Rosalind Bown

Jemma Ludley

Rebecca Johnston

Maddison Dorrell

Anna Stevens

Figure 0.1 - Falmouth fieldcourse 2012 - Group 8 on Xplorer.

Samantha Earl

Stephen Nation

Jiawei Wu

Russell Somerville

 

1. Introduction                                                                                                Return to Top of Page

This website aims to summarise the key findings and scientific procedures carried out in the waters surrounding Falmouth by Group 8 students of the University of Southampton during a two week fieldcourse. It will be based around the data collected from three practical areas:

  • an offshore coastal practical,

  • an estuary based practical and

  • a side scan habitat mapping geophysics based practical

It will hold a catalogue of raw and processed data that will then be accessible to everyone.

The aim of the fieldcourse was to further our knowledge and understanding of the physical, chemical and biological interactions/systems occurring in the Falmouth area by processing and analysing the subsequent data collected. Detailed aims of each practical are discussed below in the relevant sections. 

The Fal Estuary, also know as Carrick Roads has a total shoreline length of 127 km and stretches from its entrance at Pendennis Point and St Anthonys Head to its northern most tidal limit at Tresillian. The Fal estuary was formed due to a eustatic sea level rise at the start of the last Holocene epoch, this drowned the former river present and subsequently created the ria present now. Carrick roads is characterised by deep meandering channels, that reach depths up to 34 m; the estuary itself is also macrotidal with spring tides reaching 5.3 m, however is mesotidal at Truro with spring tides only reaching 3.5 m.

Fal Harbour is the third largest harbour in the world and is continually being exploited both recreationally and commercially; as a result these activities expose the area to various anthropogenic pressures affecting the physical, chemical and biological nature of the estuary. Anthropogenic pressures can also lead to episodes of eutrophication caused by the enrichment of nitrate, phosphate and other trace metals by inputs e.g.: metal mining in the area, agricultural runoff and sewage pumping, as well as other contaminants from Falmouth dockyard. This increase in nutrients consequently can lead to toxic dinoflagellate blooms (Alexandrium minitum).

Surrounding Falmouth waters, there are several long-term time series stations based near Plymouth (L4 and E1) allowing comparison in the area. Offshore there are also tidal fronts present where cold, well-mixed waters meet stratified waters with warm surface layers and cooler deeper layers. This allows for plankton blooms (Pingree R. D., 1975). The area is a Special Area of Conservation (SAC) due to various important biological habitats being present e.g. Maerl beds (Phymatolithon calcareum).

These various factors make the Fal Estuary and its surrounding waters an ideal place to conduct continuous surveys and collect long-term data sets in order to monitor and observe the various changes that occur.

 

Return to Top of Page

 

2. Group Schedule                                                                                         Return to Top of Page

Group 8 followed the schedule tabulated below, with additional time in the evenings occupied with further analysis and data write-up sessions.

 

Date Day Activity
 26/06/2012 Tuesday Introduction Session
27/06/2012 Wednesday Offshore Boat
 28/06/2012 Thursday Offshore Biology Lab & Chemistry Lab
 29/06/2012 Friday Website and Computer Training
 30/06/2012 Saturday Estuary Boat
 01/07/2012 Sunday Catch-up Day
02/07/2012 Monday Estuary Lab (am)
03/07/2012 Tuesday Geophysics Boat
04/07/2012 Wednesday Geophysics Lab and Pontoon
 05/07/2012 Thursday Write-up and Data Day
06/07/2012 Friday Write-up and Data Day
 07/07/2012 Saturday Finish

Figure 2.1 - Timetable of scheduled Falmouth fieldcourse 2012 activities for Group 8.

 

Return to Top of Page

 

3. Procedures                                                                                                  Return to Top of Page

Introduction

All methods used in the lab to analyse the biological and chemical samples collected on both the R.V. Callista and R.V. Bill Conway were taken from the following journals:

Manual chlorophyll, dissolved phosphate and silicon:

     Parsons T. R. Maita Y. and Lalli C., 1984, 'A manual of chemical and biological methods for seawater analysis.',       173 p. Pergamon.

 Dissolved oxygen:

 Grasshoff, K., K. Kremling, and M. Ehrhardt., 1999, 'Methods of seawater analysis.', 3rd ed. Wiley-VCH. 

Nitrate by flow injection analysis:

             Johnson K. and Petty R.L., 1983, 'Determination of nitrate and nitrite in seawater by flow injection analysis.',  Limnology and Oceanography 28 1260-1266.

 

Return to Top of Page

 

4. Offshore                                                                                                Return to Top of Page

Introduction

The R.V. Callista was taken offshore on 27th June 2012 into deeper coastal waters off Falmouth, Cornwall, UK, to investigate:

Figure 4.1 - R.V. Callista.

How vertical mixing processes in the waters off Falmouth affect, directly and indirectly, the structure and functional properties of plankton communities?

Surrounding Falmouth, in the western English Channel, waters show thermal stratification in the summer months due to an increase in solar radiation. This stratification is controlled by water depth and tidal strength, which thus determine the shear and flow characteristics in the water column.

In the western English Channel there is a prominent seasonal cycle of phytoplankton growth, where blooms can be observed in the spring and autumn. These have been investigated primarily at the L4 and E1 stations for a number of years (Western Channel Observatory, 2007). These blooms are largely controlled by vertical mixing of the surrounding waters, which in turn influence the physical and chemical properties of the surface waters.

The predictable spring bloom is a response to increased solar radiation and the resultant stratification, this ensures the phytoplankton remain in the nutrient rich surface waters which they rapidly colonise. However in the summer months nutrient levels deplete, the population collapses and the phytoplankton cells die off, being grazed upon and sinking to deeper waters (Winder & Cloern, 2010).

The stratification in the summer months controls the rate at which primary nutrients are mixed upwards across the thermocline. The rate of this process determines the variations in biological productivity at this time of year.

The changing levels of three primary nutrients: phosphate, dissolved silicon and nitrate, determine the composition of the phytoplankton community (Officer & Ryther., 1980) and vice versa.  As stipulated by Pingree et al. (1977), the fluctuating nutrient levels mirror the biological and chemical processes occurring within the water. This therefore controls the distribution, abundance and variety of organisms within a community.

Offshore an ADCP (Acoustic Doppler Current Profiler) was used to measure Eulerian flow at a series of points. This provided a cross-sectional velocity profile within the water column. These demonstrate how the seabed affect the flow.

A CTD (Conductivity Temperature Depth) probe was used to record temperature and salinity with depth, allowing  density within the water column to be calculated. From this data it can be determined whether the estuary is undergoing turbulent or laminar mixing. The Richardson number (Ri) is a dimensionless number used to show whether turbulent or laminar flow is present in the vertical water column.

 

Calculating the Richardson Number

Figure 4.2 - Formula for Richardson Number (Ri) calculation.

If the Richardson number (Ri) calculated is less than 0.25 then the water column is turbulent and there may be a strong shear present for which a lot of energy will be required to overturn the water column (weak radiance). If the number is greater than 1.00, flow is laminar, and no vertical mixing is occurring. This suggests that there is a stable flow within an area or a strong density gradient between the water layers. Thus indicating overturning will not occur under this conditions. The area between these two values (know as a 'grey area'ť) portrays real world conditions whereby there is a transition between the two water states of turbulent and laminar flow (Stewart R, 2005).

Below maps the sampled stations and the transects taken within an embedded Google Map. Station longitudes and latitudes are listed within the key beneath the map.

 

Figure 4.3 - Offshore locations of sampled stations and linear transects (between stations).

 

Equipment

R.V. Callista

The offshore vessel owned by the School of Ocean and Earth Science at the University of Southampton was used for this practical.

Acoustic Doppler Current Profiler (ADCP)

A sonar that records current velocities and backscatter as the vessel moves along a track. R.V. Callista has a hull mounted ADCP, 600 KHz profiler. Collected data is transmitted to computers on board where it can be analysed.

CTD

The CTD measured Conductivity, Temperature and Depth of the water column as it was lowered through the water column. Aboard R.V. Callista, a CTD was attached to a rosette sampler with six Niskin bottles assembled to collect water samples at set depths following analysis of data. The data is retrieved in real time via an umbilical cable linked to the vessel. Advantages of such equipment include the high resolution sampling of data. Yet conversely, disadvantages include that only one point in time is sampled, and many samples would be required to give a detailed overall picture of the water column. Calibration of such instruments against a lab or secondary standard must occur prior to use.

Niskin Bottle

A plastic cylinder sent to a certain depth allowing water to flow through the open ends until reaching the required depth. A messenger weight is sent down the attached cable to close the bottles' ends, and trap the water in the cylinder. Aboard R.V. Callista, six bottles were attached to the rosette sampler together with the CTD. Water collected from these bottles can be analysed, and the nutrient concentration of the water column calculated. The percentage of dissolved oxygen at that the sampled depth can also be determined. Advantages of using a rosette sampler with Niskin bottles include ease and simplicity of collecting multiple samples and being good for small scale sampling. Many bottles can be fired at once collecting more data and again getting a broader picture of the water column in a set time (both cost and time efficient).

Fluorometer

A sensor attached to the rosette sampler along with the CTD and bottles. Used to measure fluorescence in the water column by exciting a specific wavelength of light and monitoring returning pulses. Data can be used to calculate chlorophyll fluorescence which can be used as a proxy to indicate the amount of phytoplankton present in known water depths/samples. Once again, calibration of such instruments is required.

Transmissometer

Measures optical properties of the water column by measuring light attenuation. A beam of light is transmitted through a known path length and the time taken to reach the receiver (at the end of the path length) is recorded. The sensor is attached to the rosette along with the CTD and Fluorometer, and requires calibrations (light and dark in air calibrations).

 

Data Analysis

CTD

(a)    (b)    (c) 

(d)    (e)    (f) 

Figure 4.4 - CTD Offshore Data collected from Station 1 to Station 4 - Falmouth, UK - (a) Temperature, (b) Salinity, (c) Density, (d) Fluorescence, (e) Irradiance, and (f) ln(irradiance). Sampled on 27 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

Temperature [Figure 4.4a]

The vertical profiles at all stations offshore show a decrease in temperature from the surface waters to the bottom waters [Figure 4.4a]. At Station 1 the difference between surface (14.3 °C) and bottom (12.5 °C) waters was approximately 1.8 °C. The profile at this station showed a rapid decrease from the surface to 3.5 m, after which the temperature reduction was observed as more gradual. The rapid change is likely to be a result of the introduction of warmer, fresher river water flowing through the estuary to the mouth (where Station 1 was sampled). This warmer water is at the surface because it has lower density and so is lighter and floats at the surface. This warmer fresher water can also be seen in the salinity and density profiles for Station 1 [Figure 4.4b and Figure 4.4c].

At Station 2 and Station 3 two thermoclines were observed. One from the surface to approximately 10 m in depth and the other from 47 m to approximately 60 m in depth. The surface thermocline is likely to be a seasonal thermocline produced due to solar heating of the surface water during the summer resulting in an increase in the temperature. The difference in temperature is 0.3 °C at Station 2 and 0.6 °C at Station 3. Usually for this time of year the seasonal thermoclines are stronger with temperature differences of the order of 2 °C to 4 °C (Western Channel Observatory L4 and E1 buoy data). On the day of sampling, and during previous weeks of June 2012, the weather had been cloudy and of high precipitation, thus reducing the amount of incoming short wave radiation and a reduction in the solar heating of the surface water and so the thermocline present was weaker. Despite this the presence of this thermocline indicates that the water column is stratified. The thermocline present near the bottom of the water column has a temperature difference of 0.5 °C at Station 2 and 0.8 °C at Station 3, and at this depth could signify the convergence of two water masses at Station 2 that have different water properties such that one water mass is downwelling under the other. This idea is further supported by the presence of a high fluorescence peak at Station 2 at a depth of 49 m [Figure 4.4d]. Also witnessed at Station 3 was a temperature inversion at 25 m. This means that the water temperature has increased, in this case by 0.1 °C, and could result from mixing between the converging two water masses seen at Station 2, therefore producing changes to the properties of the water. This temperature inversion was balanced by an increase in salinity of 0.05 psu between 25 m and 30 m [Figure 4.4b] meaning the density at this depth was not significantly changed. Station 4 also shows a thermocline near to the surface decreasing to approximately 14 m with a temperature difference of 1.0 °C from 13.4 °C to 12.4 °C. The bottom thermocline is still present at Station 4 but is less prominent compared to Station 2 and Station 3.

Salinity [Figure 4.4b]

All four profiles show an increase in the salinity with depth at Station 1 showing a significant increase and Station 4 a very small increase [Figure 4.4b]. The stations are positioned at increasing distances from the shore meaning that the salinity is also increasing from the coast to offshore as the influence of the higher salinity seawater becomes more dominant. The lower salinity seen in Station 1 is the result of the input of warmer and fresher riverine water into the estuary as well as the high precipitation input experienced in the month of June 2012. This input of fresher water dilutes the water incoming from the sea thereby lowering the salinity value of the water and combined with the warmer temperature mentioned above, the resulting water density is lower (1026.3 kgm-ł).

As mentioned before there is a more rapid increase in salinity at Station 3 at 25 m depth coinciding with the temperature inversion at this depth. This balance prevents the density of the water from changing significantly and therefore means that no overturning of the water column occurs.

There are several sharp peaks seen in the vertical salinity profiles in particular near the surface at Station 1, 3 m at Station 3 and 13 m at Station 4. These peaks are known as salinity spikes and occur in all CTD profiles as a result of the time response difference between the conductivity and temperature sensors on the CTD. They commonly occur in thermocline areas where the temperature changes are rapid and thus the faster conductivity response results in a spike in the data.

Density [Figure 4.4c]

At all four stations the density of the water increases with depth [Figure 4.4c]. This is because density is controlled by the temperature and salinity characteristics of the water. Water is denser when it is colder and more saline and from the temperature and salinity graphs the coldest and most saline water is found at the bottom of the water column and therefore this is where it has the greatest density. These density profiles show that the water column at all four stations is statically stable as the lower density water is on top of the higher density water and no overturning of the water column is occurring.

Fluorescence [Figure 4.4d]

The vertical profiles for fluorescence vary with depth [Figure 4.4d]. At Station 1 surface fluorescence values were 0.10 mgm-3 and increased to a peak value of 0.17 mgm-3 at 4.2 m depth. This peak occurred due to the chlorophyll pigments present in the sample fluorescing, signifying the presence of phytoplankton at this depth. This correlates to the recorded irradiance data, where the euphotic zone at this station reaches 11 m. This is the level where enough light is available for phytoplankton to photosynthesise. At Station 2 there were two fluorescence peaks of 0.2 mgm-3 at 7 m and 0.11 mgm-3 at 49 m. The 7 m peak was due to a chlorophyll maxima resulting from phytoplankton in the euphotic zone. The peak at 49 m could be a deep chlorophyll maximum due to the convergence of the two water masses one of which contained a high chlorophyll concentration/large phytoplankton numbers and so this downwelling water has subducted the chlorophyll/plankton also. As the light level is low, it is unlikely that the phytoplankton are able to photosynthesise at this depth. There are also fluorescence peaks at Station 3 at depths of 9 m to 11 m (0.26 mgm-3) and at Station 4 at depths of 12 m (0.28 mgm-3)  and 22 m (0.23 mgm-3) respectively. The peak at 22 m at Station 4 may result from sinking phytoplankton rather than actively photosynthesising plankton due to irradiance levels of only 10 µmolm-˛s-1 which is too low for photosynthesis. The presence of a high number of zooplankton at this depth supports the theory of a chlorophyll maxima at this depth since zooplankton feed on phytoplankton. So if a high number of zooplankton are present then logically one would assume there to be a food source present to sustain them.

The graph of chlorophyll against depth derived from acetone extraction [Figure 4.11] shows peaks of chlorophyll at a similar concentration to that derived from the fluorometer [Figure 4.4d]. At Station 1, the acetone derived chlorophyll graph has a peak at 27 m with a value of 5 µgL-1. However the main chlorophyll fluorescence peak from the CTD at 4.2 m is not seen in the acetone extracted chlorophyll values. At Station 2 the deeper chlorophyll maxima perceived on the CTD is also seen in the acetone extracted chlorophyll data where values are 4.4 µgL-1 thereby signifying that this is not just an unexpected result from the CTD only. Station 3 was only a CTD cast so no water bottle samples were taken and therefore there is no acetone extracted chlorophyll values for comparison at this station. At Station 4 there is a peak in chlorophyll extracted from the acetone at 13 m with a value of 5.6 µgL-1 which once again coincides with the peak seen by the fluorometer from the CTD data. This supports the idea that there is likely to by a chlorophyll maxima at this depth.

When compared to the chlorophyll data measured by the fluorometer on the CTD the chlorophyll values extracted from acetone have higher values. The highest acetone derived chlorophyll value is 5.6 µgL-1 whereas the highest chlorophyll concentration from the fluorometer is 0.28 mgm-3. The reason for this difference is due to the chlorophyll extracted by the acetone measuring all the chlorophyll present in the water column at that particular depth whereas the fluorometer only measures the fluorescence by the chlorophyll that have been excited by the blue light emitted and therefore any other photosynthetic pigments present that are not excited by the blue light will not register on the fluorometer. This means that the acetone extracted chlorophyll values are more representative of the amount of chlorophyll and therefore more indicative of the concentration of phytoplankton present in the water column and each of the stations (chlorophyll concentration is regularly used as a proxy for phytoplankton concentration, but gives no indication on population size, as numbers, biomass, and size are all factors).

Irradiance [Figure 4.4e and Figure 4.4f]

For each station the irradiance profiles show an exponential decrease with depth, with rapid attenuation of light occurring at shallower depths (between 0 m and 10 m) and slower attenuation occurring below 10 m [Figure 4.4e]. Irradiance decreases exponentially as a result of the attenuation of light from both absorbance and scattering. Light is absorbed by the photosynthetic pigments of phytoplankton in the water column as well as by other particles in the water and the water molecules themselves. Light can also be scattered by particles in water particularly if high suspended particulate matter is present. Consequently, the amount of light reaching greater depths decreases due to these factors. The surface irradiance values for each station varied from 410 µmolm-2s-1 (at Station 1) to 1600   µmolm-2s-1 at Station 3. This variation is likely to be as a result of the weather conditions present (in particular the cloud cover affecting the irradiance reaching the surface waters). Other factors that affect the amount of radiation reaching the surface waters are the time of day which affects the height of zenith, zenith angle, absorption and scattering in the atmosphere. The 1% value of the surface irradiance can be used to give an estimate of the depth of the euphotic zone (and so the depth to which photosynthesis is occurring). At Station 1 the 1% irradiance value occurs at 11.5 m. At Station 3 the 1% irradiance value occurs at 8.7 m depth.

When the natural log of irradiance is taken the graphs produced closely resemble straight line relationships (particularly from Station 1). The plots for the other stations are however not perfect straight lines since other factors  affect light attenuation in the water column. The natural log of irradiance can be used to calculate the attenuation coefficient which helps to gain a quantitive value of how much light radiation is absorbed per metre. The vertical lines at the bottom of the plots are due to the fact the irradiance values remain consistently low from a depth of approximately 50 m.

 

ADCP

(a)   (b)    (c)    (d) 

(e)    (f)    (g)    (h) 

 Figure 4.5 - ADCP Offshore Data collected from Station 1 to Station 4 - Falmouth, UK - (a) Velocity magnitude - Station 1, (b) Average backscatter - Station 1, (c) Velocity magnitude - Station 2, (d) Average backscatter - Station 2, (e) Velocity magnitude - Station 3, (f) Average backscatter - Station 3, (g) Velocity magnitude - Station 4, (h) Average backscatter - Station 4. Sampled on 27 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

Richardson Number Graphs:

(a)    (b) 

Figure 4.6 - ADCP Offshore Data of Richardson Numbers (Ri) calculated over a 1 metre layer (a) Station 1 and Station 2 and (b) Station 3 and Station 4. Sampled on 27 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

Station 1 [Figure 4.5a and Figure 4.5b] [Figure 4.6a]

The ADCP profile of velocity magnitude shows variable velocity in flow with values from slow velocities of 0.011 ms-1 to 0.379 ms-1 with a greater proportion of the flow flowing at lower velocities at this station. The ship track profile shows that there is a minimum amount of shear in terms of flow in different directions but there is shear occurring due to flow at different velocities. The backscatter data from Station 1 shows very high average backscatter between 2 m and 10 m depths with values of greater than 80 db. The highest backscatter value recorded at this station was 93 db. These high values could be produced by zooplankton in the surface water column but a high backscattering signal can also be the result of turbulence in the water column.

Between 5 m and 8 m the Richardson number (Ri) is below 0.25 and so the water column at this point is unstable and there is turbulent flow. This is likely to be a result of the strong winds causing stress and waves on the surface and below the surface of the water. High water was at 10:20 UTC on this day, so the turbulent flow may also have been as a result of the ebbing tide. From 8 m to 11 m and 16 m to 19 m the Richardson number is above 0.25 and therefore flow is stable and laminar so the high backscatter values of 76 db to 78 db between 10 m and 20 m is more likely to be produced by zooplankton with some backscatter resulting from the turbulent flow at 11 m to 16 m. Except for 25 m to 27 m and 30 m to 32 m the rest of the water column exhibits turbulent flow [Figure 4.6a].

Station 2 [Figure 4.5c and Figure 4.5d] [Figure 4.6a]

The ADCP profile of velocity magnitude also shows variable velocity in flow at Station 2 with values from slow velocities of 0.005 ms-1 to high velocities of 0.614 ms-1 with a greater proportion of the flow flowing at lower velocities at this station. The ship track profiles show that at the shallower depths (2 m to 5 m) and the deeper depths (39 m to 48 m) of the water column there is significant shear occurring with flow travelling in different directions and at different velocities relative to one another.

The backscatter data from Station 2 also shows very high backscatter between 2 m and 10 m depths with values of greater than 80 db and high backscatter (values above 70 db) up to 20 m depths [Figure 4.5d]. The highest backscatter value recorded at this station is 106 db. Due to the waves created by the weather the very high backscatter is likely to be created by these whereas the high values lower down is more likely to be created by zooplankton. The surface waters of Station 2 have values above 0.25 with a Richardson number high of 8.5. This shows that the waters are laminar which coincides with the presence of the thermocline at the surface creating stratification and the laminar flow. Below 34 m to 44 m the Richardson number is lower than 0.25 which may be to do with the mixing occurring between the two converging water masses creating small scale turbulence. The second thermocline seen at Station 2 has formed a small laminar flow layer at 45 m where the Richardson number is above 0.25 (Ri = 1.17) [Figure 4.6a].

Station 3 [Figure 4.5e and Figure 4.5f] [Figure 4.6b]

The ADCP profile of velocity magnitude at Station 3 shows higher velocity flow than at Stations 1 and 2 with values from 0.195 ms-1 to 0.673 ms-1. The ship track profiles show that at the shallower depths (2 m to 10 m) of the water column there is significant shear occurring with flow travelling in different directions and at different velocities relative to one another but below 15 m the direction of flow is relatively constant (approximately 215°) [Figure 4.5e].

For a large proportion of the water column at Station 3 the backscatter is low with values of 63 db to 68 db with only a thin layer of water at the surface exhibiting large backscatter. Once again due to the weather this is likely to be the result of wind and wave mixing causing scattering of the sound emitted from the ADCP. This weather affect is seen in the Richardson number graph where the surface waters have a Richardson number below 0.25 representing turbulent flow at which point the water column has instability [Figure 4.6b]. The Richardson numbers with depth at Station 3 are below 0.25 for the majority of the water column except at 7.5 m to 13.7 m, 21 m to 27 m and 40 m to 46 m where the Richardson numbers are 2.67, 2.29 and 12.2 respectively, indicating that at these depths the flow is stable and laminar.

Station 4 [Figure 4.5g and Figure 4.5h] [Figure 4.6a]

The ADCP profile of velocity magnitude at Station 4 also shows higher velocity flow than at Station 1 and Station 2 with values from 0.087 ms-1 to 0.946 ms-1. The ship track profiles show that at the shallower depths (2 m to 10 m) and the deeper depths (43 m to 50 m) of the water column there is significant shear occurring with flow travelling in different directions and at different velocities relative to one another but below 12 m the direction of flow is relatively constant (approximately 235°) [Figure 4.5g].

Similar to Station 3 the average backscatter for a large proportion of the water column at Station 4 is low with values of 64 db to 70 db. There is an area at the surface with very high backscatter (115 db) which is likely caused by turbulence which is shown from the Richardson number graph with values below 0.25. The relatively high backscatter (76 db) in the top left of the backscatter contour could be related to high zooplankton numbers in this layer scattering the sound. This explanation is more likely particularly at depths of 13 m to 18 m where the Richardson number is above 0.25 and so the flow is laminar and the water column stable - thus not likely to be scattering the sound.

Dissolved Oxygen

Figure 4.7 - Offshore dissolved oxygen saturation. Sampled on 27 June 2012. Graph can be enlarged in a new window upon clicking on the relevant image.

The samples from Station 1 show percentage of dissolved oxygen (%O2) saturation increased from a depth of 1.16 m where saturation was 100.0% to 6.30 m where saturation was at 103.6%. Station 2 and Station 4 generally showed a decrease in %O2 saturation with depth. The decrease in %O2 saturation at Station 2 between the depths 0.80 m and 6.98 m with 3.6% with %O2 saturation values of 114.0% and 110.4% respectively (decreased at a rate of 0.58% m-1 depth). There was also a 3.6% decrease between the depths 6.98 m and 26.98 m with %O2 saturation values of 110.4% and 94.5% respectively (decreased at a rate of 0.79% m-1 depth), which was slightly more gradual than at Station 4. At depths of 1.34m and 13.12m Station 4 showed higher %O2 saturation values of 115.96% and 103.53% respectively (decreased at a rate 1.06% m-1 depth). Between the depths of 13.12m and  54.99m, Station 4 showed a more gradual rate of decrease in %O2 saturation with depth (0.32% m-1 depth) when compared with Station 2 (0.79% m-1 depth). Station 2 showed a slight spike in %O2 saturation between a depth of 26.983m and 47.897m from 94.5% to 99.9% (5.37% increase) after which %O2 saturation gradually decreased again with depth. Station 2 and Station 4 showed much higher surface %O2 saturation (114.0% and 115.96% respectively) than that of Station 1 (100.00%).

 

Phosphate

Figure 4.8 - Offshore dissolved phosphate concentration. Sampled on 27 June 2012. Graph can be enlarged in a new window upon clicking on the relevant image.

During the offshore investigation into phosphate concentrations Niskin Bottles were used to collect water samples at three stations. The samples were places in a spectrophotometer to analyse, this piece of equipment had a detection limit of 0.03 µmolL-1 for phosphate. From this data it was found that Station 1 had a high surface phosphate concentration of 0.333 μmolL-1 which decreased to 0.169 μmolL-1 by 6 m depth; this is the minimum which then rises again at 25 m.

These lower values correspond to the fluorescence peak observed on the CTD data which could be an indication of a chlorophyll maximum. The increased chlorophyll levels may be linked to an abundance of phytoplankton at a depth of 6 m, whereas they are not present in the same abundance at depth so phosphate concentrations increase again.

At Station 2 two thermoclines were observed along with a chlorophyll maximum at these depths. The phosphate data corresponds with this maximum, with a high value at the surface which again decreases at 6 m the same as Station 1; however the lowest value observed here is 0.075 μmolL-1 compared with 0.169 μmolL-1  at the previous site showing it has been removed more in this area than at Station 1 sampling site.

By 27 m this concentration has dropped to 0.157 μmolL-1  but increases to 0.274 μmolL-1 by 47.9 m as it is being used less at this depth compared to nearer to the surface where there is more biological activity occurring. At 60 m depth two bottle samples were taken showing both a decrease to 0.204 μmolL-1  and an increase to 0.298 μmolL-1  at the same site which could be due to them sitting on slightly different positions along the CTD rosette.

Station 4 also showed the thermocline was still present, here five bottles were fired on the CTD showing very low concentrations of phosphate just under the surface waters at only 0.110 μmolL-1, and this lowered even further by 13 m. No further readings were taken until 55 m depth where the concentration was found to be 0.298 μmolL-1 and it remained around this figure at the next sample depth of 68 m. This drop in concentration may again link to the maximum reading shown by the fluorometer during the investigation at this station.

 

Nitrate

Figure 4.9 - Offshore dissolved phosphate concentration. Sampled on 27 June 2012. Graph can be enlarged in a new window upon clicking on the relevant image.

The nitrate data shows a vertical profile for Station 1 and Station 2. Nitrate is an important macronutrient in phytoplankton growth; it is generally considered to be the limiting nutrient for plant growth in the sea. Nitrate is taken up by phytoplankton and reduced to ammonium so it can be incorporated into carbon skeletons. Therefore, it was hypothesised that when there was a low nitrate concentration, this may correlate with a peak in fluorescence on the CTD data as fluorescence is indicative of the presence of phytoplankton.

At Station 1, the nitrate concentration was 0.417 µmolL-1 in the surface sample, then decreases to a level that was undetectable by the flow injection analysis, due to removal by phytoplankton. The detection limit for the equipment used was 0.1 µmolL-1, meaning that if the concentration was  less than 0.1 µmolL-1, then this was plotted on the graph as 0.09 to show that it was below this value. The maximum nitrate concentration was measured at 27 m, and here the fluorescence levels were low, indicating low phytoplankton presence to utilise the nitrate.

At Station 2, the nitrate levels were undetectable in approximately the top 6 m of water. This minima correlates with the fluorescence peak in the CTD data of nitrate concentration was greatest, where small peaks can be seen at approximately 4 m, 6 m and 8 m. The maximum nitrate concentration is seen when the fluorescence drops to 0.10 mgm-3, as there are fewer phytoplankton to cause removal of nitrate from the water column.

The two main sources of nitrogen in the euphotic zone are either regenerated e.g. bacterial oxidation to nitrate by the benthic decomposition of organic matter (Kemp et al., 1988) below the thermocline replenishes surface waters with nitrate, or new nitrogen e.g. imported from the deep ocean or the atmosphere in the form of ammonium or urea.  Phytoplankton can only take up regenerated forms of nitrogen such as nitrate, which is why removal is seen in correlation with peaks in chlorophyll.

Other possible influences on nitrate levels could be from terrestrial origins that have washed into the estuary and entered the area offshore. The rainy weather conditions recently could have increased the amount of runoff occurring and therefore raised the nitrate levels. This would have particularly affected Station 1 which was located at the mouth of the estuary as this is closest to any terrestrial sources, and has a higher surface concentration of nitrate than Station 2 which was further offshore.

In reality there would be much more variation in nitrate concentration through the water column, but as the number of water samples that could be collected and processed was limited, the data only shows a snapshot of the vertical variations and not a continuous profile.

Dissolved Silicon

Figure 4.10 - Offshore dissolved silicon concentration. Sampled on 27 June 2012. Graph can be enlarged in a new window upon clicking on the relevant image.

Dissolved silicon is present in seawater within silicate, SiO44-. It is utilised within the frustules of diatoms as a structural component, with diatoms accounting for 90% of suspended dissolved silicon in the world's oceans (Harper., 1975). It is also important in other siliceous phytoplankton, such as dinoflagellates. As such, dissolved silicon concentration is an indicator for phytoplankton presence, and vice versa.

At Station 1, the dissolved silicon concentration was almost constant with depth. Over the 20 m sample depth range, there was a decrease of only 0.1 µmolLŻą, from 1.1 µmolLŻą at 1.2 m to 1.0 µmolLŻą at 27.0 m. The replicate analyses of the 6.3 m had a difference of 0.1 µmolLŻą between them, suggesting that the method of analysis was precise. At Station 2, the dissolved silicon concentration increased with depth. There was an overall difference in dissolved silicon concentration of 1.9 µmolLŻą over the 60 m depth range. The highest concentration was 2.4 µmolLŻą at 60.4 m. The two sets of replicates analysed from water sample taken at 27.0 m and 47.9 m both had a difference of 0.1 µmolLŻą between them, suggesting that the method of analysis was precise.

At Station 4, the dissolved silicon concentration increased with depth. There was an overall difference in dissolved silicon concentration of 2.7 µmolLŻą over the 67 m depth range. The highest concentration was 3.1 µmolLŻą at 68.2 m.

In general, there was an exponential increase in the dissolved silicon concentration with increased depth. The nearest sample taken to the surface at each station decreased in dissolved silicon concentration as the stations moved further offshore.

Chlorophyll

Figure 4.11 - Offshore chlorophyll concentration (Niskin bottle). Sampled on 27 June 2012. Graph can be enlarged in a new window upon clicking on the relevant image.

For surface waters the highest chlorophyll concentrations are found to be at Station 2 (> 5 µgL-1). The lowest surface chlorophyll concentrations can be found at Station 1 (Black Rock) (< 3 µgL-1) and intermediate chlorophyll concentration values are located at Station 4 (approximately 3.6 µgL-1). Down to a depth of 28 m chlorophyll concentration at Station 1 continues to increase but at this depth no more samples are taken and it peaks at approximately 5 µgL-1. Chlorophyll concentration at Station 4 peaks at 15 m depth where it reaches concentrations > 5.5 µgL-1. Further down in the water column at approximately 55 m depth the concentration had dropped to approximately 1. 5 µgL-1 which was the joint lowest concentration measured in the offshore practical. Between 55 m and 70 m the chlorophyll concentration increased again to above 3 µgL-1 and sampling did not continue at any greater depths. At Station 2 the highest chlorophyll concentrations are found at the surface and they generally decrease with depth until the lowest concentration is found at 60 m (approximately 1.5 µgL-1).

(a)    (b)    (c)    (d) 

Figure 4.13 - (a) CTD, (b) Water collection from a filled Niskin bottle on deck, (c) Wet chemical lab on R.V. Callista, and (d) Offshore chemical analysis. Sampled on 27 June 2012. All pictures can be enlarged in a new window upon clicking on the relevant image.

Conclusion

By studying all the nutrients together it can be seen that phosphate, nitrate and dissolved silicon all show a relationship with the fluorescence data obtained. When there is a fluorescence peak, indicating a maximum chlorophyll level and therefore large abundance of phytoplankton, this corresponds with a minimum concentration of all nutrients. Theses minimums are most prominent at Station 2 and Station 4 when the greatest peak in fluorescence was seen between 5 m and 15 m in depth. This relationship is expected due to the phytoplankton utilising the nutrients and thus depleting them.

 

Return to Top of Page

 

5. Estuary                                                                                                          Return to Top of Page

Introduction

 


View Falmouth Estuary in a larger map

Figure 5.1 - Estuary locations of sampled stations and linear transects (between stations).

Estuaries are classed as semi-enclosed bodies of water in which two natural water types with very different chemical, biological and physical characteristics mix.  Within the estuary straightforward mixing of saline and freshwater can be seen as well as other processes that subsequently effect the concentrations of various nutrients along the estuary.

Estuaries can be characterised by their salinity structure, which gives rise to three main types of structure: salt wedge, partially mixed and well mixed. A salt wedge is a highly stratified structure - it occurs when there is a large freshwater impact and weak tidal currents. These factors cause a wedge to occur as the less dense freshwater is pushed seaward and sits on top of the denser seawater. In a partially mixed estuary, seawater and freshwater mix throughout the water column although a salinity gradient may still occur. In a well-mixed estuary the salinity remains constant throughout the water column; this is caused by a low river flow combined with a strong tidal flow, which eliminates any stratification that may occur.

The aim of the estuarine fieldwork based on the R.V. Bill Conway was to develop an understanding of the lateral variations in water structure between the freshwater dominant upper estuary compared to the saltwater dominated coastal waters. These observations then allow an understanding of the various mixing processes occurring along the river in respect to nitrate, dissolved silicon and phosphate.   

The processed concentrations of dissolved silicon , nitrate and phosphate collected can then be compared against a conservative component, salinity. Salinity was continuously monitored throughout the tidal cycle to provide a recorded salinity sequence down the estuary. Salinity was then compared against the various nutrient concentrations to produce estuarine mixing diagrams. These then reflect the numerous processes occurring down the river by showing if the nutrient is behaving conservatively or non-conservatively.

The Fal Estuary is characterised as a ria or a drowned river basin with deep meandering channels reaching up to 34 m in depth. The estuary is classified as macrotidal with the mean tidal range reaching 5.3m during spring tides and with tidal currents less than 2 knots.

The aim for Group 8 in the estuary was to collect a range of data throughout a tidal cycle, so to gain a detailed "snap shot" of the estuary over that period of time. From the collected data an understanding of the physical, chemical and biological processes and how they interact can be gained.

 

Equipment

R.V. Bill Conway

The coastal vessel owned by the School of Ocean and Earth Science at the University of Southampton was used for this practical.

Niskin Bottles

Due to damage to the CTD on the R.V. Bill Conway prior to surveying the estuary, a Niskin bottle was used to collect water samples, with a temperature-salinity probe attached to record temperature and salinity data. These were used at each station. The bottle was lowered manually on a winch and hydroline off the side of the boat. For this method of deployment a brass weight messenger was manually released to close the bottles at the required depth.  

The water samples for oxygen concentration were taken via the tap on the bottle, and samples for chemical analysis were taken using the bottom lid of the bottle. Advantages of manual deployment - simple and easy to use, good for small scale sampling. Disadvantages - sampling could be done "blind" with no knowledge of the water column properties if a YSI has not been used. If on a hydroline then the depth will not be exact due to movement once in the water.

Acoustic Doppler Current Profiler (ADCP)

The R.V. Bill Conway has a hull mounted ADCP, at 600 KHz. Same system as used on R.V. Callista.

YSI 6600 V2 Probe

Cylindrical measuring probe that is lowered manually into the water and sends data back to a hand held device. There is a wide range of data obtained from the probe; Depth (m), Chlorophyll (µgL-1), pH, Dissolved Oxygen (%), Temperature (°C), Salinity were sampled from the model above.

Thermosalinograph

Device used to record spot temperatures and salinities whilst travelling through the estuary.

Figure 5.2 - Thermosalinograph. Picture can be enlarged in a new window upon clicking on the relevant image.

Secchi disk-Method

The disc is mounted on a labeled string, and lowered slowly down in the water. The depth at which the pattern on the disk is no longer visible is taken. This was performed three times to yield an average value. The time and weather conditions were also recorded.

 

Data Analysis

YSI

(a)    (b)    (c)    (d)    (e) 

Figure 5.3 - YSI Estuary Data collected from Station 1 to Station 3 - Falmouth, UK - (a) Temperature, (b) Salinity, (c) Chlorophyll, (d) Dissolved O2, and (e) pH. Sampled on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

Temperature [Figure 5.3a]

The temperature at each station decreases with increased depth. Station 1 experiences the highest temperatures overall, showing temperatures of 16.0 °C in the surface waters, and then dropping to 15.2 °C at 2 m. This station may have experienced warmer waters due to this area of the estuary being shallower than other areas. Station 2 shows the largest difference in temperature from surface waters showing temperatures of 15.7 °C to 12 m witnessing 14.0 °C. This could be caused by thermal stratification in the estuarine waters. Station 3 shows the smallest change in temperature with depth: 1°C change over 8 m and may represent the coastal waters being well mixed.

Salinity [Figure 5.3b]

The salinity profiles at all stations show a halocline present in the surface waters, about 1m down. Station 1 and 2 show a vast increase in salinity; Station 1 increases in salinity from 24 to 32 and Station 2 increases from above 29 psu to 33 psu. Station 1 shows the lowest salinity overall, this corresponds to its position in the upper part of the estuary which would experience lower salinity values due to the higher freshwater input. Station 3 also shows the higher salinity values, corresponding to its position closer to coastal waters. The lower salinity values experienced at Station 1 could also be caused by the increased freshwater input from higher levels of precipitation experienced this month.

Chlorophyll [Figure 5.3c]

Chlorophyll was also measured using a fluorometer on the YSI probe. Station 1 and Station 3 show that chlorophyll is constant throughout the water column at approximately 5 ugL-1, from the surface waters to depth. This would suggest that the waters at these stations are well mixed throughout in the vertical. Station 2 however, shows a sharp decrease in concentration from 25 ugL-1 in the surface waters, to under 5 ugL-1 at depths between 1 m and 12 m. This may represent a high proportion of phytoplankton in the surface water masses above the thermocline.

Dissolved Oxygen [Figure 5.3d]

The vertical profile derived for dissolved oxygen at each station shows the variation with depth and a longitudinal change down the estuary. Station 1 shows a lower percentage of oxygen when compared to Station 2 and Station 3. In the surface waters, Station 1 shows a 97% content of dissolved O2; where as Station 2 and Station 3 both show dissolved oxygen content of above 100%. This high percentage of O2 in the surface waters may be caused by a high number of marine phytoplankton present, which subsequently photosynthesise in the photic zone of the surface waters producing oxygen. When a large numbers of phytoplankton are present oxygen does not leave the water quick enough and can produce a dissolved oxygen percentage higher than 100%. For all stations the percentage of dissolved oxygen decreases down the water column, this is caused by phytoplankton not being able to photosynthesise at deeper depths due to a lack of irradiance. However Station 3, situated closest to coastal waters, continuously shows a higher percentage of dissolved O2 in total. This may be caused by a lack of zooplankton present which would utilise the O2 in respiration.

Density

All 3 stations show an increase in density with depth. This was expected as the temperature from the surface to the bottom decreases with depth, thus increasing the density of the water. This combined with the increasing salinity with depth, forms a layer of water of greatest density at the bottom and a less dense water layer at the surface. Overall the water column at all 3 sampled estuarine stations can be described as being statically stable since the lower density water is above the lower salinity water. The lowest density calculated (at the surface of Station 1) was 1015 kgm-ł and would have resulted from the large amount of freshwater flowing down the estuary from the River Allen and the River Kenwyn, coupled with the freshwater surface runoff due to high rainfall over the preceding weeks. The highest density recorded in the estuary was 1023.5 kgm-ł. This was calculated at Station 2 (at a depth of 12 m). The high salinity influence from the sea and the colder water at depth are explanations for this value. Station 2 shows a significant pycnocline between the surface and depth compared with the other sampled stations This is likely due to its position in the estuary where the water is influenced by both riverine and marine water. Combined inputs results in the stratified water column observed.

ADCP

Transect 1  [Figure 5.4a and Figure 5.4b]

 

Figure 5.4 - ADCP Estuary Data collected for Transect 1 - Falmouth, UK - (a) Velocity magnitude -Transect 1, (b) Average backscatter - Transect 1. Data collected on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

As with the offshore data the estuarine stations also show variability in the velocity of the flow. Velocity values along Transect 1 range from 0.058 ms-1 on the edge of the estuary to 0.497 ms-1 in the central channel. Velocities are faster in the central channel as the water here is deeper with a maximum calculated depth from the ADCP of 7.39 m. This greater depth means that less water is in contact with the estuary bed and so is less affected by bottom friction which would slow the velocity of the water. This slowing by friction both by the bottom and the sides of the estuary result in decreased velocities on the outer estuary.

Average backscatter for this transect shows a large area within the central channel from a depth of 2 m to 6 m with a low backscatter of 68 db to70 db. This coincides with the Richardson number for depths 2 m to 3 m which was much greater than 0.25 and therefore flow is laminar and stable and so is not causing disruption to the water column and resulting in backscattering. Further evidence is from the zooplankton counts which were low for this station so low numbers of phytoplankton means there are low number of particles in the water column to scatter the sound back to the transceiver.

Transect 2 [Figure 5.5a and Figure 5.5b]

 

Figure 5.5 - ADCP Estuary Data collected for Transect 2 - Falmouth, UK - (a) Velocity magnitude -Transect 2, (b) Average backscatter - Transect 2. Data collected on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

Transect 2 shows an area of lower velocity to the left side of the central channel and an area of higher velocity at the surface on the right side. The lower velocities have values of 0.005 ms-1 to 0.256 ms-1 and the higher velocities values of 0.286 to 0.619 ms-1. The reason for the higher velocities is due to the incoming flood tide (CISCAG, 2011). In the Fal estuary it has been measured that the flood tide is stronger than the ebb tide. On the sampling day of the 30th June high water was at 13:51 UTC and so when this transect was taken at 11:26 UTC the tide would have been flooding into the estuary towards high tide therefore increasing the velocity of the water.

Similar to the backscatter data collected from Transect 1, Transect 2 also has a very large area in the central channel with low backscatter values of 67 db to 70 db. It is difficult to see from the Richardson number data for this transect because it varies from below 0.25 to above 0.25 across the whole depth range at which the backscatter is low. This then means that the low backscatter is a result of low zooplankton numbers or low suspended particulate matter in the water column at this point in the estuary and so fewer particles to backscatter the sound produced by the ADCP.

Transect 3 [Figure 5.6a and Figure 5.6b]

 

Figure 5.6 - ADCP Estuary Data collected for Transect 3 - Falmouth, UK - (a) Velocity magnitude -Transect 3, (b) Average backscatter - Transect 3. Data collected on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

The velocity data for Transect 3 shows variability in the velocity across the entire transect. The lowest velocity value is 0.008 ms-1 and the highest velocity value is 0.343 ms-1. This variability in speed will result in a large amount of shear throughout the water column, both vertically and horizontally, and is possible able to form small scale eddies in the water. This transect was taken 30 minutes before high tide and therefore the tide is likely to still be flooding in some areas leading to higher velocities. In others it may have already reached high tide and be at slack water and so the velocities are slower therefore resulting in variable velocities across the transect.

Once again backscatter data for Transect 3 is relatively low across the transect with values ranging from 63 db to 71 db. This low coincides with the Richardson number data which for depths 8 m to 9 m is greatly above 0.25 meaning the flow is laminar and stable. Similar to the other two stations zooplankton numbers collected at this point were also low so there are fewer particles in the water backscattering the sound so producing a smaller signal.

Transect 4 [Figure 5.7a and Figure 5.7b]

(a)    (b) 

Figure 5.7 - ADCP Estuary Data collected for Transect 4 - Falmouth, UK - (a) Velocity magnitude -Transect 4, (b) Average backscatter - Transect 4. Data collected on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

From the velocity magnitude figure it can be seen that the velocity either side of the main channel is slower with velocities of 0.012 ms-1 to 0.235 ms-1 whereas directly in the middle of the channel there is an area of very high velocities from the surface down to 11 m depth. These high velocities range from 0.409 ms-1 to 2.185 ms-1 in this middle channel section.  The reason for this higher velocity is because the central channel is deeper there is a greater volume of water here which results in higher water velocities.

Average backscatter for Transect 4 is high at the surface with values of 90 db to 98 db and is low from 3 m down to the bottom of the central channel with values of 68 db to 70 db. Because we did not do a net sample for this transect we cannot see whether the backscatter signifies high zooplankton numbers so the other explanation for high backscatter at the surface is due to turbulence created by the wind and the waves resulting in the sound scattered in all directions.

Richardson Number (Ri) [Figure 5.8]

Figure 5.8 - ADCP Estuary Data with Richardson numbers (Ri) calculated for Station 1, Station 2, and Station 3. Data collected on 30 June 2012. Graph can be enlarged in a new window upon clicking on the relevant image.

All 3 stations in the estuary show variable Richardson numbers with depth, indicating that the flow changes from being laminar to turbulent, and vice versa at different depths and at different positions in the estuary. The Richardson number for Station 1 is above 0.25 at all depths, indicating a laminar flow at this station. This is supported by the low velocity magnitude seen on the ADCP data. The majority of the flow at Station 2 resulted in a Richardson number above 0.25, also signifying laminar flow. However, for the depths of 3.2 m to 3.8 m, 5.1 m to 5.8 m, 8.8 m to 9.2 m and below 11.4 m the Richardson number is below 0.25. This corresponds to turbulent flow, which is due to the incoming flood tide. This increases the velocity as the water flows from the sea and outer estuary into the upper estuary towards high water.

Between 1.3 m and 2.7 m at Station 3, the Richardson number was below 0.25, which indicates turbulent and unstable flow. Whilst at this station, surface wind rows and Langmuir cells were observed. These occurred as a result of the surface wind stresses present, creating surface shear at this atmospheric-water boundary. This observation correlates with the turbulent flow calculated by the Richardson number, as the wind stress creates turbulence within the water column. Between depths of 8.0 m and 8.7 m at Station 3, the Richardson number was greater than 0.25, thus the water flow at this location was stable and laminar. This correlates to the timing of the tidal cycle at which the station was sampled. The station was sampled 30 minutes before high water, at which point flood tidal currents were beginning to reduce. The tidal cycle becomes a slack tide, reducing the velocity of the water and therefore the turbulence experienced in the water column.

Thermosalinograph [Figure 5.9a and Figure 5.9b]

Throughout the estuary survey, temperature and salinity values were recorded for point latitudes and longitudes against a known time. Surfer graphing software has been used to map temperature and salinity contour plots, with a post map overlay created to demonstrate a change in these variables with time (labels in red). The latitude and longitude is in decimal degrees as opposed to Northings and Eastings. Times are labelled in red and in UTC.

(a)    (b) 

Figure 5.9 - Thermosalinograph estuary data collected in Falmouth, UK - (a) Temperature and time contour plot of latitude against longitude and (b) Temperature and time contour plot of latitude against longitude. Recorded on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

The initial track taken up the estuary was followed by a slow track down, with intermittent stops for sampling against a flood tide. As can be seen from the 08:35 UTC time stamp, temperatures of 15.6 °C and salinities of 25.7 psu were recorded. Upon reaching further up the estuary (09:25 UTC), temperatures and salinities had decreased to 16.0 °C and 22.5 psu respectively. Following the track down the estuary seawards, temperatures and salinities increased as the incoming flood tide arrived. Significant changes in these variables occurred in the time period of 1 to 2 hours (12:00 to 13:30 UTC) preceding high tide at 13:51 UTC, when the tidal flow up the estuary was at its greatest.

Dissolved Oxygen

Figure 5.10 - Estuary dissolved oxygen saturation. Sampled on 30 June 2012. Graph can be enlarged in a new window upon clicking on the relevant image.

For Station 1, which is located at the beginning of estuary only has a depth of 2.9 m. Oxygen saturation increases towards bottom water column from 99.8%, which is the lowest oxygen saturation measured of water samples from all stations, to 101.8%. Samples from Station 2 show that oxygen saturation has an increase of 1.30% from surface to 5 m, and then decreases towards the bottom with a value of 1.18%. Overall, the oxygen saturation does not change greatly from the surface to the bottom. Oxygen saturation at Station 3 decreased by 6.62% from the surface to the bottom. Station 3 showed a much higher oxygen saturation than Station 1 and Station 2. The surface water sample had a dissolved oxygen percentage of 111.8%, the highest of the water samples from all the stations. Compared to the decrease through the vertical water column, the decrease of 0.37% from 2 m to 4 m is much more gradual. The overall trend of surface oxygen saturation showed an increase from the upper estuary at Station 1 to the lower Station 3, which coincides with the increase of phytoplankton abundance. The primary source of the dissolution of oxygen is the air-sea exchange with oxygen in the atmosphere. However, factors which affect the oxygen saturation level are mainly are biological processes, including photosynthesis, respiration and biochemical oxygen demand (BOD). The decrease of oxygen saturation from the surface to the bottom may be due to less surface interaction between water and air in deep sections and increased respiration from aquatic vegetation, microorganisms and algae and decreased photosynthesis. Increased phytoplankton abundance from Station 1 to Station 3 brings higher phytoplankton primary production, which causes the increase of oxygen saturation. Station 3 is much higher than Station 1 and Station 2. This may be due to the phytoplankton community at Station 3 being in the period of growth and reproduction, and the phytoplankton community at Station 1 and Station 2 being in the period of decay, which has a higher BOD.

Phosphate [Figure 5.11a and Figure 5.11b]

 

Figure 5.11 - (a) Estuary dissolved phosphate concentration and (b) Phosphate estuary mixing diagram. Sampled on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

On the graph of phosphate concentration against depth, it can be seen that Station 1 shows a low concentration of phosphate at 1 m which increases to 0.676 µmolL-1 by 2 m down. There is then a slight drop to 0.562 µmolL-1 by 3 m. This peak in concentration may be linked to the chlorophyll levels measured, which can be an indication of the phytoplankton present, as at this station the chlorophyll was quite low with a level of 3.4 and 3.9 µgL-1. This would potentially lead to a higher phosphate concentration if the nutrients are not being utilised by the phytoplankton, as they are not present in a large abundance.

Station 2 shows the opposite to Station 1, with a decrease in concentration with depth from 0.371 µmolL-1 at the surface to 0.217 µmolL-1 at 4 m, and then to 0.189 µmolL-1 at 6 m. From the processed chlorophyll data it can be seen that, at 4 m depth, a value of 2.0 µgL-1 was recorded, so there is low abundance of phytoplankton at this depth. The surface waters registered a higher value of 3.3 µgL-1, the opposite to phosphate concentration which was higher at the surface than at depth.

At Station 3, phosphate shows almost homogenous behaviour with a surface concentration of 0.189 µmolL-1, which remains fairly constant throughout the data series. This low concentration could be linked to a higher number of phytoplankton present in the estuary at this station; however, from the processed chlorophyll data it is shown that the values do not exceed 3.3 µgL-1. This shows that there was a low phytoplankton abundance, so the phosphate concentration was not affected.

As well as phosphate concentration being affected by the abundance of phytoplankton, sediments can play a role in controlling the amount of phosphate present in estuaries (Sundby, 1992) via the decomposition or dissolving of organic/inorganic material falling to the sediment surface. Although the main controlling factor of phosphate concentration in the estuary will be phytoplankton abundance, there are other factors that also have to be taken into account when studying data sets. From studying the phosphate concentration alongside the chlorophyll data, however, there does not appear to be much correlation between the amount of phosphate and the abundance of phytoplankton.

Phosphate - Estuarine Mixing Diagram [Figure 5.11b]

When plotting the estuarine mixing diagram for phosphate concentration against salinity in the Fal Estuary, a riverine end member and marine end member (Latitude: 50’ 12.553 N. Longitude: 005’ 01.673 W) were used to create the Theoretical Dilution Line (TDL). From this line, it is evident that the phosphate is acting non-conservatively and has many data points positioned well above the TDL, meaning excess phosphate is being added into the estuary and there are processes other than dilution occurring.

At the time of sampling, there had been rainfall for a few days prior to the study, and so this will have increased the amount of land surface runoff into the rivers and tributaries that flow into the estuary. When surface run off is increased, it leads to nutrients being added to the estuary waters and so increases the amount of nutrients found when studying the estuary. The lowest concentrations of phosphate were found nearer the mouth of the estuary where the salinity was greatest; this data was taken due to the marine end member being of quite high salinity at a value of 34. The riverine end members were taken from the Rivers Allen and Kenwyn, and the Boscawen Bridge (confluence of the Allen and Kenwyn Rivers). Finally one was obtained from the Truro River. These had salinities of 0.0 psu, 4.8 psu and 16.3 psu. From the data it is clear that there is addition of phosphate along the Fal Estuary, most likely due to surface runoff caused by the rainfall occurring before and during the data collection.

Nitrate [Figure 5.12a and Figure 5.12b]

 

Figure 5.12 - (a) Estuary nitrate concentration and (b) Nitrate estuary mixing diagram. Sampled on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

The estuarine mixing diagram plots dissolved nitrate against salinity in the estuary to determine whether nitrate in the estuary is mixing conservatively or non-conservatively (Loder & Reichard, 1981). The theoretical dilution line (TDL) is created using the riverine end member (salinity = 0 psu, nitrate concentration = 36.2  µmolL-1) and marine end member (salinity = 33.1 psu, nitrate concentration = 0.848 µmolL-1). If the data points plot on the TDL it is said that the nitrate is mixing conservatively and if it plots off the TDL then mixing is non-conservative and addition (above the line) or removal (below the line) has taken place.

The graph [Figure 5.12b] indicates that, through the majority of the estuary, the nitrate was mixing conservatively, with nitrate concentration decreasing as salinity increases.  Three points deviated sufficiently from the TDL to indicate there may be some non-conservative behaviour. This could have been caused by several water masses with differing nitrate concentrations or an external source (Loder & Reichard, 1981), which would cause addition of nitrate e.g. introduced through river water via surface run off. Higher precipitation levels usually cause a higher nutrient input. The concentration of nitrate in one of the river end member samples taken at the surface and at 1 m at Station 1 (the furthest point reached up the estuary) was below the TDL enough to indicate removal. The cause of this removal is likely to be phytoplankton, which incorporates it into their skeleton through biological fixation.

Figure 5.13 - Estuary nitrate and chlorophyll concentration. Sampled on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

The vertical profile [Figure 5.13] compares the nitrate concentration (µmolL-1) and chlorophyll concentration (µgL-1) changes with depth at each station. The samples from Station 1, located furthest upstream with the lowest salinity values, contained the highest nitrate and chlorophyll concentrations overall.  The lowest nitrate concentrations were found at 1 m, which coincides with the peak in chlorophyll concentration of 5.59 µgL-1, also at 1 m. As chlorophyll is a proxy for phytoplankton presence, this indicates that the phytoplankton are removing the nitrate from the water column. A similar pattern can be seen at Station 3, although it is not as pronounced. At Station 2, however, the chlorophyll concentration profile in the top 4 m of water matches the nitrate profile, indicating that there is an alternative factor that is limiting phytoplankton growth.

The concentration of nitrate is not affected by depth, but is affected by the presence of chlorophyll. The overall nitrate concentration decreases at each station, as the salinity increases from Station 1 to Station 3.

Dissolved Silicon [Figure 5.14a and Figure 5.14b]

(a)    (b) 

Figure 5.14 - (a) Estuary dissolved silicon concentration and (b) dissolved silicon estuary mixing diagram. Sampled on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

Calibration Curve

All data points of the calibration curve plot show a very strong linear relationship with an R2 value of 0.9996.

Theoretical Dilution Line (TDL)

Generally, the concentration of dissolved silicon decreased from the riverine end member to the sea end member as salinity increased. The difference between the two end members was 92.37 µmolL-1. The dissolved silicon concentrations conformed well with the theoretical dilution line, although further up the estuary the concentration was slightly lower by 7.7 µmolL-1. The dissolved silicon in the estuary varied fairly proportionally with salinity, demonstrating conservative behaviour.

Depth Profile

At Station 1 (furthest up the estuary) and Station 2, dissolved silicon concentrations decreased rapidly with depth between 0 m and 4 m. Between the depths of 4 m and 6 m, Station 2 showed a more gradual decrease and at 6 m dissolved silicon became uniform with increasing depth. At Station 3 (furthest down the estuary) dissolved silicon varied very little with depth. There was a slight increase in concentration of  0.9 µmolL-1 between the depths of 2 m and 4 m, beyond which there was a slight decrease. Dissolved silicon was much higher at the surface at Station 1, with a value of 28.80 µmolL-1. Although dissolved silicon concentrations at Station 1 were higher at each depth, Station 1 showed the greatest decrease with depth, with the lowest value of 15.94 µmolL-1 at 5 m. Station 2 had a lower surface concentration (13.81 µmolL-1) than Station 1 but showed a less drastic decrease in concentration with depth (over the same depth range) of 9.68 µmolL-1. Station 3 had the lowest surface concentration of dissolved silicon (6.50 µmolL-1). However the dissolved silicon concentration was slightly higher at Station 1 (6.91 µmolL-1) than Station 2 (4.76 µmolL-1) at a depth of 6 m and remains higher at depth beyond 6 m.

Chlorophyll [Figure 5.15]

Figure 5.15 - Estuary chlorophyll concentration. Sampled on 30 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

For the surface water, chemical chlorophyll analysis conducted from water samples in the lab shows a trend from highest chlorophyll concentrations (> 5.5 µgL-1) in the uppermost estuary at Station 1, to  lower concentration at Station 2 (< 3.5 µgL-1) [Figure 5.13]). The concentrations continue to decrease to a minimum at Station 3 (< 3 µgL-1). Overall, Station 1 has higher chlorophyll concentrations than other stations from the surface to the depth of 2.5 m. Station 2 shows a clear decrease of chlorophyll concentration from the surface to the bottom (< 1.5 µgL-1). Station 3 has lower concentrations than Station 2 in the upper water column (0 m to 2.4 m) and higher concentrations than Station 2 in the lower water column (2.4 m to 8 m). All samples are taken from low tide at the top of the estuary. 

(a)    (b)    (c) 

Figure 5.16 - (a) Estuary R.V. Conway, (b) Chemical sampling aboard R.V. Conway and (c) Chemical sampling aboard R.V. Conway. Sampled on 30 June 2012. All pictures can be enlarged in a new window upon clicking on the relevant image.

 

Conclusion

From the obtained data, phosphate showed non-conservative behaviour, with addition occurring at Station 1. Overall the concentrations decreased with increasing depth, however there was little correlation when compared to chlorophyll levels at similar depths.

Nitrate displayed conservative behaviour with slight non-conservative behaviour at the marine end member. It did not vary with depth but showed a correlation with chlorophyll levels. The nitrate concentration also decreased at each station as the salinity increased with greater proximity to the estuary mouth.

The silicon concentrations appeared to decrease as the depth decreased and also showed a positive correlation with chlorophyll. Generally conservative behaviour was observed. However, there was slight removal at the riverine end which may be linked to the higher chlorophyll levels witnessed at this station.

 

Return to Top of Page

 

6. Offshore and Estuary Biological Analysis                                        Return to Top of Page

Introduction

Analysis of the biological content of the water column for offshore and estuary-based practicals were combined to note how the physical water properties affect the biological activity.

 

 Figure 6.1 - GoogleMap images of the location of the plankton net trawls and sampling stations for the offshore and estuary practicals.

Common plankton typical in south-west UK waters:

(a)    (b)    (c) 

(d)    (e) 

Figure 6.2 - (a) Copepod larvae, (b) Decapod larvae, (c) Hydromedusae larvae, (d) Phytoplankton species (named on figure) and (e) Phytoplankton species (named on figure).

 

Data Analysis

Phytoplankton Offshore

Figure 6.3 - Phytoplankton abundance offshore - Station 1, Station 2 and Station 4. Sampled on 27 June 2012. Graph can be enlarged in a new window upon clicking on the relevant image.

For Station 1 (Black Rock), samples were collected from three depths (1.2 m, 6.3 m, and 27.1 m). The 27.1 m water sample was entirely dominated by diatoms [Figure 6.4c]. Water samples taken at 6.3 m and 1.2 m were also dominated by diatoms and contained some dinoflagellates [Figure 6.4a and Figure 6.4b]. Dominant species are Guinardia flaccida, Eucampia, Rhizosolenia delicatula and Chaetoceros. Guinardia flaccida were found throughout this station with high density. The surface waters show higher abundance of phytoplankton than the water column below.

At Station 2, water samples were taken from 0.8 m, 7.0 m, 27.0 m and 60.4 m depths. Here diatoms are highly dominant [Figure 4.6d, Figure 4.6e, Figure 4.6f, Figure 4.6h], and the main species found are Guinardia flaccida, Guinardia striata and Rhizosolenia delicatula. The ciliate Mesodinium is found at the 60.4 m depth water sample at a low abundance. The water sample taken from 47.9 m which has the highest abundance of all water samples is dominated by dinoflagellates, of which the main species are Dinophysis and Polykrikos. Dinophysis has the highest density (5×10⁷ cells per mł of sea water). Chaetoceros is the main diatom found in this water sample.

At Station 4, water samples taken from all depths (1.3m [Figure 6.4j], 13.1 m [Figure 6.4i], 55.0 m [Figure 6.4k], and 68.2m [Figure 6.4l]) are almost entirely dominated by diatoms, however Alexandrium [Figure 6.4i] and Mesoporos perforatus [Figure 6.4k] are also found at low abundances. The main species at this station are Leptocylindrus danicus, Guinardia flaccida, Guinardia delicatula and Rhizosolenia stolterfothii. At 13.1 m depth at Station 4, Leptocylindrus danicus is present in such high abundance that it is the most abundant species found at every depth of every station. The abundance of phytoplankton increases towards the upper layer of water column.

In all, all stations are highly dominated by diatoms. The main species like Guinardia flaccida, Rhizosolenia delicatula and Leptocylindrus danicus are found throughout all stations at significantly high abundances. Dinoflagellates were mainly found at low abundance, except within the water sample taken from 47.9 m depth at Station 2 which shows slightly greater abundance of dinoflagellates than diatoms. Dinophysis and Polykrikos are only found at this station but are present in superbly high abundance. Alexandrium are found throughout all stations with low abundance. The ciliate Mesodinium rubrum is only found in the water sample found at 60.4 m depth from Station 2. Station 2 is the only station where diatoms, dinoflagellates and ciliates are all represented. It also has the highest species diversity of any station. Because Station 1 is in the mouth of the estuary, it is shallower than other stations and has an abundance peak of 1.08×10⁹ individuals per mł near the surface. Station 2 and Station 4 have abundance peaks at 47.9 m depth (1.34×10⁹ cells per mł of seawater) and 13.1 m depth (2.08×10⁹ cells per mł of sea water) respectively [Figure 6.3]. Overall, the abundance increases from inshore stations to offshore stations from surface to the depth of 30 m. Specifically, for the surface waters, Station 1 has the highest abundance and Station 2 has the lowest abundance which coincides with the same trend shown by zooplankton. This may be because Station 1 has the highest surface irradiance and Station 2 has the lowest surface irradiance. For the water column below 30 m depth, phytoplankton abundance decreases with increased depth because of decreased irradiance. The deepest samples taken have phytoplankton abundances between 2×10⁸ and 3×10⁸ cells per mł of seawater.

Factors such as light, nutrients and zooplankton abundance are crucial in determining where phytoplankton are abundant and to what degree. Water motion is in many ways the dominant factor which affects the exchange of nutrients and also drives the mixing of waters and entrainment of phytoplankton downward. The very high abundance of phytoplankton observed at 13.1 m depth at Station 4 may be due to the convergence of water by the creation of vortices from wind-driven water movement. The relatively high abundance at 47.9 m depth which coincides with the chlorophyll peak, may be due to the mixing of two water masses: inshore and offshore. The water mass from offshore has a higher abundance which takes phytoplankton down to deeper waters.

 

(a)    (b)    (c)    (d) 

(e)    (f)    (g)    (h) 

(i)    (j)    (k)    (l) 

 Figure 6.4 - Phytoplankton cells per m3 collected over a range of depths offshore - (a) Station 1 (1.164 m), (b) Station 1 (6.301 m), (c) Station 1 (27.063 m), (d) Station 2 (0.804 m), (e) Station 2 (6.984 m), (f) Station 2 (26.983 m), (g) Station 2 (47.897 m), (h) Station (60.253 m), (i) Station 4 (13.124 m), (j) Station 4 (1.342 m), (k) Station 4 (54.993 m), and (l) Station 4 (68.186 m). Sampled on 27 June 2012. All graphs can be enlarged in a new window upon clicking on the relevant image.

Zooplankton Offshore

Figure 6.5 - Multiple pie charts showing offshore zooplankton cell counts per m3 at varying depths. Sampled on 27 June 2012. Graph can be enlarged in a new window upon clicking on the relevant image.

The greatest abundance is observed at Station 1, where samples were taken from 8 m to 2 m, which is near the surface.  Zooplankton is dominated by Copepoda, Decapoda Larvae, Hydromedusae and Appendicularia with Copepoda nauplii present at low abundance. Eleven species are found here and high dominance is observed, especially exhibited by Copepoda which count for over 50% of the total abundance.

Three water samples were collected from Station 2. Figure 6.5 shows the vertical zonation of zooplankton. The deepest water sample is from 48 m to 40 m, which is dominated by Copepoda and Copepoda Nauplii. Only 4 species are found in low abundances here. A shallower water sample taken from 40 m to 20 m shows a largely increased abundances of Copepoda which has a massive dominance. Eight species are found and Cladocera was only found in any significant abundance here of all water samples. The shallowest sample is taken from 18 m to 7 m, and Copepoda Nauplii and Appendicularia are present in greater numbers compared to the deeper water column while Copepoda, Decapoda Larvae and Siphonophorae have similar abundances, although Copepoda is dominant. In all, the deepest area has a much lower abundances and fewer species compared to the area above.

A water sample collected at Station 4 was taken from 21 to 7 m. Copepoda and Copepoda Nauplii show high dominance at this location and Copepoda Nauplii had a dramatically greater density than Copepoda. Ten species were found at Station 4 and this was the only station that Gastropod were found. Station 4 has a slightly lower total abundance compared to Station 1 and a much higher density compared to Station 2.

 

Zooplankton in Estuary

Unlike zooplankton samples collected offshore, zooplankton samples collected on the estuary were collected by dragging a plankton net horizontally across the surface of the estuary at various stations. Therefore the numbers of rotations recorded on the flow meter were implemented in determining the distance the plankton net was dragged. This distance used in conjunction with the area of the plankton net opening were used to calculate the total volume of water that the plankton net filtered.

Station 1 - plankton net dragged for 13.5m. Copepoda and copepoda nauplii are dominant here. They each make up approximately 40% of the total zooplankton numbers found at this location, with combined numbers making up approximately 80%. This 80% equates to approximately 4.4x105 cells per m-ł of estuary water. Other species present are decapoda larvae, polychaeta larvae, hydromedusae and cirripedia larvae. Polychaeta larvae and cirripedia larvae are abundant to the same degree in this location, making up a low percentage of the total zooplankton count (less than 1% each). Hydromedusae and decapoda larvae are also present in the same quantities. Hydromedusae make up approximately 6% of the individual zooplankton counted and decapoda larvae account for roughly 4%. These much smaller values correlate to far fewer cells of these species present at Station 1 compares with the much more abundant copepod and copepod nauplii.

Station 2 - plankton net dragged for 275.1m. Unlike data collected from Station 1, there is an overwhelming abundance of hydromedusae here. Accounting for over 60% of the individual zooplankton counted at Station 2, around 1.6x104 individual jellyfish are found in every m-3 of estuary water. Polychaeta larvae have decreased in total population compared to Station 1 but they too make up a greater percentage of the total zooplankton numbers here than at Station 1. Copepoda and copepoda nauplii continue to make up a considerable portion of the total zooplankton numbers however it is significantly less than it was at Station 1. Cirripedia larvae and decapoda larvae are still present but make up a small percentage of the total zooplankton. While dominating Station 2, there are still more hydromedusae at Station 1 (3.3x104 per m-ł of estuary water). This suggests that rather than an increase in hydromedusae numbers, there is instead a larger decrease in other zooplankton species numbers.

Station 3 - plankton net dragged for 183.9 m. Here there continues to be a greater number of hydromedusae than other zooplankton species, contributing 45% of the individuals present. This correlates to there being 1.6x104 jellyfish per mŻł of estuary water, the same number of individual jellyfish found at Station 2. While the individual numbers remain the same, percentage has fallen, showing that there has been an increase in copepoda, copepoda nauplii, cirripedia larvae and decapoda larvae. While decapoda larvae numbers have slightly increased, copepoda nauplii numbers have more than tripled and for the first time have more individuals per cubic metre of estuary water than copepoda.

Station 1 estuary water contains the greatest number of individual zooplankton, whereas Station 2 contains the fewest individual zooplankton. While copepoda and copepoda nauplii are dominant at Station 1, all species numbers decrease at Station 2 and Station 3 and hydromedusae become dominant. This appears to mainly be due to a larger decrease in other zooplankton species numbers compared with the decrease in hydromedusae numbers.

 

Combined Offshore and Estuary Biological Interpretation

Phytoplankton

Six phytoplankton samples were collected between a mooring located a short distance downstream of Malpas (estuary Station 1), and ‘offshore Station 4’ which was located 15 miles offshore to the south.  The furthest sampling area up the estuary was Station 1 of the estuary. Here the lowest phytoplankton abundance was found and it measured at only 1.85 ×10⁸ cells per mł of estuary water. This may be due to flushing of phytoplankton caused by tidal ebb. At ‘estuary Station 2’ further downstream the total phytoplankton abundance had increased to 4.1 ×10⁸ cells per mł of estuary water and dinoflagellates became dominant. This suggests a more highly stratified water column and a lower nutrient concentration than present at ‘estuary Station 1’, as dinoflagellates prefer a more stratified environment and are more highly adapted to lower nutrient environments. This is supported by evidence of lower nitrate concentration at ‘estuary Station 2’ than at ‘estuary Station 1’. The highest total phytoplankton abundance was found at ‘offshore Station 1’, where 1.08×10⁹ cells were found per mł of sea water. This is much higher than total phytoplankton abundance found at any other sampling station suggesting greater concentrations of nutrients. The convergence of the Penryn River, Percuil River and Fal River could explain the higher nutrient concentrations that are exhibited by nitrate, phosphate and silicate data collected at all stations.  ‘Offshore station 4’ has the second highest total phytoplankton abundance and diatoms are again dominant with no dinoflagellates present suggesting a well mixed water column. From ‘estuary Station 3’, through ‘offshore Station 1’ and to ‘offshore station 2’, dinoflagellates are present although diatoms are dominant. This shows that the water column at these stations is not stratified but is not as well mixed as the water column at ‘estuary Station 1’ or ‘offshore Station 4’. Where dissolved silicon concentration increases between 2 m to 8 m depth at ‘estuary Station 3’, there is also a strong peak in chlorophyll concentration, suggesting phytoplankton growth is being limited by dissolved silicon concentrations. This trend is exhibited at all estuarine stations. The very high phytoplankton abundance at ‘offshore Station1’ could be explained by the high dissolved silicon concentration also found in the surface waters here, possibly suggesting that dissolved silicon is no longer the limiting factor and the Penryn River, Percuil River are supplying adequate quantities of dissolved silicon to support greater phytoplankton growth. Decreases in phosphate and nitrate concentrations at ‘offshore Station 1’ also suggest that they are the new limiting factors of phytoplankton growth.

While chlorophyll a concentrations are currently the best technique for indexing phytoplankton numbers (Huot et al., 2007), chlorophyll concentrations taken from each of these stations do not correspond very well with the phytoplankton counts. This may be due to differences in the levels of chlorophyll within the cells of the phytoplankton species present. For example, the ceratium furca, karenia mikimotoi and alexandrium species which are abundant to a higher degree at ‘offshore Station 1’ compared to other offshore stations, may contain smaller quantities of chlorophyll within their cells than diatom species found. This could explain why the sample taken at ‘offshore Station 1’ has a lower than expected chlorophyll concentration, and the sample taken at ‘offshore Station 4’ has a higher than expected chlorophyll concentration.

Zooplankton

Throughout the transect leading from ‘estuary Station 2’ to ‘offshore Station 4’ the trend in zooplankton abundance matches the trend in phytoplankton abundance very well. Where phytoplankton abundance is higher, as expected the zooplankton abundance is also higher and vice versa. ‘Offshore Station 1’ which exhibited the greatest phytoplankton numbers also shows the greatest zooplankton numbers (approximately 6.8×10⁶ individuals per mł of sea water). This is due to phytoplankton being the primary diet of many zooplankton species. Therefore at the locations where there are plentiful food sources zooplankton’ thrive. However, if the zooplankton correlated with phytoplankton at every station then the lowest zooplankton abundance would be expected to occur at ‘estuary Station 1’. This is not the case as there are approximately 0.7×10⁶ cells per mł of estuary water. This divergence from the expected trend can be attributed to a cycle of dominance by phytoplankton production and then by grazing zooplankton (Levinton., 2008). At ‘estuary Station 1’ grazing zooplankton are dominant whereas at every other station phytoplankton production is dominant.

(a)    (b)    (c)    (d) 

Figure 6.6 - (a) Surface changes of phytoplankton and zooplankton at all Stations offshore 27th June 2012, (b) Station 1 - Phytoplankton collected from depth of 0.5m in estuary 30th June 2012, (c) Station 2 - Phytoplankton collected from depth of 0.5m in estuary 30th June 2012, (d) Station 3 - Phytoplankton collected from depth of 0.0m in estuary 30th June 2012

Estuary discussion phytoplankton

Only one water sample from each station was collected with all the water samples collected near the surface. The water sample from Station 1 is entirely dominated by diatoms [Figure 6.6b] and the main species is Eucampia sp. which nearly accounts for 75% of total abundance. The water sample from Station 2 is dominated by dinoflagellates which account for over 70% of the total abundance [Figure 6.6c]. The main species are Alexandrium and Karenia mikimotoi. The dominant specie of diatom is Guinardia flaccida. The water sample from Station 3 shows that diatoms are highly dominant [Figure 6.6d]. The main species present are Guinardia flaccida and Rhizosolenia stolterfothii. Alexandrium is the only species of dinoflagellate found in this water sample and it has a significantly high abundance (1.2 ×10⁸ cells per mł of estuary water). The diatom Guinardia flaccida is found at all three stations with high abundance. The main dinoflagellate is Alexandrium. Station 2 and Station 3 which are closer to the mouth of the estuary show much higher total phytoplankton abundance and greater species diversity than Station 1.

 

Return to Top of Page

 

7. Geophysics                                                                                                  Return to Top of Page

Introduction

The vessel Xplorer was used on the 3rd July 2012 for surveying the Falmouth Harbour. The survey focused on an area of the channel that is a proposed site for dredging in 2013. This channel is to allow cruise ships and bigger vessels to reach port in order to increase tourism in Falmouth, which will hopefully have an advantage to businesses and industry in surround areas.


View Falmouth Geophysics in a larger map

Figure 7.1 - Geophysics locations of sampled  sidescan transect lines and drop camera transects/station.

 

Figure 7.2 - Proposed dredging site shaded pink, (Falmouth Cruise Project, 2012).

When surveying the area the biodiversity of the proposed dredging area will be studied to look at potential impacts on the organisms living in the area. The plans are controversial in the fact that the area is a habitat for the protected maerl which can be easily damaged due to the impact of trawling and pollution. Also maerl does not always return once lost (Natural England). From Environmental Impact Assessments carried out by companies such as Royal Haskoning (Link to EIA) the advantages and disadvantages are being studied, along with surveys, such as the one being done on Xplorer. This is to build up a better picture of the consequences dredging will have.

Suggestions have been made to relocate the maerl (Solly and Knowles, 2009) to areas where maerl is already known to thrive, such as further up the estuary to the North East of the dredging site. Although this would not be a definite success as it could take up to 50 years for a community to establish itself after the relocation (Marine Management Organisation, 2010) as long as no siltation occurs which could inhibit the growth.

In order to survey the area two techniques will be used, one is the Tow Fish Sidescan which maps an area of the seafloor up to 150 metres and gives a picture of the contours and bedforms present in the channel. Secondly a Video Camera that captures images of the surveyed area will be used to build a visual picture of the bed of the harbour and identify all the organisms that make this area of the estuary their home. By having different methods of analysing the harbour a data set can be built up that gives a comprehensive overview of whether the site is suitable for dredging to take place or not as well as improving scientist's knowledge of the range of species found in this area.

Geophysical Sidescan

Using the sidescan trace produced whilst out on the boat, the start and end time of each of the 3 transects carried out were highlighted. From these points the sidescan trace could be separated into each of the transects and then overlapped on top of each other to produce an overall swath width of 100m. To determine where the transects would be overlapped the Eastings of each of the transects were lined up to form a sidescan mosaic. On this mosaic 5 boundaries were mapped relating to areas of changes in colour signifying differences in the composition of the sea bed. Points marking an accurate depiction of the outline of the boundary shape were added and the times at which these points occurred determined. Once the boundaries were mapped several parameters could be calculated. These parameters were:

Figure 7.3 - Geophysics definitions.

·         Calculating the fish height from the paper trace:

(a) 

·         Calculating slant range from the paper trace:

(b) 

·         Calculating the horizontal range:

(c) 

 Figure 7.4 - Geophysics formulae (a) Calculating the fish height (m), (b) Calculating the slant range,  and (c) Calculating the horizontal true range (m).

 

Figure 7.5 - Diagram of Geophysics calculations.

Using the times that each boundary point occurred at and the horizontal range calculated for each point the boundaries could then be mapped onto the transect plot created in surfer using the scale bar and converting the distances to a size that will accurately depict them on this track plot.

Plotting the Sidescan Track Plot

The raw navigation data collected on Hydropro during the boat survey was imported into an excel spreadsheet to separate the variables into time, Eastings (m), Northings (m), Depth and Survey Line Number. From this a Post map was created in Surfer. Two post maps were created one of the position and the second for the time with Eastings (m) on the x-axis and Northings on the y-axis. A north arrow was also added to give an accurate representation of the direction. A scale bar was added to the plot which could be used later to convert the boundaries to a size that would fit on the plot whilst still accurately depicting its size, the Rh(m) was converted to Rh(mm) according to the scale. The timelines on the Post map and Rh (mm) were used to determine the various boundaries defining points previously selected from the track plot, indicating changes in seabed composition or topography. Each boundary point and connecting line was added using the symbol and line features on Surfer 8 (possibly improving the reliability of the final plot). Each boundary was labelled as a particular zone and important features were also indicated.  The key was produced on Surfer 8, and colour was added by hand later.

Equipment

Geophysics using vessel Xplorer

The geophysics vessel Xplorer was used for this practical.

Tow Fish Sidescan

GeoAcoustics limited (159D) 100 KHz model. Sends beams of sonar  out and the backscatter forms an image of the sea bed showing any present bedforms or artefacts. The returning data is printed off and the colour of the image shows whether there is coarse or fine sediment in a particular area being surveyed.

Underwater Video Camera (Drop Camera)

Underwater Video Camera and two lights position either side on a hand-made structure created by Dr John Davis. The lights were at daylight level and temperature in Kelvin. It was deployed off the side of the boat and manually lowered/raised to get the right depth. Video was streamed in real time to watch on a monitor then recorded onto a CD-ROM disk to be analysed at a later date.

(a)    (b) 

Figure 7.6 - Geophysics equipment (a) Sidescan tow fish and (b) Drop camera.

 

Data Analysis

As part of the required assessment for this fieldcourse, a geophysics poster write-up was required. This was produced in the form of a Microsoft PowerPoint slide. Please open in the link in Microsoft PowerPoint: Link to Poster or in a Adobe (.pdf) format: Link to Poster.

Geophysical Analysis

Each of the boundaries delineated on the side scan were chosen because they represented a change in tonal colour (different intensities of reflections) which often relates to a change in the lithology of the seabed or change in grain size.  Coarser particles produce higher backscatter as there are more surfaces exposed and so a darker (dark grey - black) colour is produced on the side scan trace. Finer particles produce a lighter (pale grey-white) colour on the trace as they reflect lower backscatter from their surface.

   

Figure 7.7- Sidescan sonar trace of Line 1, Line 2, and Line 3. (a) Sidescan sonar traces assembled by aligning eastings, (b) marking boundary areas identified, and (c) marking out distances of boundary/artefact from tow fish centre line. Data obtained on 03 July 2012. All photos can be enlarged in a new window upon clicking on the relevant image.

Boundary 1 contains patches of high backscatter related to coarser particles interspersed with low backscatter from finer particles. From the video analysis the seabed was composed of fine sand with bivalve shells spread in some areas. These bivalve shells could be responsible for the high backscatter seen in Boundary 1. Boundary 2 is an area of very light colouring due to low backscatter from finer sediment (fine sand or coarse silt) and also probably an area of low number of epifauna. Boundary 3 is also a boundary of low backscatter as seen by its light colour which could also mean a seabed composition of fine sediment. Boundary 4 is similar to boundary 1 with a white colour interspersed with dark clusters. So the wide scale area of the seabed is composed of fine particles but with some coarser particles and discarded organism shells on the top. Boundary 5 is a relatively dark coloured boundary with intermingled patches of lighter colour and is positioned directly next to the bedforms in transect 1. This shows a clear change in the composition and the hydrodynamic conditions responsible for these from one section of the seabed to the next. The colouring of boundary 5 indicates coarser sediment producing the higher backscatter with a few areas of finer sediment.

Due to licencing issues, no grab samples were obtained to verify the exact lithology of the seabed along the transects/lines and so the assumptions made are based only on backscattering properties of the different sized sediment particles.

The side scan trace for both Transect Line 1 (01822424.01 m E, 0033173.54 m N) and Transect 2 (0182195.58 m E, 0033273.53 m N) shows areas on the seabed where subaqueous bedforms are present. This means that the sediment here is motile and is primarily composed of sand sized sediments (estimated from the dark grey colour on the trace possibly corresponding to particles of a medium size). However as we did not take a grab sample the exact grain size cannot be determined. The presence of these bedforms was confirmed when the video camera was deployed and bedforms could be seen. There is also some evidence on the side scan trace of bifurcation of the bedforms.  Other interesting features on our side scan are an arc shaped white area at 0181743.15 m E and 00339196.75 m N which possibly depicts the shadow zone of a raised feature on the seabed. There is also evidence of anchor marks in both Transect Line 1 and Transect Line 3 which can be identified by a line of black dots. This makes sense as we were surveying in the entrance area to Falmouth Harbour.

 

     

Figure 7.8 - (a) Xplorer geophysics survey vessel, (b) Paper printed sidescan trace, (c) Live computer visulisation of geophysics sidescan trace and (d) Planned transects/lines for sidescan tow fish to be deployed over. Data obtained on 03 July 2012. All photos can be enlarged in a new window upon clicking on the relevant image.

 

Video analysis

There was high turbidity in the area in which the video transects took place, resulting in low visibility on some parts of the footage. The height of the camera above the seabed varied with the changing water depth, and by physically lowering or raising the camera to the depth desired by the observers to obtain a close up or broad-scale view.It was difficult to analyse the sediment structure purely from the video footage. Ideally this data would have been used in parallel with analysis of grab data. The transects were, however, within a Special Area of Conservation (SAC), designated under the EC Habitat Directive, so grabs were prohibited. The SAC is to protect the growth of maerl in the area, visible on the video as pink nodules on the sediment surface. The housing of the video camera was red. Reflection of light off of the housing resulted in the image appearing redder than it actually was.

On Transects 1 and 2, many mollusc shells were observed on the surface of the sediment. Mollusc shells were also observed in the troughs between bedforms in the sandy sediment. It was observed in small clumps of a variety of species on the mixed sediment containing maerl, and was not observed on sandy sediment.

Transect

Time (UTC)

Start

End

 

Start

End

Lattitude

Longitude

Lattitude

Longitude

1

09:37

10:02

50° 09.456' N

005° 03.064' W

50°09.076' N

005°02.853' W

2

10:14

10:41

50° 09.430' N

005° 03.491' W

50°09.674' N

005°03.208' W

3 (Vertical Pile)

10: 52 

11:01

50°09.419' N  

005° 03.210' W

 N/A

Figure 7.9 - Tabulated positions and times of transects for geophysics survey on 03 July 2012.

Transect 1

Algal cover ranged from 0-25%. A mixed, maerl-based sediment was observed for the first 16 minutes of the video transect. Sandy sediment was then observed for the remaining 9 minutes.

 Flora

o   Laminaria digitata (Kelp)

o   Laminaria saccharina (Sugar kelp)

o   Dilsea carnosa

o   Ulva sp.

Fauna observed on Transect 1

Latin name

English name

Abundance

Morthasterias glacialis

Spiny starfish

6

Liocarcinus depurator

Harbour crab

4

Raja clavota

Thornback Ray

1

Raja undulata

Undulate Ray

1

Myxicola infundibulum

Fanworm

1

Sabella pavonina

Peacock worm

6

Gobiusculus flavescens

Two-spotted goby

2

Pecten maximus

Giant Scallop

1

Spirobis spirobis

Polychaete worm with white, coiled shell. Attached to kelp.

3

Lophius piscatorius

Angler fish

1

Unknown

Lobster

1

Unknown

Sponge

2

Unknown

Clam

1

Unknown

Juvenile Fish

31

 Figure 7.10 - Tabulated drop camera results from geophysics Transect 1. Geophysics survey on 03 July 2012.

Transect 2

Algal cover ranged from 0-75%. A mixed, maerl-based sediment was observed for the first 9 minutes of the video transect. Sandy sediment was then observed for the last 18 minutes. The openings of burrows were observed on the surface of the sandy sediment, belonging to an unknown organism.

Flora

o   Laminaria digitata (Kelp)

o   Laminaria saccharina (Sugar kelp)

o   Dilsea carnosa

o   Ulva sp.

Fauna Observed on Transect 2

Latin name

English name

Abundance

Morthasterias glacialis

Spiny starfish

1

Asterias rubens

Common starfish

1

Liocarcinus depurator

Harbour crab:

Crab fishing pot observed

9

Raja clavota

Thornback Ray

1

Myxicola infundibulum

Fan worm

84

Sabella pavonina

Peacock worm:

observed retreating into their tubes as the camera moved over them.

37

Gobiusculus flavescens

Two-spotted goby

1

Unknown

Anemone

2

Unknown

Sponge

17

Unknown

Clam

1

Unknown

Juvenile Fish

26

 Figure 7.11 - Tabulated drop camera results from geophysics Transect 1. Geophysics survey on 03 July 2012.

 

(a)    (b)  

(c)    (d)    (e) 

Figure 7.12 - Geophysics drop camera stills with species identified - (a) Sabella pavonina - Peacock worm, (b) Morthasterias glacialis -Spiny starfish, (c) Cereus pedunculatus - Daisy anemone, (d) Juvenile fish, showing that the maerl beds act as a fish nursery, and (e) Raja undulata - Undulate Ray. Geophysics survey on 03 July 2012.

 

Figure 7.13  - Kelp above the low water mark. Geophysics survey on 03 July 2012.

Figure 7.14 - 100% algal cover on harbour wall. Geophysics survey on 03 July 2012.

Figure 7.15 - Dead-man’s fingers. Geophysics survey on 03 July 2012.

Figure 7.16 - Kelp at the base of the harbour wall. Geophysics survey on 03 July 2012.

Figure 7.17 - Seabed below harbour wall. Geophysics survey on 03 July 2012.

 

 

Transect Line 3: Vertical Transect Down Falmouth Harbour Wall

Species observed in order of depth observed at:

      Flora

o   Ulva lactuca (Sea lettuce)

o   Rhodymenia palmata (Dulse)

o   Corallina affinalis

o   Delessenia sanguinea

o   Dilsea carnosa

o   Laminaria digitata (Kelp)

o   Ulva sp.

      Fauna

o   Suberites domuncula (Sulphur sponge)

o   Gobiaus niger (Black goby)

o   Gobius paganellus (Rock goby)

o   Gibbula sp. (sea snail seen on kelp)

o   Cereus pedunculatus (Daisy anemone)

o   Monostroma grevillei (Spiny starfish)

o   Alcyonium digitatum (Dead-man’s fingers)

Above the surface of the water, Rhodymenia palmata (Dulse) and Ulva lactuca (Sea lettuce) were observed growing, together covering roughly 90% of the harbour wall. As low tide was at 11:09UTC, the video was recorded in the 10 minutes prior to low tide. This suggests that these species were living in the intertidal zone and are usually submersed.

 

 

Conclusion

From the survey carried out on Xplorer of the proposed dredging site inside the harbour it was shown that there is an abundance of organisms living in this habitat from anemones to Thornback Rays. There was also maerl present in the majority of areas we surveyed showing it is an important area for the species to thrive.

The sidescan data was used to view any bedforms present and also study the lithology of the sediment, whether it is coarse or fine, and this allows the type of organism present to be analysed as certain animals cannot thrive in the wrong sediments. For example burrowing species will be dominant in the finer sediments but will struggle to burrow in the hard, bivalve shell covered sediments found in the survey area. Bivalve shells can also indicate fishing has taken place which has been found to be helped by having maerl present in the area. 

This is an on-going issue involving different stakeholders, but if the decision is made to continue with the dredging then further surveys  post-dredging will need to be carried out, in order to assess the extent to which dredging has impacted on the harbour and the organism living there.

 

Figure 7.18 - Surfer produced habitat mapping and bedform geophysics map. Geophysics survey on 03 July 2012.

Return to Top of Page

 

8. Pontoon                                                                                                      Return to Top of Page

Introduction

The pontoon is located on the River Fal at 50° 12'58.42ť N, 005° 01'39.42ť W near the King Harry Ferry terminal, where the chain ferry crosses the river every 10 minutes. The purpose of collecting data from this site was to create a time series of data for the area using a number of parameters, to see how they varied during the tidal cycle. Data collection occurred on an ebb tide (although the wind was blowing upstream), with sampling beginning at 08:45 UTC and measurements were taken every 15 minutes at 1.0 m depth intervals to the seabed. This process was repeated 8 times, with data collection finishing at 11:33 UTC.

 

Figure 8.1 - GoogleEarth map showing location of the pontoon.

For data collection, 3 pieces of equipment were used: a YSI probe, a current meter and a light sensor. The YSI probe was used to collect data on depth, salinity, temperature, pH, O2 concentration and chlorophyll a concentration. The current meter was used to collect data on the current speed and direction relative to North. The measurements were taken as the current meter was lowered down through the water column and on the way back up to gain multiple measurements for each depth, allowing an average value to be taken. The light sensor measured the light attenuation through the water column. This had a dry surface sensor and an in-water sensor. First the measurements of both sensors were taken at the surface to show the offset between the two sensors - this could then be used to calibrate the light data. At each depth in the water, measurements were taken from both sensors so that the fluctuating amount of surface light could be taken into account.

Figure 8.2 - Photo of pontoon sampling. Pontoon sampling on 04 July 2012. All pictures can be enlarged in a new window upon clicking on the relevant image.

YSI 6600 V2 Probe 

The same version was used as on R.V. Bill Conway taking the same measurements. It was deployed manually off the side of the pontoon.

Current/Flow Meter

Measured the flow rate of the liquid and the direction of flow. The moving dial was supported by a frame and lowered manually into the water from the pontoon edge.

Light Meter

A receiver measured the amount of visible light penetrating to the given depth. It was attached to a frame then lowered into the water off the pontoon edge. A second receiver measured light in the air in order to calibrate the instruments and allow for any changes in the atmospheric conditions.

(a)    (b) 

Figure 8.3 - Pontoon sampling equipment used on 04 July 2012. (a) YSI 6600 V2 probe, and (b) Current/flow meter. All pictures can be enlarged in a new window upon clicking on the relevant image.

Data Analysis

 

Figure 8.4 - Tidal cycle for Falmouth on 4 July 2012.

     

   

Figure 8.5 - Falmouth pontoon depth against time with profile overlay (a) Temperature, (b) Salinity, (c) Chlorophyll a, (d) pH, (e) Oxygen Saturation, (f) Average flow direction, and (g) Average velocity. Sampled on 4 July 2012. All pictures can be enlarged in a new window upon clicking on the relevant image.

Temperature [Figure 8.5a]

Temperature noticeably increased in the upper part of the water column as the tide ebbed. At 8:45 UTC the temperature at 1.0 m depth is between 14°C and 15°C, this then continued to increase until the end of the time series where the temperature reached 15.5°C. The deeper parts of the water column remained at low temperatures of 14.5°C.

Salinity [Figure 8.5b]

The upper area of the water column, throughout the time series had a lower salinity (32) recorded than that in deeper waters (32-34). Salinity in the surface waters also decreased as the tide ebbed, from 32 to 30.

Chlorophyll a [Figure 8.5c]

The tide was ebbing over the tidal period studied. High chlorophyll levels of 6 µgL-1 and 7 µgL-1 were observed between 08:45 UTC and 09:45 UTC, this then decreased between 9:45 UTC and 9:50 UTC to 5 µgL-1. At 9:50 UTC there ws a notable increase in chlorophyll (7 µgL-1), between 1.5 m and 3.0 m. There was also another band of low chlorophyll at 10:45 UTC, which was followed by another patch of high chlorophyll at 11:25UTC towards the end of the sampling and close to low water at 12:00 UTC between 1.0 m - 2.0 m.

pH [Figure 8.5d]

Below 1.0 m pH remained relatively constant at 8.75pH. However between 10:50 UTC and 11:25 UTC, close to low water, pH increased throughout the whole water column, reaching 8.8 to 8.85pH. Along the time series the top meter of the water column experienced “blips” in pH levels. It showed increases up to 8.8pH at 9:25 UTC, 10:15 UTC and 11:05 UTC.

Oxygen Saturation [Figure 8.5e]

Over the time series the percentage of dissolved oxygen saturation in the water column increased as the estuary approached low water at 12:00. Between 8:45 UTC and 09:25 UTC the percentage of dissolved oxygen throughout the water column remained constant at 94%. This however began to increase at 09:30 UTC and continued to increase until the end of the data set at 11:25 UTC, where the percentage of dissolved oxygen saturation reached 98%.

Average Flow Direction [Figure 8.5f]

The current, according to the contour plot, was shown to travel right, moving from 190° to 170° as time progressed. This would suggest that the flow was moving past the pontoon to the right, which corresponds to the flow of the ebbing tide.

Average Velocity [Figure 8.5g]

The contour plot showed that at the beginning of the time series at 08:45 UTC, when the tide was ebbing, there was a faster current speed throughout the water column of 0.4 ms-1. This then dissipated to 0.2 ms-1 at 10:00 UTC, coinciding with low tide and the beginning of slack water. Current speed throughout the time series was also higher at depths of 1.0 m to 3.0 m.

 

Irradiance

(a)    (b) 

Figure 8.6 -  (a) Depth vs log(%Iradiance), and (b) Attenuation of light vs time. Sampled on 4 July 2012. All pictures can be enlarged in a new window upon clicking on the relevant image.

Irradiance exponentially decreased with depth and began to plateau at around 4.0 m. There were a few points that showed a higher than expected level of irradiance for below 1.0 m, however these were taken as anomalies (11:00 UTC) and ignored. These anomalies are likely to have been caused by errors in the procedure. The equation of the regression line calculated was: y = 0.9159 + 2.2219. Attenuation of light per meter was also calculated, this showed a decrease in the attenuation of light per meter throughout the day. This means that as the day progressed, more light reached deeper depths.

 

Conclusion

The time series of data was taken when the tide was ebbing and finished just before low tide. This suggests the increase in chlorophyll levels towards the end of the time series may have been caused by the flushing of phytoplankton down from the upper estuary with the ebbing tide, pulling these chlorophyll rich waters from the upper estuary. This would cause an increase in phytoplankton levels in lower sections of the estuary. Using chlorophyll as an indicator for phytoplankton; as the tide ebbs some phytoplankton species need to remain at certain salinities. Therefore will move with the water mass as the tide ebbs (María Trigueros. J., Orive. E., 2000).

The pH remained constant throughout the water column in the time series recorded. However the ‘patches’ of increased pH may have been caused by the usage of boats in the area, or by the chain ferry crossing the estuary upstream from the pontoon. The increased alkalinity of the water may have also been caused by an increase in precipitation the previous night causing an increase in freshwater being flushed down the estuary during the ebb and an increase in land run-off that would contribute (EPA, 2009).

As the tide ebbed, the percentage saturation of dissolved oxygen increased. The ebbing tide would cause phytoplankton to be flushed out of the estuary seaward, thus increasing the percentage of phytoplankton surrounding the pontoon. This increase in phytoplankton would cause an increase in photosynthesis, resulting in an increase in oxygen in the water column.

The increase in temperature approaching low tide could be explained by the ebbing tide flushing warm water from the upper shallower parts of the estuary downstream.  Also, due to the time of day the time series was taken, as the time series progressed the water column was subjected to increased solar radiation from morning to afternoon. There was evidence of a small amount of stratification in the water column, however this was only visible in the deeper waters before low tide.

Salinity decreased as the time series progressed. This may have been caused by the increased input of freshwater from the upper parts of the estuary as the ebbing tide would cause an increase in freshwater into the area and a decrease in coastal waters.

Irradiance exponentially decreased with depth as light increasingly dissipated. This was also shown on a time series, throughout the day, the attenuation of light per meter decreased and more light reached deeper areas of the water column.

The average velocity reflected the state of the tide when the data was collected. There was a faster flow as the tide was ebbing, which slowly dissipated as low tide was reached and the flow slacked.  The direction of flow also supported this, showing the water flowed past the pontoon in a coastal direction.

Return to Top of Page

 

9. References                                                                                                Return to Top of Page

Cornwall and the Isles of Scilly Coastal Advisory Group, 2011, ‘Zone Point to Nare Point-Introduction.’, Cornwall and Isles of Scilly Shoreline Management Plan, [Online], Available: http://www.ciscag.org/finalsmpindex.html, [accessed 2012 July 4th]

Dr Dipper, F., 2001, ‘British Sea Fishes.’, 2nd ed. Middlesex: Underwater World Publications Ltd.

EPA, 2009, ‘Volunteer Estuary Monitoring Manual, A Methods Manual Second Edition.’, [Online], Available: http://water.epa.gov/type/oceb/nep/upload/2009_03_13_estuaries_monitor_chap11.pdf, [accessed: 2012, July 6th]

Falmouth Cruise project, 2012, Available online at:   http://www.falmouthport.co.uk/commercial/html/documents/FHCExhibitionPanels.pdf , [accessed on 05/07/2012]

Gibbons, B., 2001. ‘Seashore Life of Britain and Europe.’, London: New Holland Publishers (UK) Ltd.

Gibson, R., Hextall, B., Rogers, A., 2001, ‘Photographic guide to the Sea and Shore Life of Britain and North-west Europe.’, Oxford: Oxford University Press.

 Grasshoff, K., K. Kremling, & M. Ehrhardt. (1999). ‘Methods of seawater analysis.’, 3rd ed. Wiley-VCH. 

Harper, Jr, H.E., 1975, ‘Silica, diatoms and Cenozoic radiolarian evolution.’, Geology, 3(4), p.175-177

 Huot, Y., Babin, M., Bruyant, F., Grob, C., Twardowski, M. S., and Claustre, H., 2007, ‘Relationship between photosynthetic parameters and different proxies of phytoplankton biomass in the subtropical ocean .’ [Online]. Available: http://www.biogeosciences.net/4/853/2007/bg-4-853-2007.pdf, [Accessed 2012, July 6th]

Johnson K. & Petty R.L., 1983, ‘Determination of nitrate and nitrite in seawater by flow injection analysisť.’,  Limnology and Oceanography, 28 1260-1266. 

Kemp W. M., Sampou P., Cafrey J., Mayer M., Henriksen K., Boynton W. R., 1990, ‘Ammonium recycling versus denitrification in Chesapeake Bay sediments.’, Limnol. Oceanogr., 35(7), 1545-1563

Levinton, J., 2008, ‘Marine Biology: Processes in the Open Sea.’, [Online], Available: http://www.oup.com/us/images/hesamplechapters/09-Levinton-Chap09OUP.pdf,  [Accessed 2012, July 6th] 

Loder T. C. & R. P. Reichard, 1981, ‘The Dynamics in Estuaries.’, Estuaries Vol. 4, No 1, p. 64-69

María Trigueros. J., Orive. E., 2000,Tidally driven distribution of phytoplankton blooms in a shallow, macrotidal estuary.’, J. Plankton Res., 22(5): 969-986.

Marine Management Organisation (MMO), 2010, ‘Evidence Summary Falmouth Harbour construction works, Capital Dredge, maerl mitigation.’, Available online at: http://www.marinemanagement.org.uk/licensing/public_register/cases/documents/falmouth/evidence_summary.pdf, [accessed on 05/07/2012]

Natural England, ‘Maerl Beds Information.’, Available online at; http://www.naturalengland.org.uk/ourwork/marine/mpa/mcz/features/habitats/maerl.aspx, [accessed on 05/07/2012]

Naylor, P., 2005, ‘Great British Marine Animals.’, 2nd ed. Deltor: Sound Diving Publications.

Officer C. B. & Ryther J. H., 1980, ‘The possible importance of silicon in marine eutrophication.’, Mar. Ecol. Prog. Ser 3: 83-91

Parsons T. R., Maita Y. & Lalli C., 1984, ‘A manual of chemical and biological methods for seawater analysisť.’, Pergamon.

Pingree, R. D., Pugh, P. R., Holligan, P. M.,Forster, G. R., 1975, ’Summer phytoplankton blooms and red tides along tidal fronts in the approaches to the English Channel.’, Nature, 258, 672 – 677.

Pingree R. D., Maddock L., & Butler E. I., 1977, ‘The influence of biological activity and physical stability in determining the chemical distributions of inorganic phosphate, silicate and nitrate.’

Roberge P. R., 2000, ‘Handbook of Corrosion Engineering.’, McGraw-Hill, [Online], Available: www.corrosion-doctors.org/seawater/oxygen.htm, [Assessed: 6th July 2012].

Solly, N.S and Knowles, H, 2009, ‘Maerl Surface area coverage in Fal and Helford SAC.’, Available online at:  http://www.marinemanagement.org.uk/licensing/public_register/cases/documents/falmouth/maerl_surface_area.pdf , [accessed on 05/07/2012].

Stewart R., 2005, ‘Equations of motion with viscosity’, Introduction to Physical Oceanography, [Online], Available: http://oceanworld.tamu.edu/resources/ocng_textbook/contents.html, [accessed 2012 July 1st].

Sundby et al, 1992, ‘The Phosphorus cycle in coastal marine sediments.’, Limnology & Oceanography, 37(6), pp 1129  1145.

University of Wiscosin, ‘Dissolved Oxygen’, Watershed Monitoring Program, [Online], Available: www.uwgb.edu/watershed/data/monitoring/oxygen.htm, [accessed: 6th July 2012].

Western Channel Observatory, [Online], 2007, Available: http://www.westernchannelobservatory.org.uk/data.php, [accessed 2012, June 28th].

 Widdicombe C. E., Eloire D., Harbour, R. P. Harris, & P. J. Somerfield., ‘Long-term phytoplankton community dynamics in the Western English Channel.’, J. Plankton Res. (2010), Volume 32(5): 643-655.

 

 

The views expressed in this website do not reflect the views of the University of Southampton, the National Oceanography Centre Southampton, or the Falmouth Marine School (2012).

 

Return to Top of Page