and philip a. gale*

22
Supramolecular methods: the chloride/nitrate transmembrane exchange assay Laura A. Jowett 1 and Philip A. Gale* 1 1 School of Chemistry, The University of Sydney, NSW 2006, Australia [email protected] The chloride/nitrate anion exchange assay is one of the most common methods of measuring chloride transport across lipid bilayers. In this methods paper, we provide a step-by-step procedure for using this assay based on ion selective electrodes and highlight the limitations of this approach in assessing the transmembrane chloride transport capability of anionophores. Keywords: supramolecular chemistry, anion transport, ion-selective electrode, assay Introduction Anion transport is a well-established, and still expanding field of research due to its relevance in biological systems. 1,2 There are many membrane spanning ion-channels responsible for critical biological processes, such as maintaining cellular pH and volume, and cellular signalling. 3 Defects in these channels can reduce the flux of anions through membranes and lead to life-limiting diseases known as channelopathies; examples include Bartter syndrome and the more commonly known cystic fibrosis. 4-7 Small molecule anion transporters have been developed with the potential to bypass these faulty channels and restore the transport of anions across membranes. There are various methods available for studying anion transport, including assays that use pH sensitive fluorescent dyes, voltage clamp measurements and NMR assays. However, one of the most widely used is the chloride/nitrate (Cl - /NO3 - ) exchange ion selective electrode (ISE) assay, shown in figure 1, which is used to directly measure chloride transport. 8-10

Upload: others

Post on 04-Jan-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: and Philip A. Gale*

Supramolecular methods: the chloride/nitrate transmembrane exchange

assay

Laura A. Jowett1 and Philip A. Gale*1

1School of Chemistry, The University of Sydney, NSW 2006, Australia

[email protected]

The chloride/nitrate anion exchange assay is one of the most common methods of

measuring chloride transport across lipid bilayers. In this methods paper, we provide a

step-by-step procedure for using this assay based on ion selective electrodes and

highlight the limitations of this approach in assessing the transmembrane chloride

transport capability of anionophores.

Keywords: supramolecular chemistry, anion transport, ion-selective electrode, assay

Introduction

Anion transport is a well-established, and still expanding field of research due to its relevance in

biological systems.1,2 There are many membrane spanning ion-channels responsible for critical

biological processes, such as maintaining cellular pH and volume, and cellular signalling.3 Defects in

these channels can reduce the flux of anions through membranes and lead to life-limiting diseases

known as channelopathies; examples include Bartter syndrome and the more commonly known cystic

fibrosis.4-7 Small molecule anion transporters have been developed with the potential to bypass these

faulty channels and restore the transport of anions across membranes. There are various methods

available for studying anion transport, including assays that use pH sensitive fluorescent dyes, voltage

clamp measurements and NMR assays. However, one of the most widely used is the chloride/nitrate

(Cl-/NO3-) exchange ion selective electrode (ISE) assay, shown in figure 1, which is used to directly

measure chloride transport.8-10

Page 2: and Philip A. Gale*

Figure 1. A schematic of the chloride/nitrate exchange ion selective electrode assay.

The first reports of the Cl-/NO3- exchange ISE assay were in the early 2000’s11-13 and since then it has

been used for directly monitoring chloride transport.14,15 In this assay, lipid vesicles are prepared in

sodium chloride solution and, using dialysis, the salt in the external solution can be exchanged for

sodium nitrate. This provides a vesicle suspension with a chloride concentration gradient between the

intra- and extra-vesicular solutions allowing anionophores to facilitate transport passively down the

gradient. The assay relies on the fact that 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC)

lipid bilayers are impermeable to chloride in the absence of an anionophore. Chloride efflux mediated

by a transporter is detected using a chloride selective electrode in the external solution, whilst back

transport of nitrate dissipates the charge gradient created by chloride transport and results in overall Cl-

/NO3- exchange. This assay provides direct detection of the chloride anion, allowing unambiguous

determination of chloride transport. The back transport of nitrate is assumed to be facile as it is more

lipophilic than chloride,16 however caution should be displayed when interpreting the results obtained

as it has been shown that nitrate transport can be the rate-limiting transport step for some classes of

anion transporters.17 Therefore, it is best to use a combination of assays to determine relative transport

efficiencies: the Cl-/NO3- exchange ISE assay to show direct chloride transport, and uniport assays to

measure the transport of one anion alone.

Many classes of compounds have been tested in the Cl-/NO3- exchange ISE assay demonstrating its

versatility. The Davis group have reported cholapod anion transporters, a series of compounds

comprising a cholic acid steroidal scaffold with appended anion binding groups.18 The first transporters

were tested in the Cl-/NO3- exchange ISE assay, with the best transporter 1 being shown in figure 2.12,19

As the compounds increased in complexity they became highly lipophilic resulting in poor aqueous

solubility.20 This solubility issue led to the development of alternative assays but highlighted a major

benefit of the Cl-/NO3- exchange ISE assay; the intrinsic deliverability test. The ability of the transporter

to be delivered to vesicles in aqueous solution is monitored; an important consideration when designing

Page 3: and Philip A. Gale*

potential therapeutics. Our group has used the Cl-/NO3- exchange ISE assay extensively for a wide range

of small molecule chloride transporters, including urea and thiourea tren-based compounds 2a and 2b,

figure 2,21,22 squaramide and thiosquaramide compounds,23,24 ortho-phenylenediamine compounds,25,26

and indole-based anionophores.27-29

Figure 2. Two compounds tested in the Cl-/NO3- exchange ISE assay. The cholapod anion transporter

(1) that achieved the highest percentage chloride efflux. One example of a (thio)urea tren based compound (2) that was tested in the Cl-/NO3

- exchange ISE assay.

An additional benefit of the Cl-/NO3- exchange ISE assay is the flexibility of conditions that can be

employed when studying chloride transport. The pH of the external solution can be tuned using a range

of buffer solutions allowing the transport capabilities to be determined at different pH. One study

involving phenylthiosemicarbazones showed that at pH 7.2 there was no chloride transport, but at pH

4.0 protonation of the transporter occurs resulting in a conformational change that allows the compound

to facilitate chloride transport.30 Azobenzene-based chloride transporters are another class of

compounds tested using the Cl-/NO3- exchange ISE assay. The transport was initially tested with the cis

and trans form separately. The closed cis isomer showed transport, whereas the open trans isomer did

not. In a subsequent experiment, in situ irradiation of the trans isomer induced a conformational change

to the cis isomer, switching the chloride transport on.31

Here, we give an overview of the experiments we perform, the equipment we use and a step-by-step

guide of how to conduct vesicle-based experiments in the Cl-/NO3- exchange ISE assay.

Experiments

The Cl-/NO3- exchange ISE assay is an excellent tool to study ion transport under varying pH conditions

as it can be easily modified by changing the internal and external buffer solutions to perform

experiments over a pH range. Small molecule ion transporters that can switch transport on or off

depending on pH could mimic the function of acid-sensing ion channels.32 Alternatively, tuning the pH

Page 4: and Philip A. Gale*

at which transport occurs has the potential to target ion transport in specific cell types or cellular

organelles. We have reported three classes of pH switchable anion transporters, the thiosquaramides,24

the oxothiosquaramides33 (both in collaboration with Jolliffe), and the phenylthiosemicarbazones 3 –

5.30 In these examples the Cl-/NO3- exchange ISE assay was run at pH 7.2 using sodium phosphate

buffered (5 mM) solutions and then compared to the results from the Cl-/NO3- exchange ISE assay tested

at pH 4.0 using sodium citrate buffered (5 mM) solutions. In all cases, significantly less chloride

transport was observed in near neutral conditions but upon repeating the experiments at an acidic pH, a

high percentage chloride efflux was observed, results for phenylthiosemicarbazones 3 – 5 are shown in

figure 3. One potential use for this family of compounds could be in targeting transporters to acidic

cellular organelles such as lysosomes. We have also explored the possibility of tuning the pKa of a

protonatable nitrogen atom in the perenosins,28,29 and probed the effect subtle differences in the pH

environment have on the chloride transport abilities. Here, the Cl-/NO3- exchange ISE assay was run at

pH 4.0 and 4.5 using citrate buffered (5 mM) solutions, where the higher degree of protonation resulted

in high chloride transport efficiencies. Poor chloride transport was observed under basic (pH 10

ethanolamine (5 mM) buffer and pH 8.2 sodium phosphate (5 mM) buffer) conditions, due to the lack

of protonation and subsequent chloride binding. As the pH decreased from pH 7.2 to 6.5 (both buffered

with sodium phosphate (5 mM)), or pH 7.2 to 6.2 (buffered with piperazine (5 mM)) the chloride

transport increased for all compounds. Compounds 6 – 8 showed the largest increase in transport, results

are shown in figure 3, and so display the best potential for targeting chloride transport in cancer cells

(external pH 6.734) over healthy cells (external pH 7.2-7.535).

Page 5: and Philip A. Gale*

Figure 3. Results of the Cl-/NO3- exchange ISE assay when testing in different pH environments. Left

– Phenylthiosemicarbazone compounds (3 - 5) show a large switch on in chloride transport in acidic environments, vesicle experiments compared at pH 7, open symbols, and pH 4, closed symbols. Right – Perenosin compounds (6 – 8) show an increase in chloride transport in slightly acidic environments, vesicle experiments compared at pH 7.2, open symbols and pH 6.5, closed symbols.

Chloride, bicarbonate and phosphate are the most abundant anions in the body,8 but other anions, such

as fluoride,36 sulfate37 and nitrate,38 and cations, such as sodium, potassium, magnesium and calcium,39

are biologically relevant. A major benefit of the ISE assay is its ease of modification, meaning it is

applicable for a range of transport studies, and can incorporate a large array of positively and negatively

charged species. Altering the internal and external salt solutions and/or using the corresponding ISE are

both ways to change the assay to study other systems. Recently, we have shown using a fluoride

selective electrode that a modified ISE assay can be employed to directly monitor fluoride transport.40

Sulfate has a high hydration energy41 and therefore poses a challenge for neutral receptors to bind and

transport across lipid bilayers. For this reason it has been used as an external, “un-transportable” counter

ion in the Cl-/HCO3- exchange ISE assay. The Cl-/HCO3

- exchange ISE assay is prepared by

encapsulating sodium chloride inside vesicles and then suspending them in an external solution of

sodium sulfate. The transporter is added to the vesicle suspension at t = 0 s and initially, negligible

chloride transport is observed. Addition of a bicarbonate pulse at t = 120 s allows Cl-/HCO3- exchange

Page 6: and Philip A. Gale*

to occur, which increases the external chloride concentration and is subsequently detected by chloride

ISE. From the Cl-/HCO3- exchange ISE assay, our group has shown a wide variety of compounds are

capable of Cl-/HCO3- exchange, including urea and thiourea tren-based compounds,21 ortho-

phenylenediamines25 and squaramides.23,42 In one particular study,22 with tren-based compounds, a high

degree of background transport was observed from t = 0 s to t = 120 s (i.e. before the addition of sodium

bicarbonate) resulting from Cl-/SO42- exchange. This led to further investigations into sulfate transport

and the development of a Cl-/SO42- exchange assay to monitor chloride efflux. Studies using this assay

were conducted to show (thio)urea tren-based, and strapped peptide compounds were capable of Cl-

/SO42- exchange whilst a 33S NMR assay was used to directly monitor sulfate transport.43

Some transporters are capable of facilitating ion pair transport and, when the compounds display good

activity in the Cl-/NO3- exchange ISE assay, it can subsequently be used to probe this activity. Using

NaCl, KCl, RbCl and CsCl as the internal solution and NaNO3 as the external solution, the assay was

tested with the strapped calixpyrroles; which showed ion pair transport only in the presence of CsCl.44

This experiment has also been used to rule out the possibility of M+/Cl- symport for a number of

compounds.25,45

Comparison of the anion transport ability between different classes of transporter, or the same class of

transporter with different anions, is important and can be achieved through quantifying the extent of

anion transport by obtaining EC50 values (the concentration of transporter needed to achieve 50%

chloride efflux in a set time period). Dose-response studies are performed by measuring the increase in

external chloride, using an ISE, over time at different transporter concentrations. Hill analysis46 is then

conducted to afford an EC50 value and a Hill coefficient, n, which is indicative of the stoichiometry of

the transporting complex. Results of dose response studies and subsequent Hill analysis for tren based

thiourea 9 can be found in figure 4.17 These studies have been performed routinely for a large variety

Page 7: and Philip A. Gale*

of compounds including urea and thiourea tren-based compounds,22 N,N′-

(phenylmethylene)dibenzamide compounds47 and acylthiourea compounds.48

Figure 4. Results of the Cl-/NO3- exchange ISE assay

from an alkyl thiourea tren-based compound (9) at different concentrations. Left – Dose response studies. Right –

Hill analysis of the data affording an EC50 value (k) and Hill coefficient (n).

Equipment

Before starting an anion transport assay, there are some essential equipment and materials required to

perform vesicle-based experiments.

1. The lipid

The choice of lipid is an important consideration as lipids have diverse properties which can affect the

transport experiments. In 2015 we performed a study using the Cl-/NO3- exchange ISE assay and found

that the chloride transport rate varied for four lipids tested; these were POPC, 10, see figure 5, 1-

palmitoyl-2oleoyl-sn-glycero-3-phosphoglycerol (POPG), 1-palmitoyl-2-oleoyl-sn-glycero-3-

phosphoethanolamine (POPE), and dipalmitoyl-sn-glycero-3-phosphocholine (DPPC). The chloride

transport abilities in DPPC vesicles were measured at a higher temperature due to the higher phase

Page 8: and Philip A. Gale*

transition temperature of DPPC, and vesicles of a 3:1 POPE:POPC ratio were used. POPE had to be employed

as a lipid mixture because alone it preferentially forms multilamellar vesicles which contain different lipid

phases in the same vesicle. In this particular study the chloride transport rates in POPC and DPPC vesicles

were the same (tests performed at different temperatures), in POPG vesicles the rate decreased, and in

mixed POPE/POPC vesicles the rate increased compared to POPC.49 It should be noted that these results

might be different if the temperature and lipid compositions were changed. The study concluded that

POPC was most suitable for vesicle-based experiments due to the moderate transport rates and the

reproducibility of results.

The lipid chosen for vesicle-based transport experiments should be 1) easy to work with in the

laboratory, 2) produce repeatable results, 3) within the budget allocated, and most importantly 4) mimic

biological membranes of interest (especially if the research involves developing and testing novel

therapeutics). In the Gale group, we have used POPC from Corden Pharma and Avanti Polar Lipids for

many years,22,30,48 the A.P. Davis group has also used POPC from Avanti Polar Lipids.50-52 The Matile

group has used egg yolk phosphatidylcholine (EYPC) from Avanti Polar Lipids,53-55 and more recently

the Talukdar group has used EYPC from the same supplier.56-58 These lipids are chosen because they

mimic mammalian phospholipid composition (where POPC is the main component) and hence provide

a suitable model membrane for transport experiments. EYPC is a natural lipid mixture, with POPC

being the main component.59 The EYPC mixture contains lipids that have different aliphatic tail chains

(specific information can be found on the supplier website). Within the purchased product a variety of

tail lengths and different tail chain saturations are found; and this can therefore result in slightly different

compositions between purchased batches and an average molecular weight of the lipid. POPC is a

synthetic lipid product with certified purity (see supplier website for more information). This ensures

that the purchased synthetic lipid is consistent between batches and so gives reliable experimental

results.60 Furthermore, we have found that the free fatty acid content can vary with different suppliers

of lipid and, although this does not affect the Cl-/NO3- exchange ISE assay, it can have consequences in

other transport assays.61

Figure 5. Structure of commonly used POPC lipid for vesicle-based anion transport assays.

Page 9: and Philip A. Gale*

2. The salts

Anion transport experiments can focus on anion exchange, anion uniport, or osmosis accompanying

salt transport. In all these cases buffer solutions are required for the internal and external solutions of

the vesicles. The salts we use for preparing the buffer solutions are analytical grade chemicals with high

purity, usually ≥ 99 %.

Standard assay conditions have been developed by our group for the Cl-/NO3- exchange ISE assay. We

have used sodium phosphate (5 mM) buffer for experiments conducted at pH 7.2. This is made up of

the acidic and basic form of sodium phosphate, sodium dihydrogen phosphate dihydrate and disodium

hydrogen phosphate. The internal solution contains sodium chloride (487 mM) and the external solution

contains sodium nitrate (487 mM). The total ionic strength of the internal and external solutions must

be kept constant throughout the duration of an experiment. For experiments conducted at an acidic pH,

a citrate or piperazine buffer solution is used, and for experiments performed at a basic pH either

tris(hydroxymethyl)aminomethane or ethanolamine buffer solution is used.

3. The ion selective electrode

ISEs have long been used to measure ion concentration and activity in aqueous solution.62 Some benefits

include that they are simple to use, they are very robust when properly cared for and can be used for a

wide range of applications, such as analysis of water, soil, and dairy products. However, there are

limitations to note, the electrodes do tend to have a limited shelf life, and signal interference can occur

with certain ions, such as CN- and S- if they are present as contaminants in the test solution (check the

electrode specification for more details).63,64 There are several types of ISE available with different

sensor types and operations.65 Solid state electrodes, where the electrode potential is measured across a

solid, polished crystalline membrane, have a long life time, rapid response times and minor maintenance

requirements. One example of this is the Fisherbrand accumet Chloride Combination Electrode66 we

use in the Gale group to detect chloride efflux in vesicle experiments.

Chloride selective electrodes vary depending on the manufacturer, therefore the user manual for the

electrode should be consulted before use.67,68 The combination chloride solid state electrode our group

uses contains a sensing element, the membrane, which develops an electrode potential when in contact

with the test solution containing chloride. The potential, which is dependent on the free chloride ions in

the solution, is measured against a reference with a digital pH/mV meter or a specific ion meter (we use

an Orion Star Conductivity Benchtop Meter69). The reference electrode, found either internally in the

case of combination electrodes or externally in the case of half cell electrodes, contains a known and

fixed concentration of the ion of study. The Nernst equation is used to describe the measured potential

which corresponds to the chloride ions in the test solution:

E = E0 + S log(A)

Page 10: and Philip A. Gale*

Where E is the measured electrode potential, E0 is the reference potential, S is the electrode slope and

A is the chloride ion activity level in solution. Often electrode calibration is required before use which

allows E0 and S to be determined, again consult the user manual for more information on the electrode.

The chloride ion activity is related to the free chloride ion concentration by the activity coefficient:

A = γ Cf

Where A is the chloride ion activity level in solution, γ is the activity coefficient and Cf is the free

chloride ion concentration. Activity coefficients are usually dependent on the total ionic strength but

when the background ionic strength is high and constant, relative to the sensed ion concentration (which

is the case for buffer solutions prepared in the same way as described in section 2. The Salts), the activity

coefficient is constant and therefore directly proportional to the chloride ion concentration.

Over the course of an experiment the electrode potential is recorded every 5 s with a digital pH/mV

meter (see above) connected to a PC. At the end of an experiment the recorded electrode potentials are

converted into chloride concentrations (C) using the equations described above. These concentrations

are then normalised:

Cf = C – C0

Where Cf is the final concentration, C is the calculated concentration at each electrode potential reading

and C0 is the initial concentration. Normalisation then allows a percentage chloride efflux to be

calculated at the end of each experiment.

4. Other things to think about.

As well as the important materials detailed in the above sections, there is some other equipment required

for vesicle-based anion transport experiments. Drying the lipid requires a rotary evaporator and a high

vacuum pump to ensure a smooth lipid film is created. An analytical balance is used to obtain the exact

weight of the lipid and for weighing the salts for the buffer solutions. Preparing the internal and external

buffer solutions requires access to ultrapure, Type 1 (also known as Milli-Q) water to prevent ion

contamination. A solution of Triton X-100 (11 w%, in H2O:DMSO 7:1 v:v) is readily required for lysing

the vesicles in order to calibrate the 100% chloride efflux value. Liquid nitrogen, an extrusion kit

including 200 nm polycarbonate membranes and dialysis tubing are essential for the vesicle preparation

steps. Our group also uses stopwatches, micropipettes, micro PTFE stirrer bars and single use glass

vials (10 – 20 mL), that fit the electrode in with a small amount of space for a pipette tip, when

performing the experiments. An example of the experimental set up is shown in figure 6.

Page 11: and Philip A. Gale*

Figure 6. An example of the experimental set up used by the Gale group when running a Cl-/NO3-

exchange ISE assay.

Procedure

Below is a general procedure, as used by the Gale group for preparing vesicles and running a Cl-/NO3-

exchange ISE assay at pH 7.2.

Step-by-step Solution Preparation

1. Prepare the buffer solution. Dissolve sodium dihydrogen phosphate dihydrate (0.294 g, 0.0019

mol) and disodium hydrogen phosphate (anhydrous) (1.150 g, 0.0081 mol) in ultrapure, Type

1 water (2 L) in a beaker (Figure 7).

2. Prepare the internal solution. Dissolve sodium chloride (2.846 g, 0.0487 mol) in the phosphate

buffer solution (100 mL) and when fully dissolved carefully adjust the pH to 7.20 if necessary

– using HCl to decrease the pH, or NaOH solution to increase the pH.

Page 12: and Philip A. Gale*

3. Prepare the external solution. Dissolve sodium nitrate (78.632 g, 0.9251 mol) straight into the

remaining phosphate buffer solution (1900 mL) and when fully dissolved carefully adjust the

pH to 7.2 – using HCl or NaOH solution to adjust the pH as required.

Step-by-step Vesicle Preparation

1. Dissolve POPC (1 g) in chloroform (35 mL), resulting in a 37.5 mM solution.

2. Weigh a round bottom flask and add between 0.5 – 4 mL of the chloroform solution using a

volumetric pipette (remember how much you add as this will be used in step 6), depending on

the number of experiments to run (approximately 0.15 mL of lipid per experiment to run).

3. Remove the organic solvent gently on a rotary evaporator, creating a smooth lipid film. Ensure

the flask is rotating slowly and reduce the pressure gradually until chloroform begins to

evaporate (Figure 8).

4. Dry the lipid film in a vacuum desiccator for a minimum of 6 hrs, preferably overnight.

5. Weigh the round bottom flask and lipid film, this is essential for calculating the concentration

of lipid.

6. Using an autopipette add the same volume of internal solution as the lipid chloroform solution

used in step 2.

7. Ensure the lipid is fully hydrated by vortexing the solution for 5 minutes, or until there is no

lipid remaining on the interior of the flask (Figure 9).

8. Perform nine freeze-thaw steps. Fully freeze the lipid suspension in liquid nitrogen (invert to

check all the solution is frozen), then defrost in lukewarm water and repeat. Dry the outside of

the round bottom flask before placing in liquid nitrogen (Figure 10).

9. Allow the lipid solution to rest for 30 minutes at room temperature.

10. During this time wash all parts of the extrusion kit with ethanol then water. Repeat the washing

three times and then assemble the extrusion kit with a 200 nm polycarbonate membrane.

Instructions can be found on the Avanti lipid website for help with the assembly.70

11. Set aside some of the external sodium nitrate solution (300 mL approx.) to be used later. This

can be stored in a sealed jar in a fridge for approximately 1 week, note that the solution should

be equilibrated to room temperature before opening.

12. Cut a piece of dialysis tubing to the appropriate length (5 cm per 1 mL), place a buoyant clip

on one end and put the tubing in the remaining external sodium nitrate solution (1600 mL

approx.) leaving it to stir.

13. Fill one of the syringes with 1 mL of the lipid solution and place it into one side of the extrusion

kit. Leaving the other syringe empty, place it in the opposite side. Ensure the syringe needle

tips are well tightened to prevent leakage.

14. Push the lipid solution back and forth through the extrusion kit twenty-five times then place the

extruded lipid solution in a vial, recording the volume. Repeat the extrusions using the same

Page 13: and Philip A. Gale*

syringe for the unrefined lipid and extruded lipid each time and combine the remaining lipid

(Figure 11) . Clean the extrusion kit with ethanol and then water.

15. Take the wetted dialysis tubing out of the external solution using tweezers, dry slightly on some

lab paper and then open the tubing with tweezers.

16. Whilst holding the tubing open with tweezers, carefully pipette the extruded lipid solution into

the dialysis tubing and close the top with a second buoyant clip.

17. Place the clipped dialysis tubing containing the extruded lipid solution into the external solution

and leave stirring gently for a minimum of two hours, preferably overnight. The duration of

this dialysis step depends on the pH at which experiments are to be performed. For vesicles

being tested at acidic pH, pH 4 to pH 6, this should be 2hrs, at all other pH dialysis can be left

overnight

18. After dialysis, carefully remove the tubing containing the lipid from the external solution using

tweezers and transfer the lipid solution into a volumetric flask (10 mL).

19. Top up the volumetric flask with the external solution set aside in step 11 and transfer the final

lipid solution to a capped vial.

20. Calculate the concentration of lipid in your vial, see calculations below:

Mass of lipid (g) = Mass recorded in Step 5 (g) – Mass recorded in Step 2 (g)

Moles of lipid (mol) = Mass of lipid (g) / Molecular weight of lipid (gmol-1) (760.09 gmol-1 for

POPC)

Concentration of lipid (moldm-3) = Moles of lipid (mol) / Volume of volumetric flask (L) (0.010 L)

21. Use this concentration to calculate how much lipid to use in each experiment. The desired lipid

solution has a concentration of 1 mM in 5 mL, see calculations below:

Volume of lipid for experiment (L) = (Desired concentration (0.001 M) x Final solution volume

(0.005 L)) / Lipid concentration (moldm-3)

Step-by-step Experimental Procedure

1. Prepare the electrode; in general this requires a top up of the internal electrode solution, then to

wash the electrode with water, ethanol and water, then calibrate the electrode. See the

instruction manual for more specific instructions for each electrode.

2. Prepare a known concentration DMSO solution (0.5 mL approx.) of the transporter to be tested.

Here it is best to make a concentrated stock solution (around 50 mM) that can be diluted for

subsequent measurements.

Page 14: and Philip A. Gale*

3. Prepare the experimental solution by adding the calculated volume of lipid for the experiment

(step 21 of vesicle preparation) using a micropipette to a single use glass vial equipped with a

clean micro PTFE stirrer bar. Make up the solution to 5 mL by adding the external solution

(prepared in step 3 of solution preparation and set aside in step 11 of vesicle preparation) using

a micropipette.

4. Place the electrode in the experimental solution and set the stirring to 400 rpm.

5. Prepare to begin the experiment; set the stopwatch and/or the electrode reader to zero

(depending on the equipment). Ensure the electrode reader is set to record every 5 s for the

duration of the experiment. Experiments are usually run for 300 s, but this time can be altered

to best suit the transporter being tested.

6. Simultaneously start the stopwatch and/or the electrode reader and, with the pipette tip slightly

submerged in the solution, add DMSO (10 µL) using a micropipette; this will be the control

run.

7. Allow the experiment to proceed, ensuring the electrode reader is recording. The reading on the

electrode should not change because DMSO does not facilitate chloride transport.

8. At 300 s (or at the time determined to end the experiment) add the Triton X-100 detergent

solution (50 µL) to lyse the vesicles.

9. Leave the experiment running for an additional 120 s to allow the 100% chloride efflux reading

to settle, then stop the electrode reader and stopwatch.

10. Remember to save your data.

11. At the end of the experiment dispose of the glass vial and the vesicle solution in the appropriate

waste, wash the micro PTFE stirrer bar with water, ethanol and water and dry for the next

experiment, and wash the electrode with water, ethanol and water and dry for the next

experiment.

12. Repeat steps 3 to 11, adding the transporter solution (10 µL) in place of DMSO in step 6, and

watching the reading electrode decrease as chloride is transported out of the vesicles into the

external solution in step 7. Use all the lipid for running experiments (the lipid should be used

in the same day). The same transporter can be tested at a variety of concentrations to form a

dose-response curve allowing Hill analysis, or many different transporters can be tested in the

same conditions to compare the transport ability of a series of transporters.

Photographs

Page 15: and Philip A. Gale*

Figure 7. Solution preparation step 1. Accurately weighing out salts for the buffer solution.

Figure 8. Vesicle preparation step 3. Removing the organic solvent gently on a rotary evaporator.

Page 16: and Philip A. Gale*

Figure 9. Vesicle preparation step 7. Vortexing the lipid solution to ensure it is fully hydrated.

Figure 10. Vesicle preparation step 8. Left – Fully freezing the lipid solution in liquid nitrogen. Right –

Defrosting the lipid solution in tepid water.

Figure 11. Vesicle preparation step 13/14. Extruding the lipid.

Page 17: and Philip A. Gale*

What to do if…

1. You overshoot the pH when adding acid or base during adjustment (step 2 and 3 of solution

preparation)

If this happens you can use the alternate acid/base to return the pH to 7.2. It should be noted

this is not desirable and should be avoided as it can change the ionic strength of the solution.

2. You forget to weigh the flask (step 2 and 5 of vesicle preparation)

If this happens you cannot calculate the concentration of the lipid, meaning you cannot

determine the ratio of lipid to transporter and therefore the experimental results for that batch

of lipid are void.

3. You forget to separate some of the external solution (step 11 of vesicle preparation)

If this happens and you use all the prepared external solution for dialysis, you have to make

another phosphate buffered sodium nitrate solution to use as the external solution for your

experiments.

4. You lose some of the lipid during the extrusion processes (step 13 and 14 of vesicle preparation)

If this happens you should record the volume of lipid obtained after the extrusion process and

use it to calculate the moles of lipid after extrusion, shown below:

Mass of lipid (g) = Mass recorded in Step 5 (g) – Mass recorded in Step 2 (g)

Moles of lipid (mol) = Mass of lipid (g) / Molecular weight of lipid (gmol-1) (760.09 gmol-1 for

POPC)

Moles of lipid after extrusion (mol) = (Volume of lipid in chloroform solution in Step 2 (mL) /

Volume of lipid after extrusion (mL)) x Moles of lipid (mol)

Concentration of lipid (moldm-3) = Moles of lipid after extrusion (mol) / Volume of volumetric

flask (L) (0.010 L)

5. Your transporter is not soluble in DMSO (step 2 of experimental procedure)

If this is the case, you can dissolve your compound in any solvent the compound is soluble in

as long as the solvent is miscible with water.

6. You forget to start the stopwatch/electrode reader at the beginning of an experiment (step 6 of

experimental procedure)

If this happens you are not recording the electrode reading over the course of the experiment.

It is also unclear when the experiment began and so you cannot determine when to lyse the

vesicles and when to take the final reading. Therefore, the results from this experimental run

are void.

7. Your DMSO run shows significant chloride efflux (step 6 of experimental procedure)

If this happens you should clean the PTFE stirrer bar and the electrode very well before setting

up a new experimental run to remove any chance of contamination with other compounds. If

Page 18: and Philip A. Gale*

after the second blank run there is still significant chloride efflux your vesicles might be leaky

and so are unsuitable for subsequent experiments.

8. Your transporter crashes out when you add it to the vesicle suspension (step 6 (repeated) of

experimental procedure)

If this happens it is likely that your transporter has low solubility in aqueous solution and does

not partition quickly into the membrane. If you used a high concentration DMSO solution of

transporter you can dilute the solution and try at a lower concentration (between 5 mM – 0.5

mM). If it continues to be insoluble at lower concentrations then preincorporation of your

transporter is suggested.19

9. You forget to lyse the vesicles at 300 s or at the time determined to end the experiment (step 8

of experimental procedure)

If this happens and you leave the experiment running for longer than the specified time, the

experiment is no longer consistent and so results could become inconsistent. Therefore, the

results from this experimental run are void.

10. You leave the experiment running for longer than 120 s after the addition of detergent (step 9

of experimental procedure)

If this happens then stop the experiment. For consistency in the results you should set a time

point (at least 120 s, but could be greater) at which you record the final chloride concentration

for all experimental runs.

11. You forget to save your data (step 10 of experimental procedure)

If this happens you will have to repeat the experimental run and then save the data.

12. You cannot get repeatable results

If this happens there could be some interference of your compound with the electrode. This is

sometimes the case at high concentrations, so you could dilute the sample and try at a lower

transporter concentration. It could potentially be due to the electrode, so check the age of the

electrode, make sure, if suitable, the internal solution is filled to the top, and clean the

membrane. This could also be due to poor aqueous solubility, see point 8 of frequently asked

questions

Acknowledgements

LAJ thanks The University of Sydney for a postgraduate scholarship. PAG thanks The University of

Sydney and the ARC (DP180100612) for funding.

References

(1) Busschaert, N.; Gale, P. A. Small-Molecule Lipid-Bilayer Anion Transporters for Biological Applications. Angew. Chem. Int. Ed. 2013, 52, 1374-1382.

Page 19: and Philip A. Gale*

(2) Gale, P. A.; Perez-Tomas, R.; Quesada, R. Anion Transporters and Biological Systems. Acc. Chem. Res. 2013, 46, 2801-2813. (3) Gadsby, D. C. Ion channels versus ion pumps: the principal difference, in principle. Nat. Rev. Mol. Cell. Biol. 2009, 10, 344. (4) Ashcroft, F. M.: Ion Channels and Disease; Academic Press: San Diego, 2000. (5) Vankeerberghen, A.; Cuppens, H.; Cassiman, J.-J. The cystic fibrosis transmembrane conductance regulator: an intriguing protein with pleiotropic functions. J. Cyst. Fibros. 2002, 1, 13-29. (6) Miyamura, N.; Matsumoto, K.; Taguchi, T.; Tokunaga, H.; Nishikawa, T.; Nishida, K.; Toyonaga, T.; Sakakida, M.; Araki, E. Atypical Bartter Syndrome with Sensorineural Deafness with G47R Mutation of the β-Subunit for ClC-Ka and ClC-Kb Chloride Channels, Barttin. J. Clin. Endocrinol. Metab. 2003, 88, 781-786. (7) Ashcroft, F. M. From molecule to malady. Nature 2006, 440, 440. (8) Davis, A. P.; Sheppard, D. N.; Smith, B. D. Development of synthetic membrane transporters for anions. Chem. Soc. Rev. 2007, 36, 348-357. (9) Davis, J. T.; Okunola, O.; Quesada, R. Recent advances in the transmembrane transport of anions. Chem. Soc. Rev. 2010, 39, 3843-3862. (10) Benke, B. P.; Aich, P.; Kim, Y.; Kim, K. L.; Rohman, M. R.; Hong, S.; Hwang, I.-C.; Lee, E. H.; Roh, J. H.; Kim, K. Iodide-Selective Synthetic Ion Channels Based on Shape-Persistent Organic Cages. J. Am. Chem. Soc. 2017, 139, 7432-7435. (11) Schlesinger, P. H.; Ferdani, R.; Liu, J.; Pajewska, J.; Pajewski, R.; Saito, M.; Shabany, H.; Gokel, G. W. SCMTR:  A Chloride-Selective, Membrane-Anchored Peptide Channel that Exhibits Voltage Gating. J. Am. Chem. Soc. 2002, 124, 1848-1849. (12) Koulov, A. V.; Lambert, T. N.; Shukla, R.; Jain, M.; Boon, J. M.; Smith, B. D.; Li, H.; Sheppard, D. N.; Joos, J.-B.; Clare, J. P.; Davis, A. P. Chloride Transport Across Vesicle and Cell Membranes by Steroid-Based Receptors. Angew. Chem. Int. Ed. 2003, 42, 4931-4933. (13) Koulov, A. V.; Mahoney, J. M.; Smith, B. D. Facilitated transport of sodium or potassium chloride across vesicle membranes using a ditopic salt-binding macrobicycle. Org. Biomol. Chem. 2003, 1, 27-29. (14) Spooner, M. J.; Gale, P. A. A tripodal tris-selenourea anion transporter matches the activity of its thio- analogue but shows distinct selectivity. Supramol. Chem. 2018, 1-6. (15) Jowett, L. A.; Howe, E. N. W.; Wu, X.; Busschaert, N.; Gale, P. A. New Insights into the Anion Transport Selectivity and Mechanism of Tren-based Tris-(thio)ureas. Chem. Eur. J. 2018, 24, 10475-10487. (16) Valkenier, H.; Davis, A. P. Making a Match for Valinomycin: Steroidal Scaffolds in the Design of Electroneutral, Electrogenic Anion Carriers. Acc. Chem. Res. 2013, 46, 2898-2909. (17) Yang, Y.; Wu, X.; Busschaert, N.; Furuta, H.; Gale, P. A. Dissecting the chloride-nitrate anion transport assay. Chem. Commun. 2017, 53, 9230-9233. (18) Davis, A. P.; Perry, J. J.; Williams, R. P. Anion Recognition by Tripodal Receptors Derived from Cholic Acid. J. Am. Chem. Soc. 1997, 119, 1793-1794. (19) McNally, B. A.; Koulov, A. V.; Lambert, T. N.; Smith, B. D.; Joos, J.-B.; Sisson, A. L.; Clare, J. P.; Sgarlata, V.; Judd, L. W.; Magro, G.; Davis, A. P. Structure–Activity Relationships in Cholapod Anion Carriers: Enhanced Transmembrane Chloride Transport through Substituent Tuning. Chem. Eur. J. 2008, 14, 9599-9606. (20) Judd, L. W.; Davis, A. P. From cholapod to cholaphane transmembrane anion carriers: accelerated transport through binding site enclosure. Chem. Commun. 2010, 46, 2227-2229. (21) Busschaert, N.; Gale, P. A.; Haynes, C. J. E.; Light, M. E.; Moore, S. J.; Tong, C. C.; Davis, J. T.; Harrell, J. W. A. Tripodal transmembrane transporters for bicarbonate. Chem. Commun. 2010, 46, 6252-6254.

Page 20: and Philip A. Gale*

(22) Busschaert, N.; Wenzel, M.; Light, M. E.; Iglesias-Hernández, P.; Pérez-Tomás, R.; Gale, P. A. Structure–Activity Relationships in Tripodal Transmembrane Anion Transporters: The Effect of Fluorination. J. Am. Chem. Soc. 2011, 133, 14136-14148. (23) Busschaert, N.; Kirby, I. L.; Young, S.; Coles, S. J.; Horton, P. N.; Light, M. E.; Gale, P. A. Squaramides as Potent Transmembrane Anion Transporters. Angew. Chem. Int. Ed. 2012, 51, 4426-4430. (24) Busschaert, N.; Elmes, R. B. P.; Czech, D. D.; Wu, X.; Kirby, I. L.; Peck, E. M.; Hendzel, K. D.; Shaw, S. K.; Chan, B.; Smith, B. D.; Jolliffe, K. A.; Gale, P. A. Thiosquaramides: pH switchable anion transporters. Chem. Sci. 2014, 5, 3617-3626. (25) Moore, S. J.; Haynes, C. J. E.; Gonzalez, J.; Sutton, J. L.; Brooks, S. J.; Light, M. E.; Herniman, J.; Langley, G. J.; Soto-Cerrato, V.; Perez-Tomas, R.; Marques, I.; Costa, P. J.; Felix, V.; Gale, P. A. Chloride, carboxylate and carbonate transport by ortho-phenylenediamine-based bisureas. Chem. Sci. 2013, 4, 103-117. (26) Karagiannidis, L. E.; Haynes, C. J. E.; Holder, K. J.; Kirby, I. L.; Moore, S. J.; Wells, N. J.; Gale, P. A. Highly effective yet simple transmembrane anion transporters based upon ortho-phenylenediamine bis-ureas. Chem. Commun. 2014, 50, 12050-12053. (27) Moore, S. J.; Wenzel, M.; Light, M. E.; Morley, R.; Bradberry, S. J.; Gomez-Iglesias, P.; Soto-Cerrato, V.; Perez-Tomas, R.; Gale, P. A. Towards "drug-like" indole-based transmembrane anion transporters. Chem. Sci. 2012, 3, 2501-2509. (28) Van Rossom, W.; Asby, D. J.; Tavassoli, A.; Gale, P. A. Perenosins: a new class of anion transporter with anti-cancer activity. Org. Biomol. Chem. 2016, 14, 2645-2650. (29) Jowett, L. A.; Howe, E. N. W.; Soto-Cerrato, V.; Van Rossom, W.; Pérez-Tomás, R.; Gale, P. A. Indole-based perenosins as highly potent HCl transporters and potential anti-cancer agents. Sci. Rep. 2017, 7, 9397. (30) Howe, E. N. W.; Busschaert, N.; Wu, X.; Berry, S. N.; Ho, J.; Light, M. E.; Czech, D. D.; Klein, H. A.; Kitchen, J. A.; Gale, P. A. pH-Regulated Nonelectrogenic Anion Transport by Phenylthiosemicarbazones. J. Am. Chem. Soc. 2016, 138, 8301-8308. (31) Choi, Y. R.; Kim, G. C.; Jeon, H.-G.; Park, J.; Namkung, W.; Jeong, K.-S. Azobenzene-based chloride transporters with light-controllable activities. Chem. Commun. 2014, 50, 15305-15308. (32) Yoder, N.; Yoshioka, C.; Gouaux, E. Gating mechanisms of acid-sensing ion channels. Nature 2018, 555, 397. (33) Elmes, R. B. P.; Busschaert, N.; Czech, D. D.; Gale, P. A.; Jolliffe, K. A. pH switchable anion transport by an oxothiosquaramide. Chem. Commun. 2015, 51, 10107-10110. (34) Mazzio, E. A.; Smith, B.; Soliman, K. F. A. Evaluation of endogenous acidic metabolic products associated with carbohydrate metabolism in tumor cells. Cell Biol. Toxicol. 2010, 26, 177-188. (35) Bailey, K. M.; Wojtkowiak, J. W.; Hashim, A. I.; Gillies, R. J. Targeting the Metabolic Microenvironment of Tumors. Adv. Pharmacol. 2012, 65, 63-107. (36) Stockbridge, R. B.; Lim, H.-H.; Otten, R.; Williams, C.; Shane, T.; Weinberg, Z.; Miller, C. Fluoride resistance and transport by riboswitch-controlled CLC antiporters. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 15289-15294. (37) Markovich, D. Physiological Roles and Regulation of Mammalian Sulfate Transporters. Physiol. Rev. 2001, 81, 1499-1533. (38) Addiscott, T. M.; Benjamin, N. Nitrate and human health. Soil Use and Management 2004, 20, 98-104. (39) Portelli, C. The role of the sodium, potassium, magnesium and calcium ions in the transfer of bioenergy, and the possibility of their substitution by other cations. Physiologie (Bucarest) 1977, 14, 43-46.

Page 21: and Philip A. Gale*

(40) Clarke, H. J.; Howe, E. N. W.; Wu, X.; Sommer, F.; Yano, M.; Light, M. E.; Kubik, S.; Gale, P. A. Transmembrane Fluoride Transport: Direct Measurement and Selectivity Studies. J. Am. Chem. Soc. 2016. (41) Marcus, Y. Thermodynamics of solvation of ions. Part 5.-Gibbs free energy of hydration at 298.15 K. J. Chem. Soc., Faraday Trans. 1991, 87, 2995-2999. (42) Busschaert, N.; Park, S.-H.; Baek, K.-H.; Choi, Y. P.; Park, J.; Howe, E. N. W.; Hiscock, J. R.; Karagiannidis, L. E.; Marques, I.; Félix, V.; Namkung, W.; Sessler, J. L.; Gale, P. A.; Shin, I. A synthetic ion transporter that disrupts autophagy and induces apoptosis by perturbing cellular chloride concentrations. Nat. Chem. 2017, 9, 667-675. (43) Busschaert, N.; Karagiannidis, L. E.; Wenzel, M.; Haynes, C. J. E.; Wells, N. J.; Young, P. G.; Makuc, D.; Plavec, J.; Jolliffe, K. A.; Gale, P. A. Synthetic transporters for sulfate: a new method for the direct detection of lipid bilayer sulfate transport. Chem. Sci. 2014, 5, 1118-1127. (44) Yano, M.; Tong, C. C.; Light, M. E.; Schmidtchen, F. P.; Gale, P. A. Calix[4]pyrrole-based anion transporters with tuneable transport properties. Org. Biomol. Chem. 2010, 8, 4356-4363. (45) Berry, S. N.; Soto-Cerrato, V.; Howe, E. N. W.; Clarke, H. J.; Mistry, I.; Tavassoli, A.; Chang, Y.-T.; Perez-Tomas, R.; Gale, P. A. Fluorescent transmembrane anion transporters: shedding light on anionophoric activity in cells. Chem. Sci. 2016, 7, 5069-5077. (46) Hill, A. V. The Combinations of Haemoglobin with Oxygen and with Carbon Monoxide. I. Biochem. J. 1913, 7, 471-480. (47) Clarke, H. J.; Van Rossom, W.; Horton, P. N.; Light, M. E.; Gale, P. A. Anion transport and binding properties of N Nʹ-(phenylmethylene)dibenzamide based receptors. Supramol. Chem. 2016, 28, 10-17. (48) Haynes, C. J. E.; Busschaert, N.; Kirby, I. L.; Herniman, J.; Light, M. E.; Wells, N. J.; Marques, I.; Felix, V.; Gale, P. A. Acylthioureas as anion transporters: the effect of intramolecular hydrogen bonding. Org. Biomol. Chem. 2014, 12, 62-72. (49) Spooner, M. J.; Gale, P. A. Anion transport across varying lipid membranes - the effect of lipophilicity. Chem. Commun. 2015, 51, 4883-4886. (50) Hussain, S.; Brotherhood, P. R.; Judd, L. W.; Davis, A. P. Diaxial Diureido Decalins as Compact, Efficient, and Tunable Anion Transporters. J. Am. Chem. Soc. 2011, 133, 1614-1617. (51) Lisbjerg, M.; Valkenier, H.; Jessen, B. M.; Al-Kerdi, H.; Davis, A. P.; Pittelkow, M. Biotin[6]uril Esters: Chloride-Selective Transmembrane Anion Carriers Employing C—H···Anion Interactions. J. Am. Chem. Soc. 2015, 137, 4948-4951. (52) Dias, C. M.; Li, H.; Valkenier, H.; Karagiannidis, L. E.; Gale, P. A.; Sheppard, D. N.; Davis, A. P. Anion transport by ortho-phenylene bis-ureas across cell and vesicle membranes. Org. Biomol. Chem. 2018, 16, 1083-1087. (53) Sakai, N.; Matile, S. The determination of the ion selectivity of synthetic ion channels and pores in vesicles. J. Phys. Org. Chem. 2006, 19, 452-460. (54) Jentzsch, A. V.; Emery, D.; Mareda, J.; Nayak, S. K.; Metrangolo, P.; Resnati, G.; Sakai, N.; Matile, S. Transmembrane anion transport mediated by halogen-bond donors. Nat. Commun. 2012, 3, 905. (55) Macchione, M.; Tsemperouli, M.; Goujon, A.; Mallia, A. R.; Sakai, N.; Sugihara, K.; Matile, S. Mechanosensitive Oligodithienothiophenes: Transmembrane Anion Transport Along Chalcogen-Bonding Cascades. Helv. Chim. Acta 2018, 101, e1800014. (56) Roy, A.; Saha, D.; Mukherjee, A.; Talukdar, P. One-Pot Synthesis and Transmembrane Chloride Transport Properties of C3-Symmetric Benzoxazine Urea. Org. Lett. 2016, 18, 5864-5867. (57) Roy, A.; Biswas, O.; Talukdar, P. Bis(sulfonamide) transmembrane carriers allow pH-gated inversion of ion selectivity. Chem. Commun. 2017, 53, 3122-3125. (58) Roy, A.; Gautam, A.; Malla, J. A.; Sarkar, S.; Mukherjee, A.; Talukdar, P. Self-assembly of small-molecule fumaramides allows transmembrane chloride channel formation. Chem. Commun. 2018, 54, 2024-2027.

Page 22: and Philip A. Gale*

(59) Avanti Polar Lipids Inc. https://avantilipids.com/product/840051 (accessed 09/08/2018 2018). (60) Avanti Polar Lipids Inc. https://avantilipids.com/product/850457 (accessed 10/08/2018 2018). (61) Wu, X.; Gale, P. A. Small-Molecule Uncoupling Protein Mimics: Synthetic Anion Receptors as Fatty Acid-Activated Proton Transporters. J. Am. Chem. Soc. 2016, 138, 16508-16514. (62) Fischer, R. B. Ion-selective electrodes. J. Chem. Educ. 1974, 51, 387. (63) De Marco, R.; Clarke, G.; Pejcic, B. Ion-Selective Electrode Potentiometry in Environmental Analysis. Electroanalysis 2007, 19, 1987-2001. (64) Covington, A. K.: Ion Selective Electrode Method; CRC Press: Boca Raton, 1979. (65) Thermo Fisher Scientific Inc. http://www.thermofishersci.in/lit/Fisherbrand%20Accumet%20Electrodes%20Handbook.pdf (accessed 23/11/2018 2018). (66) Thermo Fisher Scientific Inc. https://www.fishersci.com/shop/products/thermo-scientific-orion-chloride-combination-electrode-mercury-free-chloride-combination/13620627 (accessed 28/11/2018 2018). (67) Thermo Fisher Scientific Inc. https://assets.thermofisher.com/TFS-Assets/LSG/manuals/D12671~.pdf (accessed 22/11/2018 2018). (68) Mettler Toledo. https://www.mt.com/dam/non-indexed/po/ana/titration/ApplGuide_ISE_Cl_30253765_V04.15.pdf (accessed 23/11/2018 2018). (69) Thermo Fisher Scientific Inc. https://www.fishersci.co.uk/shop/products/orion-star-a211-ph-benchtop-meter/p-4529651 (accessed 27/11/2018 2018). (70) Avanti Polar Lipids Inc. https://avantilipids.com/divisions/equipment-products/mini-extruder-assembly-instructions/ (accessed 17/08/2018 2018).

Table of contents figure