 2018a). On the other hand, authors’ models is not supported by oxygen production rate or C assimilation of algae yet. The kinetic based ETRk by FIRe system (FIRe-ETRk) is a novel method to improve the errors from the parameters (Gorbunov & Falkowski, 2021). The relationships between the FIRe-ETRk and the growth rate (net primary production, NPP) of two model species, Thalassiosira pseudonana, and Dunaliella tertiolecta were examined, but not in natural community yet. Also the relationships between FIRe-ETRk and GPP is not examined yet. Therefore, the use of authors’ kinetic model in natural algal communities without reference measured productivity (O2 or C) must be considered carefully, even if it is likely analogous to the FIRe-ETRk. Because authors’ models is not supported by oxygen production rate of algae yet, the absolute value of ETRk cannot compare to ETRa. For example, when these two ETR values are used as estimators of GPP and NPP, respectively, in the study area, the results are still questionable. Because net productivity does not include respiration, ETRk must be lower than ETRa on the same phytoplankton assemblage. However, present study clearly showed ETRk > ETRa in Lancaster Sound (Fig. 8). This paradoxical results may be due to underestimation of ETRa, and/or overestimation of ETRk. The former is likely due to inadequate dark acclimation time (5 min, P11 L211). For rapid light curve, there is no consensus rule but typically 10~20 min of dark adaptation period is used (Schuback et al., 2021).The latter is likely due to the underestimation of turnover rate, as authors pointed out. If authors primarily focus to the comparison of relative oxygen productivity between Lancaster Sound and Barrow Strait, use Fv/Fm and ETRa but not ETRk.”

The Reviewer makes excellent points but appears to have mis-interpreted the aims and scope of our study. We agree that electron transport rates (ETR) are not equivalent to oxygen evolution or carbon uptake. Indeed, as noted by the reviewer, there are many photosynthetic and metabolic processes that can lead to the decoupling of electron transport from oxygen evolution and downstream carbon assimilation. However, it was not our intention in this article to use ETR as an absolute metric of primary productivity. Rather, as we state in the introduction (P6 L150), our goal is to compare different approaches to estimating ETR, examining how the two contrasting algorithms perform across a range of hydrographic regimes. As pointed out by the Reviewer, there is a significant need for work of this kind in natural environments, moving beyond current results that have focused on laboratory studies. Even without ‘calibration’ against independent productivity bench-marks (O2 evolution, CO2 uptake), ETR is an important diagnostic of photosynthetic capacity in its own right. The reducing power produced by this process fuels a number of metabolic functions, and presents an upper limit for carbon fixation by the Calvin-Benson cycle. We understand that the comments of the reviewer may have reflected a lack of clarity in our initial wording. We therefore have altered our wording throughout from ‘primary productivity’ to ‘primary photochemistry’ to remove any suggestion that ETR is interchangeable with carbon uptake or oxygen evolution. We wholeheartedly agree that validating ETR measurements with parallel measures of O2 evolution would enable us to assess the accuracy of the different ETR algorithms. This has now been further emphasized in the discussion section, where we suggest such work for a follow up study (P3 L862). The reviewer also raises some concerns about the effects of respiration on our results, suggesting that ETRk and ETRa measure NPP and GPP, respectively. This is, in fact, incorrect, as both ETR algorithms provide estimates of GPP. The kinetic ETR algorithm recently developed by Gorbunov and Falkowski (2021) evaluates the rate of electron transport through Photosystem II (PSII) as the normalized photochemical yield at a given light intensity multiplied by the rate at which the primary electron acceptor, Qa is capable of turning over electrons from the photosynthetic Reaction Center II to the secondary electron acceptor, Qb. As such, ETRk measures primary photochemistry and is theoretically equivalent to ETRa, which measures the rate of electron transport through PS II as the product of the photochemical energy conversion efficiency, the functional light absorption area of PSII, and the provided light intensity. Although Gorbunov and Falkowski (2021) used the high correlation between lab-grown phytoplankton

growth rates and ETRk to validate their novel kinetic-based ETR algorithm, ETRk is not a measure of NPP, which is the difference between GPP and respiration.

Minor points

P2 L25 fast repetition rate fluorometry

->No change, we prefer to keep the capitalization for consistency with the literature.

P5 L93 photosynthetically active radiation

->Resolved. (P5 L106)

P8 L157 Provide company name and city for every instruments of the manuscript.

->Provided company name where missing. Did not include city name, to keep with style of

previous PLOS One publications (see Schuback et al., 2015)

P8 L178 Unify the unit of whole manuscript (s, min).

->No change (unless the Editor and Editorial process requires) since our results capture processes

occurring on differing time-scales, we prefer to use different units of time, as appropriate.

P10 L207 Provide the version and company of the software.

->Resolved company name. Software version/name provided.

P10 L209 Table 1 What is the ChlF?

->Replaced with Chla

Ų PSII−1, Definition of Fq’/Fm’ (max) should not be “under 150 μmol quanta m-2 s-1”.

->Resolved.

P11 L217 Provide the duration of each acquisitions and total time per sample.

->Resolved. P10 L253

P11 L218 Use mol instead of E.

->Resolved.

P11 L226 Provide the duration of each light step. For rapid light curve, less than 30 s is

recommended (Perkins et al., 2010).

->Resolved. P10 L260

P13 L254-258 Ų PSII−1

->Resolved.

P13 L262 Provide the reference for 3 component multi-exponential model.

->Resolved. P13 L325

Equation (2) What are the F(t) and CQA(t)?

->Resolved. Measured fluorescence and fraction of RCIIs closed by excitation flashlets. P14 L345

P13 L273-276 QA, QB

->Set to Qa and Qb throughout, removed all instances of QA and QB

P16 L319 Provide version, company and city of the software.

->Added company. Version included. (P13 L342)

P17 L349 Table 2 Correct notation Chl a.

->Resolved. Table 2 moved to the supplements.

P17 L354 Table 3 Does the zero mean true zero or lower than detection limit?

->Zeros do indicate below detection limit. Values below detection limit in have been updated to <

0.02 uM. Table 3 has been moved to the Supporting Information section (S2 Table)

P19 L387 Spearman’s rank correlation. Use ρ (rho) or “rs” instead of R. See Schober et al. 2018.

Correlation Coefficients: Appropriate Use and Interpretation, Anesthesia & Analgesia: Volume

126, 5, 1763-1768. doi: 10.1213/ANE.0000000000002864

->Resolved throughout.

P20 L403 Use 25th and 75th percentile, or range with median value, instead of SD.

->Deviation from the median reported is the median absolute deviation, not the SD.

P20 L410-415 Integrate this part into discussion.

->We prefer to retain this aspect in the results, as these lines provide a quick operational note,

rather than a discussion of key findings.

P20 L417 Explanation is needed why 150 μmol quanta m-2 s-1 is used.

->Explanation added to Methods (P11 L283)

P21 L423-426 Integrate this part into discussion.

->Agreed. We have moved the majority of text outlining our interpretations of Fig 5 to P24 L672-

678 in the Discussion section.

P22 L460 Fig. 8

->Resolved

P23 L485 Spatial variation of photophysiology?

->Adopted suggestion. (P22 L609)

P24 L512 Provide the reference of “model predictions”.

->Resolved. (P23 L637)

P24 L517 Correct notation Fv/Fm as in Table 1.

->Resolved throughout.

P28 L598-625 These paragraphs should be included in Methods

->After some consideration, we are confident that these paragraphs belong in the discussion, since

they provide critical context to understand how differing numerical procedures may affect

results. Further, these lines include a description of an analogous method that we did not apply -

including this in the Methods could lead to some confusion in what we actually do, as opposed to

what we contrast against.

P32 L683- Follow the style in journal guidelines.

->Resolved as per Editor’s comments.

Responses to Reviewer 2:

“The analysis is, to my knowledge sound and robust, but I have some issues with the narrative. I think this stems to some extent to the introduction not really clearly presenting the authors’ hypotheses. I don’t think all science needs to be hypothesis-based but here we can identify several research questions but I don’t feel I got a good sense of what was anticipated and why. The end of the introduction would be where I would expect to find this, rather than the recap of key methods/results/discussion in this iteration. I would also recommend highlighting the biggest finding in the title instead of a general description of what was done. The title had me looking forwards to the “assemblages” part but since it wasn’t possible to address community composition in any way with this dataset I wonder if there isn’t a more just way to phrase this. Another question that I would like to raise relates to the relationship between photochemical efficiency, environmental stress and species composition. In my experience, using lab cultures, fluorescence can vary a lot between taxonomic groups despite ressource replete and exponential growth conditions (notably I’d consider FvFm values to be higher for diatoms and greens and ower for (pico)cyanobacteria and heterotrophs). I guess I really feel like knowledge of species composition at lower resolution is necessary at this point to distinguish whether we are measuring stress or community changes as a result of changes in the environment (including stress!). Although I don’t think there is a whole lot of literature discussing this, I would be happy to see some support to whatever position is taken. Finally I would have been interested in knowing how high resolution the data needs to be to have adequate dark acclimation. How many fewer samples would be analyzed with a longer dark

acclimation? (10, 20, 30+ min) and how would this affect the results (maybe this adds too much to the text but comparing the statistical power of the tests used and by subsampling the current set based on different scenarios would be one way to elegantly present this...). If PLOS ONE does supplementary material: Do the hydrographic variables for each station need to be presented in the main text? This data is great to make available but I might stick to presenting the data directly relevant to the analyses (graphically to highlight differences between LS and BS) and leave the rest out of the main text.”

The Reviewer raises important points that we have addressed:

1. We have added our research question and objective to the introductory paragraph (P6 L150), and refined our title, as suggested, to ensure more transparency in what was anticipated and why.

2. We agree entirely with the Reviewers point, and have shown previously from broad analysis of numerous cultures (e.g. Suggett et al. 2009) that diatoms and greens result in very different inherent FRRf parameterization (e.g. higher Fv/Fm) than pico and nano phytoplankton, but also that these trends are hard to deconvolve from patterns of resource availability for photosynthesis. We therefore argue that variability in Fv/Fm results from both direct physiological effects of nutrient limitation, and indirect effects that increase the abundance of smaller phytoplankton taxa (with typically lower Fv/Fm) in nutrient limited environments. We make this point on P24 L680-714, and acknowledge the unfortunate lack of species composition data in our study to fully resolve this issue.

3. We recently conducted additional experiments to address the question of dark acclimation time-scales. (These experiments were conducted this past year, i.e. on a different cruise from that presented in this paper). We allowed samples to dark acclimate over 30 minutes and took FRRf measurements every 2 minutes to assess Fv/Fm and 𝜎!"## recovery throughout the course of the dark acclimation period. Samples required a mean recovery time of 9.65 ± 1.53 min (n = 34) to reach stable, maximum light absorption efficiencies. The sampling approach used in this current study (2019 data) used a 5 min acclimation time. Increasing this to 10 minutes would increase the overall sampling interval to 17 minutes, which would have reduced our sample size from 481 to 340, representing a 29.5% reduction in measurement resolution. These preliminary results have been added to the Discussion section (P26 L749-764).

4. We prefer to keep the hydrographic data in the main text to provide context for the physical properties of our sampling region. We have moved the table of correlation statistics between hydrographic variables to the supplemental section, as we agree that it is not critical to the narrative presented. All the hydrographic data is publicly available on the Polar Data Catalogue (doi: 10.5884/12715).

Detailed comments:

Make sure the tables and figures are presented in order of appearance (I think Fig 8 is mentioned very early on). Spell check the tables.

->Resolved.

line 318: Is it necessary to say data wasn’t corrected for autocorrelation? Otherwise specify why this wasn’t necessary.

->We have removed this line.

Figures are a bit fuzzy...can this be improved?

->High resolution versions of the figures can be downloaded by clicking the link at the top right corner of the figure page.

Fig1: Should have insert of greater geographical scope to better situate sites

->Resolved.

Fig2 and 3: grey highlight should be under curve

->Resolved.

Fig 3: match axis colors to the colors of the curve? Improve axis clarity (put long labels on 2 lines?)

->Resolved.

Fig4 and 5: If possible add confidence interval? Why weren’t the rank correlations calculated for BS and LS separately if we are interested in comparing these two locales

->Resolved. Computing separate correlations for BS and LS has led to some new discussion of the differences between these two regions, particularly regarding Fq’/Fm’(150) and PAR relationship. We interpret the positive relationship between Fq’/Fm’ (150) and PAR in Barrow Strait as an indicator of the increased ability of phytoplankton in Barrow Strait to acclimate to higher light levels. In contrast, the lack of a relationship between Fq’/Fm’ (150) and PAR in Lancaster Sound is further evidence of possible photodamage to reaction centers, likely due to insufficient N

required to repair non-functional reaction centers. We discuss these results on P24 L672-678.

Fig 7: Fv/Fm seems related to the relationship between ETRk and ETRa...except for the steepest

slope of points which has very high Fv/Fm values. This is presented in the results but isn’t

discussed as far as I can tell. I would be curious to know more about why this site is different.

->The unusual site where Fv/Fm is very high, but has a large decoupling ratio between ETRk:ETRa is unique in that it has a very low Fq’/Fm’emax value ( < 0.04) despite its high Fv/Fm value ( > 0.42). Fv/Fm and Fq’/Fm’emax are highly correlated (𝜌 = 0.93 in our dataset), so it is likely this site is an outlier given we have one of our highest Fv/Fm and lowest Fq’/Fm’emax values here. This explanation is now given on P27 L792-809.

Fig8: hard to interpret. Could sites close by be averaged to better distinguish between day/night sites? Which could then be kept of similar size?

->After spending some time rethinking how to best represent the data presented in Fig 9 (we believe Reviewer 2 is actually commenting on Fig 9), we’ve resolved to keep the figure as is. We believe that averaging samples across broad areas would obscure small-scale features in photophysiological spatial variability, which we believe is one of the unique and valuable aspects of our study.

Responses to Reviewer 3:

“One concern I have with the paper is the use of the term “primary productivity” throughout the manuscript when speaking of the FRRf measurements. Generally (though not exclusively), this term is used to refer to either oxygen production or carbon fixation. While electron transport rates are correlated with these, they are not synonymous with them. As the authors are very well aware given their previous work, the difficulty involved in converting measurements of electron transport rates to carbon fixation has been the major stumbling block to utilizing shipboard FRRf measurements to estimate productivity in oceanographic studies. Thus, the use of the term

primary production to refer to electron transport measurements derived from either of the two ETR algorithms is misleading. This is particularly so given that no effort is made in the study to directly measure photosynthetic oxygen production or carbon fixation or to determine which of the two algorithms provides a better estimate of either (as the authors point out, this is important for future work). As such, I believe it would be better to use terms like “electron transport rates” or “primary photochemistry” when referring to the measurements derived from the FRRf.”

In agreement also with Reviewer 1’s first comment, we have adopted the Reviewer 3’s suggestion of switching our language from ‘primary productivity’ to ‘primary photochemistry’ to avoid confusion with carbon or oxygen-based productivity measures.

Line 261: Change “ETRk” to “ETRk”.

->Resolved.

Line 309: There are several instances in the paper where Fv/Fm is not italicized. Please make sure that italics are consistently used.

->Resolved.

Line 460: Change “Fig8” to “Fig 8”.

->Resolved.

Line 487: This paragraph could be cleaned up slightly. Specifically, the sentence in line 500, “we infer that nitrogen, rather than iron deficiency was the most likely cause of low photoefficiencies”, is something of a repeat of the sentence at line 493, “suggest that the low Fv/Fm values we observed likely reflect nitrogen deficiency.”

->Removed repetitive statements.

Line 584: Change “F’q/F’m” to “Fq’/Fm’ ”.

->Resolved

Line 591: Change “nutrient limiting” to “nutrient-limiting”.

->Resolved.

Line 657: Change “NQP” to “NPQ”.

->Resolved.

Line 660: Change “F’q/F’m” to “Fq’/Fm’ ”.

->Resolved.

Figure 3: The y-axis label for Figure E (Fq’/Fm’ 150) is somewhat confusing to read. The 150 is in line with the label for Figure D and at first it was unclear which figure it belonged to. Adjusting the figure so that the labels are not in line with or so close to each other would make it easier to read.

->Resolved.

Figure 4: correct uE to μE in x-axis labels.

->Resolved.

Once again, we thank each of the Reviewers for their insightful questions and comments, which we feel have enhanced the quality of this manuscript.

Attachment

Submitted filename: Response to Reviewers.pdf

Decision Letter 1

Matheus C Carvalho

17 Nov 2021

Irradiance and nutrient-dependent effects on photosynthetic electron transport in Arctic phytoplankton: a comparison of two Chlorophyll fluorescence-based approaches to derive primary photochemistry

PONE-D-21-24983R1

Dear Dr. Sezginer,

We’re pleased to inform you that your manuscript has been judged scientifically suitable for publication and will be formally accepted for publication once it meets all outstanding technical requirements.

Within one week, you’ll receive an e-mail detailing the required amendments. When these have been addressed, you’ll receive a formal acceptance letter and your manuscript will be scheduled for publication.

An invoice for payment will follow shortly after the formal acceptance. To ensure an efficient process, please log into Editorial Manager at http://www.editorialmanager.com/pone/, click the 'Update My Information' link at the top of the page, and double check that your user information is up-to-date. If you have any billing related questions, please contact our Author Billing department directly at authorbilling@plos.org.

If your institution or institutions have a press office, please notify them about your upcoming paper to help maximize its impact. If they’ll be preparing press materials, please inform our press team as soon as possible -- no later than 48 hours after receiving the formal acceptance. Your manuscript will remain under strict press embargo until 2 pm Eastern Time on the date of publication. For more information, please contact onepress@plos.org.

Kind regards,

Matheus C. Carvalho

Academic Editor

PLOS ONE

Additional Editor Comments (optional):

Reviewers' comments:

Reviewer's Responses to Questions

Comments to the Author

1. If the authors have adequately addressed your comments raised in a previous round of review and you feel that this manuscript is now acceptable for publication, you may indicate that here to bypass the “Comments to the Author” section, enter your conflict of interest statement in the “Confidential to Editor” section, and submit your "Accept" recommendation.

Reviewer #1: All comments have been addressed

Reviewer #2: All comments have been addressed

Reviewer #3: All comments have been addressed

**********

2. Is the manuscript technically sound, and do the data support the conclusions?

The manuscript must describe a technically sound piece of scientific research with data that supports the conclusions. Experiments must have been conducted rigorously, with appropriate controls, replication, and sample sizes. The conclusions must be drawn appropriately based on the data presented.

Reviewer #1: Yes

Reviewer #2: Yes

Reviewer #3: (No Response)

**********

3. Has the statistical analysis been performed appropriately and rigorously?

Reviewer #1: Yes

Reviewer #2: Yes

Reviewer #3: (No Response)

**********

4. Have the authors made all data underlying the findings in their manuscript fully available?

The PLOS Data policy requires authors to make all data underlying the findings described in their manuscript fully available without restriction, with rare exception (please refer to the Data Availability Statement in the manuscript PDF file). The data should be provided as part of the manuscript or its supporting information, or deposited to a public repository. For example, in addition to summary statistics, the data points behind means, medians and variance measures should be available. If there are restrictions on publicly sharing data—e.g. participant privacy or use of data from a third party—those must be specified.

Reviewer #1: Yes

Reviewer #2: Yes

Reviewer #3: (No Response)

**********

5. Is the manuscript presented in an intelligible fashion and written in standard English?

PLOS ONE does not copyedit accepted manuscripts, so the language in submitted articles must be clear, correct, and unambiguous. Any typographical or grammatical errors should be corrected at revision, so please note any specific errors here.

Reviewer #1: Yes

Reviewer #2: Yes

Reviewer #3: (No Response)

**********

6. Review Comments to the Author

Please use the space provided to explain your answers to the questions above. You may also include additional comments for the author, including concerns about dual publication, research ethics, or publication ethics. (Please upload your review as an attachment if it exceeds 20,000 characters)

Reviewer #1: The new title is explaining the concept of this paper concisely. The issue of the uncertainty in the relationship between ETR and productivity has been solved by shifting to “primary photochemistry” instead of “primary productivity”. The methods and results are well-documented. I think this revised manuscript is good to publish after fixing some minor points such as follows.

Minor points:

Title: chlorophyll

Table 1 and main body: The term p can be easily confused with the p of the p-value. Please change either of these.

L183: 30 s

Table 2: Use the minus sign despite the hyphen.

L420: CO2

L840-L851: CQa(t)

Reviewer #2: My concerns with the manuscript have been adequately addressed. I might suggest another mention of the potential role of taxonomy to better unravel the relationship between photophysiology, fluorescence and productivity in the conclusion section but this is potentially more indicative of my particular biases.

Reviewer #3: (No Response)

**********

7. PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #1: Yes: Takehiro Kazama

Reviewer #2: No

Reviewer #3: No

Acceptance letter

Matheus C Carvalho

1 Dec 2021

PONE-D-21-24983R1

Irradiance and nutrient-dependent effects on photosynthetic electron transport in Arctic phytoplankton: a comparison of two Chlorophyll fluorescence-based approaches to derive primary photochemistry

Dear Dr. Sezginer:

I'm pleased to inform you that your manuscript has been deemed suitable for publication in PLOS ONE. Congratulations! Your manuscript is now with our production department.

If your institution or institutions have a press office, please let them know about your upcoming paper now to help maximize its impact. If they'll be preparing press materials, please inform our press team within the next 48 hours. Your manuscript will remain under strict press embargo until 2 pm Eastern Time on the date of publication. For more information please contact onepress@plos.org.

If we can help with anything else, please email us at plosone@plos.org.

Thank you for submitting your work to PLOS ONE and supporting open access.

Kind regards,

PLOS ONE Editorial Office Staff

on behalf of

Dr. Matheus C. Carvalho

Academic Editor

PLOS ONE

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Supplementary Materials

    S1 Table. Correlation of underway hydrographic variables.

    Results of Spearman Rank correlation analyses between each underway hydrographic variable are displayed. ** is used to indicate p values < 0.001. In all instances n = 7200.

    (PDF)

    S2 Table. CTD profiling stations.

    Sampling station locations are indicated as LS for Lancaster Sound or BS for Barrow Strait. Station mixed layer depth (MLD), mean mixed layer nitrate and nitrite concentration, and Chl a concentration within the mixed layer are shown.

    (PDF)

    S3 Table. Surface PAR and underway photophysiological variable regression analyses.

    All p values were < 0.001. Standard error is reported for the regression intercept and coefficient, respectively.

    (PDF)

    Attachment

    Submitted filename: PONE-D-21-24983_Review.docx

    Attachment

    Submitted filename: Response to Reviewers.pdf

    Data Availability Statement

    Salinity and temperature data collected by Amundsen Science are available in the Polar Data Catalogue (doi: 10.5884/12715). Surface photosynthetically active radiation data are available in the Polar Data Catalogue (doi: 10.5884/12518), alongside additional meteorological data provided by the Amundsen Science group of U. Laval. CTD station conductivity, temperature, and salinity data provided by the Amundsen Science group of U. Laval are available in the Polar Data Catalogue (doi: 10.5884/12713). Underway oxygen saturation and ΔO2/Ar data are available in the Polar Data Catalogue (doi: 10.5884/13242). All Fast Repetition Rate Fluorometry measurements of phytoplankton photophysiology are available in the Polar Data Catalogue (doi: 10.5884/13254). The Canadian Cryospheric Information Network reference number for this dataset is 13254, and will be trackable using this number once published. A doi will be provided as soon as available. River data collected by the Canadian GEOTRACES program is available from the PANGEA database (https://doi.org/10.1594/PANGAEA.908497).


    Articles from PLoS ONE are provided here courtesy of PLOS

    RESOURCES