The rate of hypoxia induction (RHI) is an important but overlooked dimension of environmental hypoxia that may affect an organism's survival. We hypothesized that, compared with rapid RHI, gradual RHI will afford an organism more time to alter plastic phenotypes associated with O2 uptake and subsequently reduce the critical O2 tension (Pcrit) of the rate of O2 uptake (O2). We investigated this by determining Pcrit values for goldfish exposed to short (∼24 min), typical (∼84 min) and long (∼480 min) duration Pcrit trials to represent different RHIs. Consistent with our predictions, long duration Pcrit trials yielded significantly lower Pcrit values (1.0–1.4 kPa) than short and typical duration trials, which did not differ (2.6±0.3 and 2.5±0.2 kPa, respectively). Parallel experiments revealed these time-related shifts in Pcrit were associated with changes to aspects of the O2 transport cascade that took place over the hypoxia exposures: gill surface areas and haemoglobin–O2 binding affinities were significantly higher in fish exposed to gradual RHIs over 480 min than fish exposed to rapid RHIs over 60 min. Our results also revealed that the choice of respirometric technique (i.e. closed versus intermittent) does not affect Pcrit or routine O2, despite the significantly reduced water pH and elevated CO2 and ammonia levels measured following closed-circuit Pcrit trials of ∼90 min. Together, our results demonstrate that gradual RHIs result in alterations to physiological parameters that enhance O2 uptake in hypoxic environments. An organism's innate Pcrit is therefore most accurately determined using rapid RHIs (<90 min) so as to avoid the confounding effects of hypoxic acclimation.

Environmental hypoxia is a common characteristic of many aquatic systems and is becoming increasingly prevalent, severe and long-lasting because of anthropogenic and climate change effects (Friedrich et al., 2014; IPCC, 2014; Smith et al., 2006). Many studies have examined the physiological impacts of hypoxia exposure on a diverse array of fish species, but these have focused almost exclusively on either the severity of the hypoxic exposure [i.e. partial pressure of O2 in water (PwO2)] or its duration. However, a third dimension of hypoxic exposure, the rate of hypoxia induction (RHI), has received very little attention and is rarely even controlled for (or at least reported) when environmental hypoxia is experimentally induced (Rogers et al., 2016). This is unlike other abiotic variables such as temperature, which are typically altered at consistent rates across studies [e.g. 0.2–0.3°C min−1 for the determination of critical thermal maxima (CTmax)] owing to the effects they have on organismal responses (e.g. temperature tolerance in fishes; Mora and Maya, 2006). Similarly, RHIs may influence the physiological responses of fishes to hypoxia, particularly, time-dependent responses related to environmental O2 extraction.

Most fishes possess mechanisms that enhance O2 extraction and delivery to tissues as PwO2 is reduced, such as increased haemoglobin (Hb) synthesis (Gracey et al., 2001) and concentration in the blood (Affonso et al., 2002), increased haematocrit (Lai et al., 2006; Turko et al., 2014), increased Hb–O2 binding affinity (Turko et al., 2014), increased ventilation frequency and amplitude (Holeton and Randall, 1967; Itazawa and Takeda, 1978; Tzaneva et al., 2011; Vulesevic and Perry, 2006), and a redistribution of blood supply to critical tissues (Sundin et al., 1995). Some fishes, including goldfish and numerous other species, also have the ability to dramatically increase lamellar surface area in response to hypoxia exposure through apoptotic reductions to the inter-lamellar cell mass (ILCM; Anttila et al., 2015; Borowiec et al., 2015; Crispo and Chapman, 2010; Dhillon et al., 2013; Ong et al., 2007; Sollid et al., 2003, 2005; Turko et al., 2012). While these modifications to different parts of the O2 transport cascade function to improve O2 uptake at low PwO2, the time courses over which these modifications are enacted differ and may potentially impact the critical PO2 (Pcrit) of the rate of O2 uptake (O2).

Pcrit is defined as the PwO2 at which a fish's O2 transitions from being regulated at some stable level independent of PwO2 (i.e. oxyregulation) to being dependent upon PwO2 (i.e. oxyconformation). At Pcrit, the fish's aerobic scope is theoretically zero and at PwO2 values below Pcrit, the fish's ability to generate ATP aerobically is limited (Farrell and Richards, 2009). Pcrit therefore reflects a fish's ability to acquire and use environmental O2 as a function of PwO2, with a lower Pcrit indicating a greater ability to extract O2 to maintain aerobic metabolism in hypoxic environments. A low Pcrit is beneficial because it allows the animal to maintain a routine level of function and activity in hypoxic environments while avoiding a reliance on anaerobic glycolysis and/or metabolic rate depression. Indeed, we have recently shown that goldfish prioritize their use of aerobic metabolism in hypoxic environments over their exceptional ability to induce metabolic rate depression, which they reserve for anoxic environments (Regan et al., 2017). Goldfish also appear to enhance their ability to extract environmental O2 over relatively short time periods in hypoxia, which in theory should result in a lowering of their Pcrit value (Regan et al., 2017). Because this ability is influenced by a suite of O2 extraction mechanisms that are both plastic and time-dependent, we hypothesized that gradual RHIs would allow fish to induce plastic mechanisms that enhance O2 extraction, resulting in lower Pcrit values than those of fish exposed to rapid RHIs.

List of symbols and abbreviations
     
  • CO2

    oxygen content

  •  
  • CTmax

    critical thermal maxima

  •  
  • Hb

    haemoglobin

  •  
  • ILCM

    inter-lamellar cell mass

  •  
  • O2

    oxygen consumption rate

  •  
  • NTP

    organic phosphates (ATP and GTP)

  •  
  • OEC

    oxygen equilibrium curve

  •  
  • P50

    PO2 at which Hb is 50% saturated with oxygen

  •  
  • PCO2

    partial pressure of carbon dioxide

  •  
  • Pcrit

    critical partial pressure of oxygen for O2

  •  
  • PO2

    partial pressure of oxygen

  •  
  • PwCO2

    partial pressure of carbon dioxide in water

  •  
  • PwO2

    partial pressure of oxygen in water

  •  
  • RBC

    red blood cell

  •  
  • RHI

    rate of hypoxia induction

We tested this hypothesis by determining the Pcrit values of goldfish exposed to progressive reductions in PwO2 (from normoxia to near-anoxia; referred to as Pcrit trials) over different durations: ∼24 min to represent rapid RHI, ∼84 min to represent typical RHI (Pcrit trials in the literature typically last 60 to 120 min; see Rogers et al., 2016) and ∼480 min to represent gradual RHI. We also ran parallel hypoxic exposures of different RHIs to investigate morphological and physiological traits of goldfish that might play causal roles in a time-related shift in Pcrit, including (among other traits) gill morphometry and Hb–O2 binding affinity. Furthermore, our use of different respirometric techniques allowed us to disentangle the effects of time and technique on the determination of Pcrit, thereby addressing a longstanding concern over the use of closed-chamber respirometry and its associated metabolic end-product accumulation for the determination of Pcrit (Keys, 1930; Rogers et al., 2016; Snyder et al., 2016; Steffensen, 1989). And finally, we chose goldfish as our study species because they have well-characterized responses to hypoxia exposure (Dhillon et al., 2013; Mitrovic et al., 2009), including a well-resolved Pcrit as determined by closed-chamber respirometry (Dhillon et al., 2013; Fry and Hart, 1948; Fu et al., 2011; Regan et al., 2017), which could aid our analysis of how RHI might influence the underlying physiology of Pcrit.

Animals

Goldfish (Carassius auratus auratus Linnaeus 1758; 2.87±0.14 g wet mass; N=84; sex unknown) were purchased from a commercial supplier (The Little Fish Company, Surrey, BC, Canada) and held under a 12 h:12 h light:dark cycle in 100-litre recirculating systems of well-aerated, dechlorinated, 17°C water (replaced weekly) at the University of British Columbia (UBC). Stocking density was <0.3 g l−1. Fish were fed to satiation daily (Nutrafin Max Goldfish Flakes) except for 24 h before transfer to the experimental apparatus, when feeding ceased. UBC's Animal Care Committee approved all procedures (protocol A13-0309).

Respirometry

We exposed goldfish to Pcrit trials of short (∼24 min), typical (∼84 min, representing a typical closed-chamber Pcrit trial's duration) and long (∼480 min) durations to represent progressively reduced RHIs. These different RHIs were achieved using different respirometric techniques (details below), while the respirometer chambers, animal transfer protocol, habituation period and mean fish mass remained consistent across all trials. Each fish was used only once.

We used two 32 ml flow-through respirometer chambers made from stainless steel as described in Regan et al. (2013). For each trial, we inserted a fish into the chamber and held it under flow-through conditions for ≥16 h prior to commencing the Pcrit trial. The fish chamber was supplied with flow-through water at a rate of 190 ml h−1 and maintained at 17°C. Inflowing water was drawn from a well-mixed reservoir held at ∼26 kPa (manually controlled using compressed N2 and O2) and pumped to the respirometer chamber via a peristaltic pump (Gilson Minipuls 3, Middleton, WI, USA) through a combination of stainless steel tubing and gas-impermeable Tygon peristaltic tubing. The PwO2 of the inflowing water was maintained slightly hyperoxic to ensure that the outflowing water was always at or slightly above normoxic PwO2. Following the habituation period, we conducted our respirometry experiments.

For the typical duration Pcrit trials (84±8 min), we used closed-circuit respirometry. To start the trial, the inflowing and outflowing water supply lines were short-circuited to create a closed loop, with water recirculating through the chamber by the peristaltic pump at the same rate (190 ml h−1) as during the habituation period to ensure minimal disturbance to the fish and good mixing of the chamber's water volume. Chamber PwO2 was then allowed to decrease as a result of the fish's respiration. An O2 optode placed within the chamber (see Regan et al., 2013) continuously measured PwO2, and O2 was calculated according to:
formula
(1)

where CO2 is the O2 content of the water converted to μmol l−1 from PwO2 using the solubility factor of 14.485 μmol l−1 kPa−1 (Boutilier et al., 1984), T is the time period over which the change in CO2 is calculated (5 or 2 min; see below), V is the fish chamber volume (32 ml) plus the volume of the closed-circuit water lines minus the volume displaced by the fish itself, and M is the mass of the fish. The trials were ended when PwO2 reached 0 kPa, at which point the short-circuit was dismantled and flow-through conditions were reestablished to return chamber PwO2 to habituation period conditions.

For the short duration Pcrit trials (24±2 min), we again used closed-circuit respirometry as described for the typical Pcrit trials. To shorten the trial and hasten the PwO2 decline, we made initial normoxic O2 readings and then manually replaced the entire water volume of the respirometry chamber and its water supply lines over ∼5 min with water equilibrated to 5.3 kPa PwO2 using a 60 ml syringe. PwO2 was therefore reduced from normoxia to ∼5.3 kPa not by the fish's O2, but by the active replacement of the water volume. At this point, we attached the water supply lines to the peristaltic tubing, turned the pump back on to 190 ml h−1, and allowed the fish to deplete the closed system's O2 through its own respiration (typically over a ∼20 min period). We chose 5.3 kPa as our replacement PwO2 for two reasons: first, it allowed for reliable O2 measurements starting at ∼4.8 kPa, which provided enough O2 data points above Pcrit to construct robust Pcrit traces; and second, the amount of time required for the fish to reduce PwO2 from 5.3 kPa to anoxia put the overall duration of these Pcrit trials within our targeted duration of between 20 and 30 min. Although these procedures reduced the overall duration of the Pcrit trial, we must point out that the RHI below 5.3 kPa was similar to that of the typical duration trials. If mechanisms of enhanced O2 extraction are only induced at PwO2 <5.3 kPa, then these two techniques could result in similar Pcrit values.

Prior to actively replacing the water volume, we converted the system to closed-circuit and made a series of normoxic O2 readings between 25 and 19 kPa to aid in our calculation of Pcrit (see below). Upon reaching 19 kPa, we converted the system back to flow-through, reestablished a normoxic PwO2 of ∼21 kPa, and then commenced the active water volume replacement.

For the long duration Pcrit trials, we used three different respirometric techniques to ensure the mean Pcrit values were the result of Pcrit trial duration and not respirometric technique per se. These trials varied in average duration from 434 to 562 min depending on the technique used. We chose a time duration of ∼480 min because it was significantly longer than the typical trial durations, but likely shorter than would be required to induce gene expression acclimation responses. It is also in line with some of the longer Pcrit trial durations observed in the literature (see Rogers et al., 2016).

For our first technique, we used closed-circuit respirometry where we added a 250 ml water reservoir to reduce the rate at which the fish's respiration depleted the system's O2. This reservoir was a glass bottle placed immediately after the peristaltic pump. Water leaving the respirometer chamber was pumped into the reservoir directly over a stir bar that mixed the water volume to prevent O2 stratification in the bottle. Water flowed out of the reservoir through a stainless steel line that punctured the bottle's rubber stopper and went directly into the stainless steel line supplying the respirometer chamber. All materials used were gas-impermeable glass or stainless steel. Attaching this reservoir to the closed-circuit system took ∼2 min, after which the peristaltic pump was turned back on and the Pcrit trial was run according the closed-circuit technique described for the typical duration Pcrit trials. The average duration for these closed-circuit trials was 434±56 min.

Second, we used flow-through respirometry where O2 was calculated according to:
formula
(2)

where CiO2 and CoO2 are O2 content of inflowing and outflowing water, respectively, converted from PwO2 as described above (we used a single PO2 optode for these measurements) and f is water flow rate (190 ml h−1). We held fish at approximately 26, 16, 5.3, 2.7, 1.3, 0.7 and 0 kPa, each PwO2 in series, in that order and for 1 h, and at each PwO2 we measured O2 at 10, 30 and 60 min (10 min was the minimum time required to ensure PwO2 had equilibrated across the respirometer and the upstream and downstream PwO2 measurement chambers). Because the calculated O2 at each PwO2 was nearly identical at each of the three time points, we averaged across the time points and calculated Pcrit from those averaged O2 values for each individual. The average duration for these flow-through trials was 562±19 min, including the time required to reach target PwO2 values.

Third, we used a variation on intermittent flow respirometry that combined flow-through and closed-circuit respirometry. We used flow-through conditions to manually reduce PwO2 from normoxia to ∼2.8 kPa over ∼430 min and then commenced a period of closed-circuit respirometry, which took an additional ∼15 min. We chose a target PwO2 of 2.8 kPa to start the closed-circuit portion of the trial based upon our earlier short-term Pcrit trials (which used the same combined respirometric technique) that suggested we could reliably determine Pcrit from this PwO2. Upon reaching 2.8 kPa, we converted to the closed-circuit setup and allowed the fish's respiration to deplete the remaining O2 in the closed system as described previously. This combination of techniques allowed for closed-circuit O2 measurements with a reduced accumulation of metabolic end-products. As with the rapid RHI Pcrit trials, we used closed-circuit respirometry to make a series of normoxic O2 readings between 25 and 19 kPa prior to the active (but in this case gradual) reduction of PwO2 to aid in our calculation of Pcrit (see below). Upon reaching 19 kPa, we converted the system back to flow-through, reestablished a PwO2 of ∼21 kPa, and then commenced the active water volume replacement. The average duration of these combined flow-through and closed-circuit trials was 444±12 min.

Parallel hypoxic exposures for physiological measurements

We ran two separate but identical parallel sets of hypoxic exposures to investigate potentially causal physiological factors in a time-dependent reduction in Pcrit. These parallel exposures involved manually reducing PwO2 of aquaria from normoxia to anoxia over 60 and 480 min periods to represent rapid and gradual RHIs, respectively. We also ran normoxic control exposures during which PwO2 remained normoxic for 480 min following the habituation period. Each exposure was run in two 10 litre aquaria housing four fish each, and each aquarium was fitted with a screen just below the water surface to prevent the fish from accessing the air–water interface. We mimicked the respirometric Pcrit trials described above as closely as possible, with exposures run at 17°C at the same time of day (each trial commenced at ∼09:00 h) following a ≥16 h habituation period, and in complete darkness. Fish from the first set of parallel exposures were sampled to assess gill morphology and haematological parameters, and fish from the second set of parallel exposures were sampled to measure plasma [lactate].

At the end of each exposure, fish were euthanized by inconspicuously introducing anaesthetic (buffered MS-222, final concentration of 200 mg l−1) to the water. Once fish ceased to respond to a tail pinch, individuals were removed, weighed, and then blood was sampled and gills were dissected. To sample blood from the fish in the first set of parallel exposures, the fish's tail was severed and blood was collected from the caudal preduncle using a 60 μl heparinized capillary tube. Ten microlitres of blood was pipetted into 1 ml Drabkins reagent for determination of [Hb], 20 μl of blood was mixed with 10 μl of heparinized Cortland's saline plus 80 μl of 3% perchloric acid for determination of the concentration of red blood cell (RBC) organic phosphates (ATP and GTP; [NTP]), and 10 μl of blood was mixed with 5 μl of heparinzed Cortland's saline for determination of Hb–O2 binding affinity. The entire right gill basket was then removed from the fish and immediately immersed in 1 ml of Karnovsky's fixative (25% glutaraldehyde, 16% formaldehyde, 0.15 mol l−1 sodium cacodylate, pH 7.4). Twenty-four hours later, the gill basket was transferred to 0.15 mol l−1 sodium cacodylate and stored at 4°C until use. This procedure was repeated for all four fish in each tank, and then duplicated for the second tank of four fish, yielding N=8 for each treatment. For the second set of parallel exposures, fish were euthanized and blood was collected in the same manner as before, but the plasma was separated from the red blood cells by centrifugation and immediately assayed for plasma [lactate] (see below).

The goal of these parallel exposures was to assess the effects of RHI on physiological adjustments that may explain differences in Pcrit, but there are differences between the Pcrit trials and the parallel exposures that the reader should be made aware of. The main difference was vessel size (respirometer chamber was 32 ml and exposure aquaria were 10 litres), which could have affected the ability of the fish to move throughout the exposure. However, observations of the fish in the 10-litre aquaria suggest that goldfish do not increase activity during progressive hypoxia exposure. Furthermore, the parallel exposures were terminated when PwO2 reached 0 kPa. As the samples were taken at this point, the haematology and gill morphology measurements were not taken precisely at the point at which we observed differences in Pcrit, and this could affect our ability to relate the two studies. However, the fish used for the haematology and gill morphology analyses were only exposed to an additional ∼7 to ∼15 min of progressively deepening hypoxia (for rapid and gradual RHI, respectively) beyond what they were exposed to by the time Pcrit had been reached. Thus we do not believe these relatively minor differences in time would affect our ability to directly relate these components of our study.

Gill morphometrics

Gill samples were randomly assigned an alphanumeric code by an independent party so analysis could be performed blindly. The second gill arch of each gill basket was isolated and its anterior hemibranch imaged using light microscopy (Olympus Stereomicroscope SZX10; 6.3× magnification, 10× zoom; image capture using cellSens Software). The images were used in combination with ImageJ v2.0.0 software to measure filament length and number, and lamellar height (distance from base to the distal edge of the lamellae), length (distance lamellae runs along the filament) and frequency (number of lamellae per unit distance of filament). We made the lamellar measurements by dividing the length of the gill arch into five sections, and then isolating a filament from each of these sections. Each filament was imaged from the top and the side, providing clear views of the height and length of its lamellae that we later measured. Specifically, we measured the height, length and width of three lamellae per filament (one from the filament's base, one from its middle and one from its tip), as well as the width of inter-lamellar channels in these three regions. We then estimated each filament's lamellar frequency (lamellae μm−1) by dividing filament length by the sum of that filament's average channel and lamellar widths. Total lamellar surface area (SAL,total) for each fish was then calculated according to:
formula
(3)

where SAF is the mean lamellar surface area of the five analyzed filaments, F is the number of filaments per gill arch, and 16 is the product of two hemibranchs per gill arch, four gill arches per gill basket and two gill baskets per individual fish (according to Wegner, 2011).

Blood analyses

Hb–O2 binding affinity was determined within 60 min of blood sampling by constructing an oxygen equilibrium curve (OEC) using the thin film spectrophotometric technique (Lilly et al., 2013) and a 96-well microplate spectrophotometer (Molecular Devices, Sunnyvale, CA, USA). A Wostoff gas mixing pump (H. Wösthoff Messtechnik GmbH, Bochum, Germany) mixed compressed O2 and N2 to each of nine PO2 values between 0 and 21 kPa PO2, always starting with 0 kPa and working toward 21 kPa, and each PO2 was maintained for 20 min, during which Hb–O2 saturation was determined spectrophotometrically. A sigmoidal OEC was fit through the % Hb–O2 saturation versus PO2 data for each fish, and Hb P50 (the PO2 at which Hb is 50% saturated with O2) was determined using SigmaStat 11.0.

We measured whole blood [Hb] spectrophotometrically at 17°C and 540 nm after conversion to cyanomethemoglobin using Drabkin's reagent (Sigma-Aldrich). The measurements were made using a Shimadzu UV-160 spectrophotometer and a millimolar extinction coefficient of 11.

We measured RBC [NTP] spectrophotometrically at 17°C using the GAPDH- and PGK-catalyzed reactions converting glycerate 3-phosphate to glyceraldehyde 3-phosphate, where the oxidation of NADH to NAD+ was measured at 340 nm (Bergmeyer et al., 1983). Finally, we measured plasma [lactate] spectrophotometrically at 17°C using the LDH-catalyzed reaction converting lactate to pyruvate, where the reduction of NAD+ to NADH was measured at 340 nm (Bergmeyer et al., 1983). [NTP] and [lactate] were measured using a 96-well microplate spectrophotometer (Molecular Devices).

CO2 and nitrogenous end-product measurements

The accumulation of metabolic end-products during closed-chamber/circuit respirometry is regarded as a major shortcoming of the technique (e.g. Keys, 1930; Rogers et al., 2016; Snyder et al., 2016; Steffensen, 1989), but measurements of metabolic end-products in the respirometer chamber are rarely, if ever, reported. To address this knowledge gap, we ran a separate set of closed-circuit Pcrit trials (91±10 min) to measure accumulated levels of CO2 and nitrogenous end-products (NH3+NH4+). For each of four fish, we took water samples from the respirometer chamber at three time points: start of the habituation period; end of a 16 h habituation period immediately prior to starting the Pcrit trial; and end of the Pcrit trial as soon as the respirometer's PwO2 reached 0 kPa. PwCO2 was determined using the Henderson–Hasselbalch equation and measurements of total CO2 content in the water (CO2+HCO3; Corning 965 Carbon Dioxide Analyzer, Corning, NY, USA) and pH (probe: SaS gK2401C, Radiometer Analytical, France; meter: VWR Symphony SB70P, VWR, Radnor, PA, USA). Total ammonia (NH3+NH4+) was measured using an API ammonia test kit, and unionized ammonia (NH3) was calculated from this value in combination with the particular trial's water pH and temperature (17°C).

Pcrit calculation

Pcrit is defined as the PwO2 at which an organism's stable O2 transitions from being independent of to being dependent upon PwO2. There are different methods to calculate Pcrit, but analyses performed by Mueller and Seymour (2011) suggest that most of the methods used yield comparable values. We therefore decided to use a variation on a two-segment linear regression model (details below) to identify Pcrit as the PwO2 at which the two linear trend lines (one representing the PwO2 range of oxyregulation, the other of oxyconformation) intersect on a graph plotting O2 as a function of PwO2 (BASIC program of Yeager and Ultsch, 1989). This method is employed widely throughout the literature (see Rogers et al., 2016) and has been used on goldfish (Fu et al., 2011; Dhillon et al., 2013; Regan et al., 2017).

We calculated O2 values by measuring the change in PwO2 over sequential time intervals: 5 min between 25 and 5.3 kPa and 2 min between 5.3 and 0 kPa. To standardize our estimates of a stable, oxyregulated O2, we used the mean of each fish's calculated O2 values between 21 and 18.7 kPa PwO2. This represented a normoxic routine O2 that was likely close to standard O2 as a result of it being made following a habituation period that was ≥16 h. We then determined Pcrit as the intersection of this horizontal line with a linear regression through the O2 values that were >15% below the mean routine O2 value. This technique was carried out according to McBryan et al. (2016).

Data analysis and statistics

We compared all average values of Pcrit, normoxic O2, blood properties, gill morphometrics and accumulated PwCO2 and nitrogenous end-products using one-way, two-tailed ANOVAs with a critical α=0.05 (repeated measures for the water pH, PwCO2 and nitrogenous end-products comparisons). Post hoc Tukey tests were used to test for differences between treatment groups. Any data set that did not meet the assumptions of normality or equal variance were log-transformed prior to analysis. All analyses were performed using SigmaStat 11.0. Values reported in the text are presented as means±s.e.m.

Respirometry

Long duration Pcrit trials resulted in Pcrit values that were approximately half those of short and typical duration Pcrit trials (ANOVA, P<0.001; Fig. 1A, Fig. S1). Pcrit values determined by short and typical trial durations did not differ from one another, nor did the Pcrit values determined by the three respirometric techniques used for the long duration trials (Fig. 1A). Each of the five respirometric techniques yielded statistically similar normoxic O2 values (ANOVA, P=0.276; Fig. 1B).

Fig. 1.

The effect of Pcrit trial duration on the average Pcrit and normoxic O2 values of goldfish. (A) Average Pcrit values of the individual fish comprising each set of respirometry experiments. (B) Average normoxic O2 values of the individual fish comprising each set of respirometry experiments while those fish were exposed to normoxic PwO2 (18 to 26 kPa). Details on the respirometric techniques are included in the Materials and methods, but briefly, ‘short’ used combined flow-through/closed-circuit intermittent flow respirometry (N=5; 24±2.2 min), ‘typical’ used closed-circuit respirometry (N=6; 84±8 min) and ‘long’ from left to right used closed-circuit respirometry (N=6; 434±56 min), combined flow-through/closed-circuit intermittent flow respirometry (N=4; 444±12 min) and flow-through respirometry (N=6; 562±19 min). Error bars are s.e.m.; bars that share a letter are not significantly different (one-way ANOVA, aP<0.001, bP=0.276).

Fig. 1.

The effect of Pcrit trial duration on the average Pcrit and normoxic O2 values of goldfish. (A) Average Pcrit values of the individual fish comprising each set of respirometry experiments. (B) Average normoxic O2 values of the individual fish comprising each set of respirometry experiments while those fish were exposed to normoxic PwO2 (18 to 26 kPa). Details on the respirometric techniques are included in the Materials and methods, but briefly, ‘short’ used combined flow-through/closed-circuit intermittent flow respirometry (N=5; 24±2.2 min), ‘typical’ used closed-circuit respirometry (N=6; 84±8 min) and ‘long’ from left to right used closed-circuit respirometry (N=6; 434±56 min), combined flow-through/closed-circuit intermittent flow respirometry (N=4; 444±12 min) and flow-through respirometry (N=6; 562±19 min). Error bars are s.e.m.; bars that share a letter are not significantly different (one-way ANOVA, aP<0.001, bP=0.276).

Effect of RHI on gill morphology

RHI significantly affected the mass-specific lamellar surface areas of goldfish (ANOVA, P=0.004; Fig. 2), whereby fish exposed to gradual RHIs had ∼60% larger lamellar surface areas than fish exposed to rapid RHIs and normoxic controls, which did not differ.

Fig. 2.

The effect of the rate of hypoxia induction (RHI) on the mass-specific lamellar surface area of goldfish. Mass-specific lamellar surface areas of goldfish exposed to rapid and gradual RHIs (normoxia to anoxia in 60 and 480 min, respectively) and normoxic controls (N=8 for each; one-way ANOVA, P=0.004). Error bars are s.e.m.; bars that share a letter are not significantly different.

Fig. 2.

The effect of the rate of hypoxia induction (RHI) on the mass-specific lamellar surface area of goldfish. Mass-specific lamellar surface areas of goldfish exposed to rapid and gradual RHIs (normoxia to anoxia in 60 and 480 min, respectively) and normoxic controls (N=8 for each; one-way ANOVA, P=0.004). Error bars are s.e.m.; bars that share a letter are not significantly different.

Effect of RHI on Hb–O2 binding affinity, [Hb] and RBC [NTP]

RHI significantly affected Hb–O2 binding affinity (ANOVA, P=0.007; Fig. 3). Goldfish exposed to rapid RHIs had ∼60% higher Hb P50 values than goldfish exposed to gradual RHIs and normoxic controls, which did not differ.

Fig. 3.

The effect of RHI on the oxygen equilibrium curve (OEC) and the Hb P50 value of goldfish. (A) OECs for the extracted whole blood of goldfish exposed to rapid and gradual RHIs (normoxia to anoxia in 60 and 480 min, respectively) and normoxic controls. The blood was collected from the fish immediately upon PwO2 reaching ∼0 kPa (for rapid and gradual RHIs), and the spectrophotometric determination of Hb–O2 saturation was begun as soon thereafter as possible (<1 h). Each OEC is a trendline through the data points of eight blood samples exposed to nine PwO2 values between 0 and 21 kPa PO2. Grey horizontal dashed line highlights the 50% Hb–O2 saturation point. (B) Average Hb P50 values (PO2 at which Hb is 50% saturated with O2) for each treatment group (one-way ANOVA, P=0.007). Error bars are s.e.m.; bars that share a letter are not significantly different.

Fig. 3.

The effect of RHI on the oxygen equilibrium curve (OEC) and the Hb P50 value of goldfish. (A) OECs for the extracted whole blood of goldfish exposed to rapid and gradual RHIs (normoxia to anoxia in 60 and 480 min, respectively) and normoxic controls. The blood was collected from the fish immediately upon PwO2 reaching ∼0 kPa (for rapid and gradual RHIs), and the spectrophotometric determination of Hb–O2 saturation was begun as soon thereafter as possible (<1 h). Each OEC is a trendline through the data points of eight blood samples exposed to nine PwO2 values between 0 and 21 kPa PO2. Grey horizontal dashed line highlights the 50% Hb–O2 saturation point. (B) Average Hb P50 values (PO2 at which Hb is 50% saturated with O2) for each treatment group (one-way ANOVA, P=0.007). Error bars are s.e.m.; bars that share a letter are not significantly different.

RHI did not affect whole-blood [Hb] (ANOVA, P=0.334; Fig. 4A), but it did affect RBC [NTP] (ANOVA, P=0.001; Fig. 4B), whereby gradual RHI fish had RBC [NTP] values that were approximately half those of the rapid RHI and normoxic control fish, which did not differ.

Fig. 4.

The effect of RHI on blood parameters of goldfish. (A) Average values for whole blood [Hb] measured spectrophotometrically using Drabkins reagent (N=6–8; one-way ANOVA, P=0.334). (B) Average values for red blood cell [NTP] (N=6–8; one-way ANOVA, P=0.001). (C) Average values for plasma [lactate] (N=6–8; one-way ANOVA, P=0.001). Error bars are s.e.m.; bars that share a letter are not significantly different.

Fig. 4.

The effect of RHI on blood parameters of goldfish. (A) Average values for whole blood [Hb] measured spectrophotometrically using Drabkins reagent (N=6–8; one-way ANOVA, P=0.334). (B) Average values for red blood cell [NTP] (N=6–8; one-way ANOVA, P=0.001). (C) Average values for plasma [lactate] (N=6–8; one-way ANOVA, P=0.001). Error bars are s.e.m.; bars that share a letter are not significantly different.

Effect of RHI on plasma lactate

Goldfish exposed to rapid and gradual RHIs both accumulated similar concentrations of plasma lactate to a level significantly higher than that observed in normoxic control fish (ANOVA, P=0.001; Fig. 4C).

Metabolic end-product accumulation

Respirometer chamber PwCO2 doubled over the course of a 16 h habituation period under flow-through conditions, and increased 6.5-fold over the course of a typical duration closed-circuit Pcrit trial (91±10 min; ANOVA, P<0.001; Fig. 5A). Water pH was concomitantly reduced from 7.61 to 6.93 over the course of the Pcrit trial (ANOVA, P<0.001; Fig. 5B). The concentration of total ammonia (NH3+NH4+) in the chamber also increased (ANOVA, P<0.001; Fig. 5C). Unionized ammonia (NH3) accumulated in a different way because of pH changes of the water, with [NH3] tripling over the 16 h habituation period, then falling to an intermediate value by the end of the Pcrit trial (ANOVA, P<0.001; Fig. 5D).

Fig. 5.

The effects of closed-circuit respirometry on water chemistry and the buildup of metabolic end-products. (A) Chamber PwCO2, measured as total CO2 and converted to PwCO2 using water pH and the Henderson–Hasselbalch equation (N=4; one-way ANOVA, P<0.001). (B) Water pH (N=4; one-way ANOVA, P<0.001). (C) Total ammonia concentration (NH3+NH4+; N=4; one-way ANOVA, P<0.001). (D) Unionized ammonia concentration, calculated using water pH and temperature (N=4; one-way ANOVA, P<0.001). Error bars are s.e.m.; bars that share a letter are not significantly different.

Fig. 5.

The effects of closed-circuit respirometry on water chemistry and the buildup of metabolic end-products. (A) Chamber PwCO2, measured as total CO2 and converted to PwCO2 using water pH and the Henderson–Hasselbalch equation (N=4; one-way ANOVA, P<0.001). (B) Water pH (N=4; one-way ANOVA, P<0.001). (C) Total ammonia concentration (NH3+NH4+; N=4; one-way ANOVA, P<0.001). (D) Unionized ammonia concentration, calculated using water pH and temperature (N=4; one-way ANOVA, P<0.001). Error bars are s.e.m.; bars that share a letter are not significantly different.

We hypothesized that gradual RHIs would allow goldfish to induce time-dependent plastic phenotypes that enhance O2 uptake. This hypothesis predicted that the Pcrit of goldfish exposed to long duration Pcrit trials would be lower than those of goldfish exposed to short or typical duration Pcrit trials, and our results agree with these predictions regardless of the respirometric technique used. Furthermore, fish exposed to gradual RHIs developed greater lamellar surface areas and Hb–O2 affinities over their exposures than fish exposed to rapid RHIs, both characteristics that would enhance O2 extraction under hypoxia and therefore likely explain the lower Pcrit associated with the long duration trials. Taken together, our results suggest that time (more precisely, RHI) is a significant determinant of Pcrit in goldfish.

The majority of Pcrit measurements are made using closed-chamber respirometry over the course of 60 to 120 min (Rogers et al., 2016). Here, our representative closed-circuit Pcrit trials lasted ∼84 min and resulted in a Pcrit of 2.5±0.2 kPa (Fig. 1A). Our values are in general agreement with values previously reported for goldfish [∼3.6 kPa (Fry and Hart, 1948); 3.0 kPa (Fu et al., 2011); 3.3 kPa (Dhillon et al., 2013); and 3.0 kPa (Regan et al., 2017)], though slightly lower owing to a possible combination of experimental temperature differences and the fact that our study used closed-circuit respirometry as opposed to static closed-chamber respirometry. Reducing the trial duration to ∼24 min did not affect Pcrit (Fig. 1A), which may not be surprising considering the RHI below a PwO2 of 5.3 kPa was similar between the short and typical duration Pcrit trials (see Materials and methods for details). However, our results clearly indicate that increasing the trial duration from ∼84 min (i.e. reducing its RHI) to ∼480 min resulted in significantly lower Pcrit values. The reasons for this variation could be related to time, technique or some combination of the two, and we will explore these possibilities below.

Effects of time on the physiology of O2 uptake

Goldfish exposed to gradual RHIs developed significantly larger lamellar surface areas than those of normoxic controls and goldfish exposed to rapid RHIs, which did not differ. Hypoxia-induced gill remodeling was first observed in goldfish and the closely related crucian carp (Carassius carassius) 13 years ago (Sollid et al., 2003, 2005) and in numerous fish species since [e.g. mangrove killifish, Kryptolebias marmoratus (Ong et al., 2007; Turko et al., 2012); African cichlids (Crispo and Chapman, 2010); various carp species (Dhillon et al., 2013); Atlantic salmon Salmo salar (Anttila et al., 2015); and Atlantic killifish Fundulus heteroclitus (Borowiec et al., 2015)]. Dhillon et al. (2013) observed a near-doubling in the lamellar surface area of goldfish following 8 h acclimation to a constant PwO2 of 0.7 kPa, but to our knowledge, the present study is the first time gills have been shown to remodel over such short time scales under progressively decreasing PwO2. Increases to lamellar surface area are typically the result of apoptotic reductions to the ILCM (Sollid et al., 2003). ILCM reductions per se also enhance the gill's diffusion capacity (Bindon et al., 1994; Greco et al., 1995) and contribute to a reduced Pcrit in crucian carp (Sollid et al., 2003) and Atlantic killifish (Borowiec et al., 2015). However, a study that examined (among other things) O2 diffusion across the gills of two groups of goldfish with temperature-induced differences in gill surface area found that the differences in gill surface area had no effect on arterial PO2 when acutely exposed to hypoxia (Tzaneva et al., 2011). Although this seems to run counter to what Fick's first diffusion law would predict, the authors speculated that the goldfish that started hypoxia exposure with a smaller gill surface area may have been rapidly remodeling their gills to increase lamellar surface area over the course of the acute hypoxia exposure. Our gill morphometric results lend empirical support to this speculation.

The mean Hb P50 value of gradual RHI goldfish (1.6 kPa, similar to long duration Pcrit of ∼1.2 kPa) was significantly lower than that of rapid RHI goldfish (2.4 kPa, similar to typical duration Pcrit of 2.5 kPa), but no different than normoxic controls (1.4 kPa). This implies that rapid hypoxia induction reduces Hb–O2 binding affinity, but gradual induction does not. The underlying mechanism(s) might involve RBC [NTP] and/or pH. Nucleoside triphosphates (ATP and GTP, collectively NTPs) reduce Hb–O2 binding affinity by binding to sites on the Hb tetramer that stabilize its deoxygenated conformation and consequently increase the P50 (Jensen et al., 1998; Wood and Johansen, 1972). The significantly lower [NTP] values of our gradual RHI fish (Fig. 4B) at least partly explain their lower Hb P50 values compared with those of the rapid RHI fish (Fig. 3), but the similar [NTP] values in the rapid RHI and normoxic control fish exclude this mechanism as the cause of the rapid RHI fish's elevated Hb P50 values. Another possible mechanism for the rapid RHI fish's reduced Hb–O2 binding affinity is RBC pH. We did not measure RBC pH, but if protons accumulated to higher concentrations in the RBCs of rapid RHI goldfish than gradual RHI goldfish, then this would likely have reduced Hb–O2 binding affinity via goldfish's Bohr/Root effect (Rodewald and Braunitzer, 1984). Regardless of the causal mechanism(s), the different Hb–O2 binding affinities of the rapid and gradual RHI fish are likely to at least partly explain their different Pcrit values.

Respirometric technique and end-product accumulation

Respirometric techniques can be broadly categorized as closed (closed-circuit or static closed-chamber), flow-through or intermittent flow. Though none of these techniques are ideal for all experimental scenarios, intermittent flow respirometry is generally regarded as superior because it avoids the potential accumulation of metabolic end-products presumed to occur in closed respirometry and it has greater temporal resolution than flow-through respirometry (reviewed by Clark et al., 2013; Steffensen, 1989; Svendsen et al., 2016). It has been suggested that the choice of respirometric technique used to determine Pcrit may influence the results, and indeed Pcrit in shiner perch (Cymatogaster aggregata) shifted from ∼9.9 kPa to ∼6.1 kPa when using closed-chamber versus intermittent flow respirometry, respectively (Snyder et al., 2016). The authors attribute this to technique, but also discuss the possibility that duration of the Pcrit trials (∼1 h for closed-chamber, ∼5 h for intermittent flow) may play a role (Snyder et al., 2016). In the present study, we used modified versions of all three respirometric techniques for our long duration Pcrit trials, and, despite technique specific-differences and challenges (e.g. flow-through trials demanded a step-wise reduction in PwO2; closed-circuit trials resulted in higher O2 values in the mid-PwO2 range; Fig. S1), each technique yielded nearly identical Pcrit values, which were all lower than the typical or short duration Pcrit trials. This suggests that the different Pcrit values observed between our short and typical duration Pcrit trials and those of the long duration trials are the result of RHI rather than technique, and this may also be the case with the results of Snyder et al. (2016).

The fact remains that closed respirometry leads to end-product accumulation (Fig. 5), and this could theoretically influence Pcrit (Keys, 1930; Snyder et al., 2016; Steffensen, 1989). However, within our long duration Pcrit trials, closed-circuit trials (where CO2 accumulated) and combined flow-through/closed-circuit trials (where CO2 did not accumulate) resulted in nearly identical Pcrit values (Fig. 1A). This suggests that the levels of metabolic end-products that accumulate with closed-circuit respirometry are not high enough per se to have a significant effect on Pcrit.

Conclusions

Our results demonstrate that RHI significantly alters the Pcrit of goldfish, whereby long duration Pcrit trials (i.e. gradual RHIs) yield lower Pcrit values than short duration Pcrit trials (i.e. rapid RHIs). These reduced Pcrit values are caused by time-dependent effects on mechanisms that enhance environmental O2 extraction, including gill morphology and Hb–O2 binding affinity. Fishes generally possess numerous time-dependent mechanisms that enhance O2 extraction in response to hypoxia, so RHI is likely an important factor to consider in all fish species when carrying out Pcrit trials and experimental hypoxia exposures. Because longer duration Pcrit trials allow for some degree of acclimation that may consequently reduce Pcrit, shorter duration Pcrit trials are likely to best represent the innate abilities of a hypoxia-exposed fish to extract and use O2 at the time of analysis. Thus, similar to the standardized rate of temperature change used when determining a fish's CTmax, an RHI should be chosen that is fast enough to avoid acclimation during the trial. Our data suggest that Pcrit trials of <90 min are probably sufficient to achieve this.

We thank Tara McBryan for her assistance with gill morphometrics, and Phillip Morrison for his assistance with OEC measurement and construction. We also thank the two reviewers for their helpful, insightful comments.

Author contributions

Conceptualization: M.D.R.; Methodology: M.D.R., J.G.R.; Formal analysis: M.D.R.; Investigation: M.D.R.; Resources: M.D.R.; Writing - original draft: M.D.R.; Writing - review & editing: M.D.R., J.G.R.; Funding acquisition: J.G.R.

Funding

This work was supported by a Natural Sciences and Engineering Research Council of Canada (NSERC) Discovery Grant to J.G.R. M.D.R. was supported by an NSERC postgraduate scholarship and a University of British Columbia Zoology Graduate Fellowship.

Data availability

Data are available from the Dryad Digital Repository (Regan and Richards, 2017): http://dx.doi.org/10.5061/dryad.bp20p

Affonso
,
E. G.
,
Polez
,
V. P.
,
Correa
,
C. F.
,
Mazon
,
A. F.
,
Araujo
,
M.
,
Moraes
,
G.
and
Rantin
,
F. T.
(
2002
).
Blood parameters and metabolites in the teleost fish Colossoma macropomum exposed to sulfide or hypoxia
.
Comp. Biochem. Physiol. C
133
,
375
-
382
.
Anttila
,
K.
,
Lewis
,
M.
,
Prokkola
,
J. M.
,
Kanerva
,
M.
,
Seppänen
,
E.
,
Kolari
,
I.
and
Nikinmaa
,
M.
(
2015
).
Warm acclimation and oxygen depletion induce species-specific responses in salmonids
.
J. Exp. Biol.
218
,
1471
-
1477
.
Bergmeyer
,
H. U.
,
Bergmeyer
,
J.
and
Grassl
,
M.
(
1983
).
Methods of Enzymatic Analysis
.
Weinheim, FL
:
Verlag Chemie
.
Bindon
,
S.
,
Gilmour
,
K.
and
Fenwick
,
J.
(
1994
).
The effects of branchial chloride cell proliferation on respiratory function in the rainbow trout Oncorhynchus mykiss
.
J. Exp. Biol.
197
,
47
-
63
.
Borowiec
,
B. G.
,
Darcy
,
K. L.
,
Gillette
,
D. M.
and
Scott
,
G. R.
(
2015
).
Distinct physiological strategies are used to cope with constant hypoxia and intermittent hypoxia in killifish (Fundulus heteroclitus)
.
J. Exp. Biol.
218
,
1198
-
1211
.
Boutilier
,
R. G.
,
Heming
,
T. A.
and
Iwama
,
G. K.
(
1984
).
Appendix: physicochemical parameters for use in fish respiratory physiology
.
Fish Physiol.
10
,
403
-
430
.
Clark
,
T. D.
,
Sandblom
,
E.
and
Jutfelt
,
F.
(
2013
).
Aerobic scope measurements of fishes in an era of climate change: respirometry, relevance and recommendations
.
J. Exp. Biol.
216
,
2771
-
2782
.
Crispo
,
E.
and
Chapman
,
L. J.
(
2010
).
Geographic variation in phenotypic plasticity in response to dissolved oxygen in an African cichlid fish
.
J. Evol. Biol.
23
,
2091
-
2103
.
Dhillon
,
R. S.
,
Yao
,
L.
,
Matey
,
V.
,
Chen
,
B.-J.
,
Zhang
,
A.-J.
,
Cao
,
Z.-D.
,
Fu
,
S.-J.
,
Brauner
,
C. J.
,
Wang
,
Y. S.
and
Richards
,
J. G.
(
2013
).
Interspecific differences in hypoxia-induced gill remodeling in carp
.
Physiol. Biochem. Zool.
86
,
727
-
739
.
Farrell
,
A. P.
and
Richards
,
J. G.
(
2009
).
Chapter 11: Defining hypoxia
.
Fish Physiol.
27
,
487
-
503
.
Friedrich
,
J.
,
Janssen
,
F.
,
Aleynik
,
D.
,
Bange
,
H. W.
,
Boltacheva
,
N.
,
Çagatay
,
M. N.
,
Dale
,
A. W.
,
Etiope
,
G.
,
Erdem
,
Z.
,
Geraga
,
M.
, et al. 
(
2014
).
Investigating hypoxia in aquatic environments: diverse approaches to addressing a complex phenomenon
.
Biogeosciences
11
,
1215
-
1259
.
Fry
,
F. E. J.
,
Hart
,
J. S.
(
1948
).
The relation of temperature to oxygen consumption in the goldfish
.
Biol. Bull.
94
,
66
-
77
.
Fu
,
S.-J.
,
Brauner
,
C. J.
,
Cao
,
Z.-D.
,
Richards
,
J. G.
,
Peng
,
J.-L.
,
Dhillon
,
R.
and
Wang
,
Y.-X.
(
2011
).
The effect of acclimation to hypoxia and sustained exercise on subsequent hypoxia tolerance and swimming performance in goldfish (Carassius auratus)
.
J. Exp. Biol.
214
,
2080
-
2088
.
Gracey
,
A. Y.
,
Troll
,
J. V.
and
Somero
,
G. N.
(
2001
).
Hypoxia-induced gene expression profiling in the euryoxic fish Gillichthys mirabilis
.
Proc. Natl. Acad. Sci. USA
98
,
1993
-
1998
.
Greco
,
A. M.
,
Gilmour
,
K. M.
and
Fenwick
,
J. C.
(
1995
).
The effects of softwater acclimation on respiratory gas transfer in the rainbow trout Oncorhynchus mykiss
.
J. Exp. Biol.
198
,
2557
-
2567
.
Holeton
,
G. F.
and
Randall
,
D. J.
(
1967
).
The effect of hypoxia upon the partial pressure of gases in the blood and water afferent and efferent to the gills of rainbow trout
.
J. Exp. Biol.
46
,
317
-
327
.
IPCC
(
2014
).
Climate Change 2014: Impacts, Adaptation and Vulnerability. Part B: Regional Aspects
. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (eds
V. R.
Barros
,
C. B.
Field
,
D. J.
Dokken
,
M. D.
Mastrandrea
,
K. J.
Mach
,
T. E.
Bilir
,
M.
Chatterjee
,
K. L.
Ebi
,
Y. O.
Estrada
,
R. C.
Genova
, et al. 
).
Cambridge
:
Cambridge University Press
.
Itazawa
,
Y.
and
Takeda
,
T.
(
1978
).
Gas exchange in the carp gills in normoxic and hypoxic conditions
.
Resp. Physiol.
35
,
263
-
269
.
Jensen
,
F. B.
,
Fago
,
A.
and
Weber
,
R. E.
(
1998
).
Chapter 1: Hemoglobin structure and function
.
Fish Physiol.
17
,
1
-
40
.
Keys
,
A. B.
(
1930
).
The measurement of the respiratory exchange of aquatic animals
.
Biol. Bull.
59
,
187
-
198
.
Lai
,
J. C. C.
,
Kakuta
,
I.
,
Mok
,
H. O. L.
,
Rummer
,
J. L.
and
Randall
,
D.
(
2006
).
Effects of moderate and substantial hypoxia on erythropoietin levels in rainbow trout kidney and spleen
.
J. Exp. Biol.
209
,
2734
-
2738
.
Lilly
,
L. E.
,
Blinebry
,
S. K.
,
Viscardi
,
C. M.
,
Perez
,
L.
,
Bonaventura
,
J.
and
McMahon
,
T. J.
(
2013
).
Parallel assay of oxygen equilibria of hemoglobin
.
Anal. Biochem.
441
,
63
-
68
.
McBryan
,
T. L.
,
Healy
,
T. M.
,
Haakons
,
K. L.
and
Schulte
,
P. M.
(
2016
).
Warm acclimation improves hypoxia tolerance in Fundulus heteroclitus
.
J. Exp. Biol.
219
,
474
-
484
.
Mitrovic
,
D.
,
Dymowska
,
A.
,
Nilsson
,
G. E.
and
Perry
,
S. F.
(
2009
).
Physiological consequences of gill remodeling in goldfish (Carassius auratusc) during exposure to long-term hypoxia
.
Am. J. Physiol.
297
,
R224
-
R234
.
Mora
,
C.
and
Maya
,
M. F.
(
2006
).
Effect of the rate of temperature increase of the dynamic method on the heat tolerance of fishes
.
J. Therm. Biol.
31
,
337
-
341
.
Mueller
,
C. A.
and
Seymour
,
R. S.
(
2011
).
The regulation index: a new method for assessing the relationship between oxygen consumption and environmental oxygen
.
Physiol. Biochem. Zool.
84
,
522
-
532
.
Ong
,
K. J.
,
Stevens
,
E. D.
and
Wright
,
P. A.
(
2007
).
Gill morphology of the mangrove killifish (Kryptolebias marmoratus) is plastic and changes in response to terrestrial air exposure
.
J. Exp. Biol.
210
,
1109
-
1115
.
Regan
,
M. D.
and
Richards
,
J. G.
(
2017
).
Data from: Rates of hypoxia induction alter mechanisms of O2 uptake and the critical O2 tension of goldfish
.
Dryad Digital Repository
.
Regan
,
M. D.
,
Gosline
,
J. M.
and
Richards
,
J. G.
(
2013
).
A simple and affordable calorespirometer for assessing the metabolic rates of fishes
.
J. Exp. Biol.
216
,
4507
-
4513
.
Regan
,
M. D.
,
Gill
,
I. S.
and
Richards
,
J. G.
(
2017
).
Calorespirometry reveals that goldfish prioritize aerobic metabolism over metabolic rate depression in all but near-anoxic environments
.
J. Exp. Biol.
220
,
564
-
572
.
Rodewald
,
K.
and
Braunitzer
,
G.
(
1984
).
Die Primärstruktur des Hämoglobins vom Goldfisch (Carassius auratus)
.
Hoppe-Seyler's Z. Physiol. Chem.
365
,
95
-
104
.
Rogers
,
N. J.
,
Urbina
,
M. A.
and
Reardon
,
E. E.
(
2016
).
A new analysis of hypoxia tolerance in fishes using a database of critical oxygen level (Pcrit)
.
Conserv. Physiol.
4
,
1
-
19
.
Smith
,
V. H.
,
Joye
,
S. B.
and
Howarth
,
R. W.
(
2006
).
Eutrophication of freshwater and marine ecosystems
.
Limnol. Oceanogr.
51
,
351
-
355
.
Snyder
,
S.
,
Nadler
,
L. E.
,
Bayley
,
J. S.
,
Svendsen
,
M. B. S.
,
Johansen
,
J. L.
,
Domenici
,
P.
and
Steffensen
,
J. F.
(
2016
).
Effect of closed v. intermittent-flow respirometry on hypoxia tolerance in the shiner perch Cymatogaster aggregata
.
J. Fish Biol.
88
,
252
-
264
.
Sollid
,
J.
,
De Angelis
,
P.
,
Gundersen
,
K.
and
Nilsson
,
G. E.
(
2003
).
Hypoxia induces adaptive and reversible gross morphological changes in crucian carp gills
.
J. Exp. Biol.
206
,
3667
-
3673
.
Sollid
,
J.
,
Weber
,
R. E.
and
Nilsson
,
G. E.
(
2005
).
Temperature alters the respiratory surface area of crucian carp Carassius carassius and goldfish Carassius auratus
.
J. Exp. Biol.
208
,
1109
-
1116
.
Steffensen
,
J. F.
(
1989
).
Some errors in respirometry of aquatic breathers: How to avoid and correct for them
.
Fish Physiol. Biochem.
6
,
49
-
59
.
Sundin
,
L.
,
Nilsson
,
G. E.
,
Block
,
M.
and
Lofman
,
C. O.
(
1995
).
Control of gill filament blood flow by serotonin in the rainbow trout, Oncorhynchus mykiss
.
Am. J. Physiol. Regul. Integr. Comp. Physiol.
268
,
R1224
-
R1229
.
Svendsen
,
M. B. S.
,
Bushnell
,
P. G.
and
Steffensen
,
J. F.
(
2016
).
Design and setup of intermittent-flow respirometry system for aquatic organisms
.
J. Fish Biol.
88
,
26
-
50
.
Turko
,
A. J.
,
Cooper
,
C. A.
and
Wright
,
P. A.
(
2012
).
Gill remodelling during terrestrial acclimation reduces aquatic respiratory function of the amphibious fish Kryptolebias marmoratus
.
J. Exp. Biol.
215
,
3973
-
3980
.
Turko
,
A. J.
,
Robertson
,
C. E.
,
Bianchini
,
K.
,
Freeman
,
M.
and
Wright
,
P. A.
(
2014
).
The amphibious fish Kryptolebias marmoratus uses different strategies to maintain oxygen delivery during aquatic hypoxia and air exposure
.
J. Exp. Biol.
217
,
3988
-
3995
.
Tzaneva
,
V.
,
Gilmour
,
K. M.
and
Perry
,
S. F.
(
2011
).
Respiratory responses to hypoxia or hypercapnia in goldfish (Carassius auratus) experiencing gill remodelling
.
Resp. Physiol. Neurobiol.
175
,
112
-
120
.
Vulesevic
,
B.
and
Perry
,
S. F.
(
2006
).
Developmental plasticity of ventilatory control in zebrafish, Danio rerio
.
Resp. Physiol. Neurobiol.
154
,
396
-
405
.
Wegner
,
N. C.
(
2011
).
Gill respiratory morphometrics
. In
Encyclopedia of Fish Physiology Volume 2: Gas Exchange, Internal Homeostasis and Food Uptake
(ed.
A. P.
Farrell
), pp.
803
-
811
.
San Diego, CA
:
Elsevier
.
Wood
,
S. C.
and
Johansen
,
K.
(
1972
).
Adaptation to hypoxia by increased Hb-O2 affinity and decreased red cell ATP concentration
.
Nature
237
,
278
-
279
.
Yeager
,
D. P.
and
Ultsch
,
G. R.
(
1989
).
Physiological regulation and conformation: a BASIC program for the determination of critical points
.
Physiol. Zool.
62
,
888
-
907
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information