Subscriber access provided by University of Winnipeg Library
Article
Lipid-based inhibitors act directly on GlyT2 Alexandra Schumann-Gillett, and Megan L O'Mara ACS Chem. Neurosci., Just Accepted Manuscript • DOI: 10.1021/acschemneuro.8b00586 • Publication Date (Web): 05 Dec 2018 Downloaded from http://pubs.acs.org on December 10, 2018
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
1
Lipid-based inhibitors act directly on GlyT2
2
Alexandra Schumann-Gillett1 and Megan L. O’Mara1*
3 4
1Research
5
2601, Australia
6
*
[email protected] (corresponding author)
School of Chemistry, The Australian National University, Canberra ACT
7 8
ACS Paragon Plus Environment
1
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
9
Page 2 of 37
Abstract
10 11
The endogenous lipids N-arachidonylglycine and oleoyl-L-carnitine, are potential
12
therapeutic leads in the treatment of chronic pain through their inhibition of the
13
glycine transporter GlyT2. However, their mechanism of action is unknown. It has
14
been hypothesised that these “bioactive” lipids either inhibit GlyT2 indirectly, by
15
significantly perturbing the biophysical properties of the membrane; or directly, by
16
binding directly to the transporter (either from a membrane-exposed or solvent-
17
exposed binding site). Here, we used molecular dynamics simulations to study the
18
effects of the lipids anandamide, N-arachidonylglycine and oleoyl-L-carnitine on a)
19
the biophysical properties of the bilayer, and b) direct binding interactions with
20
GlyT2. During the simulations, the biophysical properties of the bilayer itself—for
21
example the area per lipid, bilayer thickness and order parameters—were not
22
significantly altered by the presence or type of bioactive lipid, regardless of the
23
presence of GlyT2. Our work, together with previous computational and experimental
24
data, suggests that these acyl-inhibitors of GlyT2 inhibit the transporter by directly
25
binding to it. However, these bioactive lipids bound to various parts of GlyT2 and did
26
not prefer a single binding site during 4.5 μs of simulation. We postulate that the
27
binding site is located at the solvent-exposed regions of GlyT2. Understanding the
28
mechanism of action of these, and related bioactive lipids is essential in effectively
29
developing high-affinity GlyT2 inhibitors for the treatment of pain.
30 31
Abbreviations
32 ATB
Automated topology builder
EL4
Extracellular loop 4
GlyT2
Glycine transporter 2
MD
Molecular dynamics
PC
Phosphatidylcholine
PE
Phosphatidylethanolamine
POPC
1-Palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine
SLC6
Solute carrier 6
33 34
Keywords
ACS Paragon Plus Environment
2
Page 3 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
35 36
Anandamide; GlyT2; lipid inhibitor; molecular dynamics; N-arachidonylglycine;
37
oleoyl-L-carnitine; symporter
38
ACS Paragon Plus Environment
3
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
39
Page 4 of 37
Introduction
40 41
Glycine is an inhibitory neurotransmitter that is important in regulating pain sensory
42
neurons 1. The glycine transporter, GlyT2, regulates the concentration of glycine in
43
the synapse by the transport of glycine from the synaptic cleft into the pre-synaptic
44
neuron. GlyT2 activity can be altered by signalling pathways that influence GlyT2
45
expression levels, or via the action of inhibitors. For example, GlyT2 expression is
46
modulated by glycogen synthase kinase 3 2, neuronal plasma membrane Ca2+-ATPase
47
(isoforms 2 and 3) and Na+/Ca2+-exchanger 1 3, and Na+/K+-ATPase 4 . The inhibition
48
of GlyT2 has become the target for lipid-based inhibitors to treat chronic pain 5.
49
Glycine transport by GlyT2 is reversibly and non-competitively inhibited by N-
50
arachidonylglycine (IC50 5.1 μM)
51
Transport is inhibited up to 15-fold more by oleoyl-L-carnitine (IC50 340 nM) 9; a
52
lipid that is widely distributed throughout the body. In contrast, the structurally
53
related lipid and precursor to N-arachidonylglycine, anandamide, has no measurable
54
effects on GlyT2 function 6. Figure 1 shows the chemical structure of anandamide, N-
55
arachidonylglycine and oleoyl-L-carnitine. Anandamide and N-arachidonylglycine
56
differ only in their head groups, where anandamide has an aminoglycol group and N-
57
arachadonylglycine has a glycine group. Oleoyl-L-carnitine has a similar overall
58
structure (acyl chain with a polar head group) but both a unique head group
59
(carnitine) and acyl chain (monounsaturated, 18-carbon length). N-arachidonylglycine
60
reduces pain in rodent models, with minimal side effects 10,11.
6,7;
a lipid present mainly in the spinal cord 8.
61 62
Two main theories have been posed to explain how N-arachidonylglycine and oleoyl-
63
L-carnitine
64
biophysical properties of the bilayer, or directly, by binding to a specific binding site
65
that is accessible via the membrane or the extracellular solution. For example, as
66
acylcarnitines can increase the permeability of the membrane to drugs, it has been
67
suggested that they may perturb the lipid order in the bilayer
68
influence the transport characteristics of the embedded membrane transporter
69
Moreover, endocannabinoids have been associated with lipid raft formation, which
70
could impact the function of neighbouring proteins
71
studies have shown that the addition of fatty acids to model bilayers or liposomes
72
alter the biophysical properties of the membrane, such as the phospholipid phase,
inhibit GlyT2. They could act on GlyT2 indirectly, by perturbing the
ACS Paragon Plus Environment
14–17.
9,12,
which in turn may 13.
Previous experimental
4
Page 5 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
73
fluidity and thickness, however the precise nature of these changes depend on the acyl
74
tail length and saturation 18.
75 76
While there is a great deal of supporting evidence for a direct and specific binding
77
interaction of lipid-based inhibitors with GlyT2, the mechanism of the interaction is
78
unclear. Oleoyl-L-carnitine is a relatively potent inhibitor of GlyT2 (IC50 of 340 nM)
79
9,
80
suggestions from experiments are that a) oleoyl-L-carnitine interacts with a
81
membrane-exposed site on GlyT2
82
interactions with solvent-exposed extracellular loop 4 (EL4). The so-called lipid
83
phase interaction of oleoyl-L-carnitine with GlyT2 has been explored using β-
84
cyclodextrin; a cyclic sugar oligosaccharide that removes lipids such as cholesterol
85
from cell cultures. In experiments, the binding of oleoyl-L-carnitine to GlyT2 is
86
essentially irreversible. However when β-cyclodextrin is added to the preparation,
87
oleoyl-L-carnitine binding is reversible. Because β-cyclodextrin removes lipids from
88
bilayers, it has been suggested that the oleoyl-L-carnitine binding site on GlyT2 may
89
be exposed to the lipid membrane. The cell membrane has also been postulated to
90
form a reservoir in which oleoyl-L-carnitine can accumulate 9. Previous investigations
91
have shown that anandamide binds to the G protein-coupled cannabinoid receptor
92
CB1 via the membrane 19,20, and there is evidence that this binding mechanism is used
93
by other highly lipophilic molecules to G-protein coupled receptors and ion channels
94
21–23.
95
potent GlyT2 inhibitors like oleoyl-L-carnitine to that of common membrane
96
phosphatidylcholine (PC) and phosphatidylethanolamine (PE) lipids, it has been
97
hypothesised that N-arachidonylglycine, oleoyl-L-carnitine and other inhibitors may
98
inhibit GlyT2 by first inserting into the membrane before associating with GlyT2 5.
that displays intriguing behaviour on the transporter. Two seemingly conflicting 5,9;
and b) oleoyl-L-carnitine inhibits GlyT2 via
Based on these previous studies, and the similarity of the unsaturated C18 tail of
99 100
However, mutagenesis data suggests that the activity of oleoyl-L-carnitine and N-
101
arachidonylglycine on GlyT2 also depends on interactions with specific extracellular
102
protein residues, particularly I545, located between extracellular loop 4 (EL4)a and
103
EL4b. The I545L mutation significantly reduced oleoyl-L-carnitine and N-
104
arachidonylglycine inhibition of GlyT2. It has been proposed that oleoyl-L-carnitine
105
and N-arachidonylglycine directly interact with EL4 to inhibit the conformational
106
movements required for transport 9. Independent docking studies suggest that
ACS Paragon Plus Environment
5
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 6 of 37
107
anandamide and the related lipid, 2-arachidonoyl glycerol, bind to the cannabinoid
108
receptor CB1 between the extracellular loops
109
importance in mediating protein/ligand interactions.
24,
further demonstrating their
110 111
Whether lipid-based inhibitors act on GlyT2 indirectly by changing the biophysical
112
properties of the membrane in which the transporter is embedded, or directly by
113
binding to a specific location on GlyT2 is still a matter of considerable debate. To this
114
end, we performed atomistic molecular dynamics simulations in which our
115
experimentally characterised GlyT2 model 25 was embedded in bilayers that contained
116
1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) and cholesterol, and
117
either anandamide (negligibly active), N-arachidonylglycine (active) or oleoyl-L-
118
carnitine (potent), as illustrated in Figure 2. We show how these “bioactive” lipids
119
affect the biophysical properties of the bilayer in which they are embedded, with and
120
without GlyT2; and we present their membrane-accessible binding interactions with
121
GlyT2 embedded in bilayer containing POPC and cholesterol.
122 123
Results and discussion
124 125
With the emergence of a new class of effective acyl GlyT2 inhibitors, there has been
126
speculation as to how these drugs elicit inhibition on GlyT2. Theories include a)
127
indirect action on GlyT2 by altering the environment (membrane) in which GlyT2 is
128
embedded; b) direct binding to GlyT2 from the membrane; and c) direct binding to
129
solvent-exposed areas of GlyT2. Here, we performed 12 μs of atomistic molecular
130
dynamics simulations of POPC bilayers enriched with 20 mol % cholesterol, lacking
131
and containing membrane embedded GlyT2 and 5 mol % of each of the following
132
bioactive lipids: anandamide, N-arachidonylglycine and oleoyl-L-carnitine, to address
133
theories a) and b).
134 135
Membrane-embedded anandamide, N-arachidonylglycine and oleoyl-L-carnitine
136
have little effect on the biophysical properties of the bilayer
137 138
The bilayer thickness and POPC order parameter were not significantly altered in the
139
simulations containing or lacking the bioactive lipids, or GlyT2, as shown in Figure 3
140
and Table 1. The thickness of the control POPC/cholesterol bilayer remained
ACS Paragon Plus Environment
6
Page 7 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
141
unchanged in the presence (42.8 ± 0.2 Å) or absence (42.8 ± 0.5 Å) of GlyT2. The
142
thickness of POPC bilayers enriched with 20 mol % cholesterol is 43 Å in other
143
simulations 26 and in experiments (containing 20 mol % cholesterol and a
144
phospholipid similar to POPC, SOPC)
145
arachidonylglycine or oleoyl-L-carnitine into the bilayers did not substantially alter
146
the bilayer thickness, as shown in Figure 3A and Table 1. Likewise, the addition of
147
GlyT2 or 5 mol % of any one of the bioactive lipids had an almost negligible effect on
148
the ordering of the POPC tails, as shown in Figure 3C–D. In the absence of GlyT2,
149
there was a slight increase in the order parameter for the bilayers containing 5 %
150
bioactive lipid (maximum increase of approximately 0.001 units between N-
151
arachidonylglycine and control bilayer lacking GlyT2 for carbon 9 of sn-1 tail). The
152
POPC order parameter was not altered by adding 5 mol % bioactive lipid to the
153
bilayers containing GlyT2. The order parameters for the control system are similar to
154
those obtained previously 28.
27.
Incorporating anandamide, N-
155 156
In the absence of GlyT2, the mean area per POPC of the control POPC/cholesterol
157
bilayer was 53.0 ± 0.7 Å2. In comparison, the mean area per POPC decreased in all
158
cases when 5 % bioactive lipid was added, to give values between 49.2 ± 0.8 Å2 and
159
51.2 ± 0.6 Å2 as shown in Table 1 and Figure 3B. When GlyT2 was present in the
160
membrane, the mean area per POPC increased by 3.5 ± 0.9 Å2. The mean area per
161
lipid of the control POPC/cholesterol bilayer increased from 53.0 ± 0.7 Å2 to
162
55.2 ± 0.9 Å2 when GlyT2 was present. Table 1 shows similar trends in the presence
163
of GlyT2 for the bilayers containing 5 mol % anandamide, N-arachidonylglycine and
164
oleoyl-L-carnitine. The average experimental area per lipid for SOPC, which has the
165
same head group as POPC, in a bilayer containing 20 mol % cholesterol is
166
approximately 56 Å2 27.
167 168
It should be noted that the bilayers without GlyT2 contained fewer lipids than the
169
bilayers with embedded GlyT2, as shown in Table 2. Previous studies have suggested
170
that the increase in the area per POPC molecule that we observed when GlyT2 was
171
present could be due to finite size effects 29. To ensure that finite size effects were not
172
influencing the mean area per POPC observed in the absence of GlyT2, we simulated
173
a larger bilayer that contained POPC, cholesterol and 5 mol % anandamide (without
174
GlyT2) that had the same box dimensions as the bilayer containing GlyT2. The mean
ACS Paragon Plus Environment
7
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 8 of 37
175
area per POPC for the large anandamide-containing bilayer that lacked GlyT2 was
176
51.0 Å ± 0.4 Å (c.f. 50.5 ± 0.6 Å for the smaller bilayer) and its thickness was
177
42.9 ± 0.1 Å (c.f. 43.0 ± 0.3 Å for the smaller bilayer). The order parameters for
178
POPC in the larger anandamide-containing bilayer are shown in Supporting
179
Information Figure S1. There is no difference between the order parameters in the
180
smaller bilayer and the larger bilayer. The area per POPC and thickness of the larger
181
bilayer are within error for the smaller GlyT2-lacking bilayer (Table 1), therefore
182
finite size effects are unlikely to have impacted our results. Simulations of the lipid 2-
183
hydroxyoleic acid, which exhibits anti-tumour activity, embedded in a POPC bilayer
184
show that incorporating it into a bilayer has little effect on the area per lipid and
185
bilayer thickness (on the order of tenths of Å2 or Å) 30. We did not observe any lipid
186
raft formation in our simulations. Our data suggest that anandamide, N-
187
arachidolylglycine and oleoyl-L-carnitine do not alter the biophysical properties of the
188
bilayer. Recent experimental data support this finding, for example membrane
189
capacitance measurements show that insertion of endocannabanoids has limited effect
190
on the thickness of phospholipid bilayers 31.
191 192 193
Anandamide snorkelled into the bilayer but N-arachidonylglycine and oleoyl-L-
194
carnitine remained at the lipid/water interface
195 196
During MD simulations, anandamide “snorkelled” into the centre of the bilayer
197
regardless of the presence of GlyT2. This behaviour is reflected in the normalised
198
number density for the location of the head group atoms of each bioactive lipid and
199
GlyT2, shown in Figure 4A–B. At the start of the simulations, all lipids, including the
200
bioactive lipids, were embedded into the bilayer with their head group at the
201
lipid/water interface and their tail group elongated approximately parallel to the
202
bilayer normal. The head group atoms chosen (the phosphorus atom of POPC, the
203
oxygen of cholesterol and the nitrogen of the bioactive lipids) remained at the
204
lipid/water interface for all lipids except anandamide. As illustrated in Figure 4C,
205
anandamide was able to “snorkel” into the membrane and interact with GlyT2 near
206
the membrane-embedded TM helices. In the absence of GlyT2, the normalised
207
number density when the anandamide head groups were in the centre of the bilayer
208
was 0.2, and when GlyT2 was present it was 0.3. The interactions between
ACS Paragon Plus Environment
8
Page 9 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
209
anandamide and GlyT2 were largely transient and occurred on the scale of tens of
210
nanoseconds.
211 212
To quantify the flexibility of the bioactive lipids in each membrane, we measured
213
their start-to-end intramolecular distance, shown in Figure 5A. In bilayers lacking
214
GlyT2, this distance was 18–21 Å for N-arachidonylglycine and oleoyl-L-carnitine,
215
and 14–18 Å for anandamide. The shorter distance for anandamide was due to the
216
acyl chain folding over on itself, as observed in previous studies
217
conformation facilitated the removal of the anandamide head group from the
218
lipid/water interface, allowing it to “snorkel” to the centre of the bilayer. In contrast,
219
the head groups of N-arachidonylglycine and oleoyl-L-carnitine remained at the
220
lipid/water interface (Figure 4) and these lipids adopted an elongated confirmation as
221
in Figure 5C. The intra-molecular flexibility of anandamide and N-arachidonylglycine
222
was largely unaltered when GlyT2 was present or absent in the bilayer, as shown in
223
Figure 5B (with some outliers in the spread of the data due to a larger sample size
224
when GlyT2 was present). However, the intra-molecular flexibility of oleoyl-L-
225
carnitine increased to 19–23 Å when GlyT2 was present, with the majority of the data
226
clustered at approximately 22 Å. Therefore, oleoyl-L-carnitine displayed an extended
227
conformation, at least as elongated as that shown in Figure 5C, in the bilayer that
228
contained GlyT2.
32.
This folded
229 230 231
Anandamide, N-arachidonylglycine and oleoyl-L-carnitine do not display a
232
conserved binding site on GlyT2
233 234
In light of our above findings that the bioactive lipids studied here do not significantly
235
alter the biophysical properties of the membrane, we hypothesised that these bioactive
236
lipids could act on GlyT2 by directly binding to it from the membrane. To assess the
237
structural stability of GlyT2 embedded in our membrane systems, we measured its
238
backbone root-mean-square deviation. During the during the first 25 ns of all
239
simulations, GlyT2 adjusted to the modified lipid bilayer and adopted a conformation
240
that was within 2–4 Å of the starting structure, then plateaued, as shown in
241
Supporting Information Figure S2, indicating that the conformation of GlyT2 was
242
stable in all simulations.
ACS Paragon Plus Environment
9
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 10 of 37
243 244
To examine the interactions between the membrane-embedded bioactive lipids and
245
GlyT2, we recorded the regions of GlyT2 that the bioactive lipids interacted with
246
during the final 400 ns of each simulation replicate. Of the three bioactive lipids,
247
anandamide bound to GlyT2 the most frequently (Supporting Information Figure
248
S3A), followed by N-arachidonylglycine (Supporting Information Figure S3B), then
249
oleoyl-L-carnitine (Supporting Information Figure S3C). In fact, anandamide bound to
250
GlyT2 almost eight times more frequently than oleoyl-L-carnitine, and N-
251
arachidonylglycine bound to GlyT2 almost three times more than oleoyl-L-carnitine.
252 253
The binding locations in Supporting Information Figure S3 show little conservation
254
between lipid type, but some conservation between replicate simulations. To identify
255
any conserved binding locations on GlyT2, we combined the replicate binding data
256
for each bioactive lipid and averaged it to calculate the binding propensity for each
257
GlyT2 residue. Figure 6 shows the binding propensity data for anandamide (Figure
258
6A), N-arachidonylglycine (Figure 6B) and oleoyl-L-carnitine (Figure 6C), coloured
259
in greyscale and mapped onto the structure of GlyT2. Regions of GlyT2 that bioactive
260
lipids did not associate with are coloured by helix. Supporting Information Figure S3
261
and Figure 6 show that none of the lipids consistently bound to GlyT2 in the solvent-
262
exposed regions (i.e. the light salmon coloured parts of GlyT2 in Figure 6).
263
Furthermore, anandamide and N-arachidonylglycine consistently bound to different
264
regions of GlyT2, with little or no overlap in the binding regions. For example,
265
anandamide displayed widespread adsorption to all TM helices and consistently
266
bound to the extracellular region of TM9 (Figure 6A; right side view), whereas N-
267
arachidonylglycine never bound there. Although N-arachidonylglycine remained in
268
the bilayer, oriented with its head group at the lipid/water interface (c.f. atom density
269
data and Figure 4), it bound to the extracellular portion of TM5 and the intracellular
270
portion of TM1 during all simulations (Figure 6B). Oleoyl-L-carnitine—the most
271
experimentally potent bioactive lipid studied here—did not consistently bind to any
272
specific regions of GlyT2 within the timeframe of the simulations.
273 274
Previous studies have found that anandamide does not perturb the properties of the
275
bilayer 19,31 and interacts specifically with the G protein-coupled cannabinoid receptor
276
CB1 at a membrane-exposed site 19. Intriguingly, although the same number of each
ACS Paragon Plus Environment
10
Page 11 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
277
of the three bioactive lipids were used in the simulations, and all three bound to the
278
GlyT2 transmembrane domains, the most potent inhibitor, oleoyl-L-carnitine,
279
interacted the least with GlyT2 and did not display a conserved binding site between
280
the three simulation replicates (see Supporting Information Figure S3C and Figure
281
6C). This could simply be due to lipid diffusion or insufficient sampling; however the
282
concentration of bioactive lipid used in each simulation was an order of magnitude
283
larger than the half maximal inhibitory concentration for oleoyl-L-carnitine on GlyT2.
284
Our simulations contained 5 mol % bioactive lipid embedded in the bilayer; a
285
concentration of approximately 3000 nM of inhibitor with respect to the entire
286
solute/solvent system. The experimental IC50, for oleoyl-L-carnitine is 340 nM 5,9. The
287
relatively large concentration could explain why anandamide, which does not inhibit
288
GlyT2 6, adsorbed all over transmembrane domains, and at conserved locations
289
between the replicates, for example at the extracellular portion of TM9 and at residues
290
on TM10, TM11 and TM12. The number of binding interactions between anandamide
291
and GlyT2 could also be attributed to the flexible “snorkelling” behaviour
292
anandamide displayed, where it folded over and moved in the centre of the bilayer as
293
shown in Figure 4C (c.f. Figures 4A–B and 5). Alternatively, the behaviour displayed
294
by anandamide could be considered to be similar to that of POPC, with widespread
295
non-specific associations. Conversely, oleoyl-L-carnitine did not display many non-
296
specific associations, but did not show specific binding behaviour in the timescale of
297
the simulations.
298 299
Compared to anandamide, the GlyT2 inhibitors N-arachidonylglycine and oleoyl-L-
300
carnitine remained in an elongated conformation (see Figure 5) and their head groups
301
continually resided at the lipid/water interface (see Figure 4). Polyunsaturated fatty
302
acids are relatively flexible and can form folded structures, whereas monounsaturated
303
fatty acids adopt elongated structures with a local kink at the site of unsaturation
304
During all simulations, N-arachidonylglycine, which inhibits GlyT2 15-fold less than
305
oleoyl-L-carnitine, bound to the extracellular portion of TM1 and the intracellular
306
portion of TM5 (Figure 6B, greyscale colouring) while oleoyl-L-carnitine bound
307
three-fold less than N-arachidonylglycine and did not bind to a conserved location on
308
GlyT2. While GlyT2 can be modulated by the presence of lipid raft formation 34 and
309
neighbouring proteins 2–4, recent experimental data suggest that the head group of a
310
lipid inhibitor could determine the stability of GlyT2 in the bilayer 5. The consistent
ACS Paragon Plus Environment
33.
11
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 12 of 37
311
extracellular interactions between N-arachidonylglycine and the extracellular portions
312
of TM1 and TM5 support the postulation by Carland and co-workers that the head
313
groups of N-arachidonylglycine and oleoyl-L-carnitine could interact with GlyT2 at
314
the extracellular lipid/water interface of the TM domains, and influence the
315
movement of EL4 during transport 9.
316 317
There is strong evidence that acyl inhibitors bind to GlyT2 from solution. EL2 and
318
EL4 are required for N-arachidonylglycine to inhibit GlyT2 7, and EL4 is essential for
319
oleoyl-L-carnitine to inhibit GlyT2 9. Figure 6 shows EL4 (shaded in salmon) is
320
located next to the extracellular region of TM5 (cyan) where N-arachidonylglycine
321
consistently bound. In fact, N-arachidonylglycine bound at the extracellular loops of
322
the transmembrane lysophosphatidic acid receptor 5 (formerly known as GPR92) in a
323
study that combined docking with site-directed mutagenesis 35. Docking studies have
324
also suggested that anandamide and related lipids bind near the extracellular loops of
325
cannabinoid receptors 24,36 and ion channels 37.
326 327
We have shown that, at concentrations one order of magnitude higher than those used
328
experimentally, the incorporation of anandamide, N-arachidonylglycine and oleoyl-L-
329
carnitine into the bilayer does not significantly perturb the biophysical properties of
330
the bilayer. Our results also show that membrane-embedded anandamide, N-
331
arachidonylglycine and oleoyl-L-carnitine do not interact with GlyT2 at a specific
332
location that is accessible from within the membrane within a 500 ns timeframe. We
333
therefore hypothesise that the bioactive lipid binding site of GlyT2 may not be
334
accessible from the hydrophobic core of the membrane. We are currently
335
investigating the binding patterns of other lipid-based inhibitors to GlyT2 to
336
determine whether uptake occurs from solution or from the lipid/water interface. By
337
identifying the mechanism of action of these bioactive lipid inhibitors of GlyT2, we
338
move closer to developing targeted, selective and reversible treatments for chronic
339
pain.
340 341 342
Methods
343
Molecular dynamics simulations
344
ACS Paragon Plus Environment
12
Page 13 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
345
We simulated bilayers that contained POPC and cholesterol, and added bioactive
346
lipids and GlyT2. The specific compositions of the bilayers, expressed as molar ratios,
347
were approximately: 5 % anandamide with 20 % cholesterol and 75 % POPC, 5 % N-
348
arachidonylglycine with 20 % cholesterol and 75 % POPC, and 5 % oleoyl-L-
349
carnitine with 20 % cholesterol and 75 % POPC. As controls, we simulated a bilayer
350
containing GlyT2, 20 % cholesterol and 80 % POPC, and these bilayers without
351
GlyT2. The number of lipids in each bilayer is given in Table 2. Note that we
352
simulated the bilayers without GlyT2 for 0.5 μs first, and then used the coordinates
353
from the end of a 0.5 μs simulation of each bilayer type to build the bilayers that
354
contained GlyT2. Snapshots of each bilayer, lacking and containing GlyT2, are shown
355
in Figure 2.
356 357 358
Molecular coordinates and topologies
359 360
The coordinates of our experimentally validated homology model of GlyT2 (in the
361
outward-occluded conformation) were taken from Subramanian et al. 25. To determine
362
whether inhibitor interactions were competitive or non-competitive with the glycine
363
substrate, all simulations were performed on apo GlyT2, in the absence of substrate or
364
bound ions. The united atom coordinates and parameters for anandamide, N-
365
arachidonylglycine and oleoyl-L-carnitine were developed using the Automated
366
Topology Builder and Repository (ATB)
367
available for download from the ATB (anandamide MoleculeID: 250693, N-
368
arachidonylglycine MoleculeID: 250689, oleoyl-L-carnitine MoleculeID: 296337).
369
Note that the force constant associated with cis double bonds that the ATB produced
370
was routinely low enough to allow for cis/trans isomerisation around the double bond
371
at room temperature—a process not found under biological conditions. To prevent
372
non-physiological cis/trans isomerisation, the the force constants of all double bonds
373
in the ATB parameters for anandamide, N-arachidonylglycine and oleoyl-L-carnitine
374
were altered from their original value of 5.86 kJ/mol/rad2 to a revised value of 41.80
375
kJ/mol/rad2. To ensure that the cis double bond conformations were maintained, each
376
lipid was simulated for 1 ns in a box of water prior to incorporating it into a bilayer.
377
The POPC parameters were taken from Poger and Mark
378
parameters were taken from the ATB 38–40. As classical MD simulations implement a
38–40.
The coordinates and topologies are
ACS Paragon Plus Environment
41
and the cholesterol
13
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 14 of 37
379
fixed protonation state, rather than an assigned pH, all lipids were simulated in the
380
protonation state in which they would most likely be found at physiological pH
381
(pH 7). The pKa of the POPC phosphonium group is 1.86, and the pKa’s of the
382
carboxylic acid groups of N-arachidonylglycine and oleoyl-L-carnitine are 3.98 and
383
4.22, respectively, as calculated using the method implemented in ChemAxon 42.
384
Thus, POPC and oleoyl-L-carnitine were simulated as zwitterions and N-
385
arachidonylglycine was simulated in its deprotonated form, as shown in Figure 1. The
386
number of lipids in each bilayer is included in Table 2.
387 388
System Setup
389 390
The simulations were prepared and performed using version 2016.1 of the
391
GROMACS engine 43,44 with the GROMOS54A7 force field for proteins and lipids 44.
392
An equal number of bioactive lipids were incorporated into the upper and lower
393
leaflets of each respective bilayer system, and initially positioned so that they were
394
evenly distributed throughout the leaflet prior to energy minimisation and
395
equilibration. All bilayers were oriented in the x-y plane of a solvated rectangular box.
396
Water was simulated explicitly using the simple point charge water model 45. Na+ and
397
Cl- were included to neutralise the system, and 0.15 M of these ions was added to
398
mimic physiological conditions. Prior to production simulations, periodic boundary
399
conditions were applied and all systems were energy minimised using a steepest
400
descent algorithm.
401 402
All bilayers (lacking or containing GlyT2) underwent a 5 ns equilibration simulation
403
after the initial energy minimisation, to visually assess the coordinates of each system.
404
For the systems containing GlyT2, a series of five consecutive 1 ns position restrained
405
simulations were performed, where the positions of the heavy protein atoms were
406
restrained by a harmonic potential that was successively lowered. The force constant
407
of this potential was initially 1,000 kJ/mol/nm for the first position restrained
408
simulation, and was progressively decreased to 500 kJ/mol/nm, 100 kJ/mol/nm,
409
50 kJ/mol/nm and finally 10 kJ/mol/nm for the following four position restrained
410
simulations. The coordinates of each system after the initial 5 ns simulations
411
(unrestrained for bilayers without GlyT2; restrained for bilayers with GlyT2) were
ACS Paragon Plus Environment
14
Page 15 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
412
used as the starting conformation for three independent, 0.5 μs simulations per
413
system.
414 415
Simulation Details
416 417
The simulation conditions were consistent for all simulations, regardless of the
418
presence of GlyT2, and are documented by Carland and co-workers
419
velocities were assigned at the start of each simulation replicate. The simulations were
420
performed in the NPT ensemble, where the temperature was maintained by coupling
421
the bilayer (solute) and the solvent separately to an external temperature bath at
422
300 K. The Bussi-Donadio-Parrinello velocity rescale thermostat
423
with a coupling constant, τT = 0.1 ps. Similarly, the pressure was maintained at 1 bar
424
by weakly coupling the system to an external pressure bath using the Berendsen
425
thermostat. The pressure coupling was semi-isotropic, isotropic in the plane of the
426
bilayer (x-y). The isothermal compressibility was 4.5 × 10-5 bar and the coupling
427
constant was τT = 0.5 ps. The lengths of the covalent solute bonds were constrained
428
using the LINCS algorithm
429
constrained using the SETTLE algorithm
430
bonded interactions were updated every time step. The electrostatic interactions were
431
calculated using particle mesh Ewald summation and the Lennard-Jones interactions
432
were calculated with a 1.0 nm cut-off. Images of the systems were produced using the
433
Visual Molecular Dynamics software 50.
48
47
46.
Unique
was employed,
and the geometry of the water molecules was 49.
Both the electrostatic and the non-
434 435
Analysis
436 437
In each case, the first 100 ns of each simulation was treated as an equilibration period,
438
unless otherwise stated. The final 400 ns of each simulation (i.e. 100 ns–500 ns) were
439
used for data analysis. The output was processed so that 400 ns corresponded to 400
440
frames. All data were calculated for each of these frames unless otherwise stated, and
441
averages were calculated over all frames and replicates.
442 443
Bilayer thickness
444
ACS Paragon Plus Environment
15
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 16 of 37
445
We calculated the bilayer thickness using the FATSLiM software 51, and selected the
446
phosphorus atoms on POPC as the reference atoms. The program performs a
447
neighbour search and identifies neighbouring lipids, for each lipid, and iteratively
448
calculates the normal. A single lipid is chosen to serve as a modulator for the
449
reference position when its normal is within 10o of the reference (bilayer) normal.
450
This treatment is important for undulating membranes. An averaged position is
451
calculated based on the lipid’s neighbours and the local normal is used with this
452
averaged position to estimate the inter-leaflet distance. Once an averaged position is
453
calculated for one leaflet, a neighbour search is performed on the other leaflet, where
454
lipids close to the reference position and lipids belonging to the other leaflet are
455
identified. We used a neighbour search cut-off distance of 6 nm. Lipids are selected if
456
the distance vector between the neighbour and the reference position is within 10o of
457
the reference normal. Again, an averaged position is calculated. The bilayer thickness
458
is calculated using the distance vector between the two averaged positions. The mean
459
and standard deviation of the combined replicate trajectories (for 100 ns–500 ns) are
460
reported in Table 1.
461 462
Area per lipid
463 464
The FATSLiM software was used to calculate the area per lipid 51. First, a neighbour
465
search is performed for each lipid, using a 3 nm cut-off radius. The curvature within
466
this cut-off is assumed to be negligible. Neighbour lipids are projected onto the x-y
467
plane with the z-coordinate of the reference atom. These projected lipids are then used
468
to calculate the Voronoi cell. The lipid’s accessible area is approximated based on its
469
Voronoi cell. Importantly, FATSLiM takes membrane proteins into account—any
470
protein atoms that lie within the Voronoi cell are projected onto the same plane as the
471
lipids (x-y in our case), and the centre of geometry for those protein atoms inside the
472
Voronoi cell is calculated and removed from the Voronoi cell. An updated cell is used
473
to calculate the area of the lipid in question. The area per lipid was calculated for all
474
lipids in the system, and is reported for POPC for comparison to existing data. The
475
mean and standard deviation of the combined replicate trajectories, for 100 ns–500 ns,
476
are given in Table 1.
477 478
Deuterium order parameter
ACS Paragon Plus Environment
16
Page 17 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
479 480
We calculated the deuterium order parameter with the gmx order tool provided in the
481
GROMACS software. Note that this tool mistreats double bonds unless particular care
482
is taken
483
carbon atoms as recommended by Piggot and co-workers
484
previous work
485
according to equation 1:
486
𝑆𝐶𝐷 =
(1)
494 495
Lipid head group location
496 497
We determined the location of the lipid head groups by calculating the number
498
density of specific head group atoms for each lipid in the simulation box, and GlyT2
499
if it was present in the bilayer. After centring the bilayer, we performed the number
500
density calculation on the phosphorus atom of POPC, the oxygen of cholesterol and
501
the nitrogen of the bioactive lipids along the z-axis (perpendicular to the plane of the
502
membrane). We performed this calculation for each frame. Because our bilayers
503
contained a heterogeneous number of constituents, we normalised the data with
504
respect to unity so that we could clearly show the locations of the atoms on a single
505
axis.
506 507
Bioactive lipid conformation
508 509
To characterise the conformations that each bioactive lipid adopted in the bilayers, we
510
calculated the pairwise distance between a defined carbon atom on either end of each
511
bioactive lipid as shown in Figure 5C. We averaged the inter-atom distance data over
ACS Paragon Plus Environment
17
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 18 of 37
512
the final 400 ns of each simulation replicate for each bioactive lipid molecule, in each
513
system.
514 515
Root-mean-square deviation
516 517
We calculated the root-mean-square deviation (RMSD) of the backbone of GlyT2 by
518
first fitting the backbone to a reference structure, and then computing the deviation of
519
the backbone from the reference using the method of Maiorov and Crippen
520
reference structure was the first frame of the unrestrained simulation. The mean
521
RMSD and standard deviation (for 100 ns–500 ns) are given in Table 1.
56.
The
522 523
Lipid binding
524 525
To characterise how membrane-incorporated bioactive lipids, and cholesterol, bind to
526
GlyT2, we calculated a) the binding location of each lipid to GlyT2 over time, and b)
527
the binding propensity of each lipid to the structure of GlyT2. For a), we recorded
528
every time that any bioactive lipid was within 4 Å of a protein residue, for each frame
529
in each replicate. For b), we combined the binding data from a), averaged it and
530
calculated its proportion so that we had a single binding propensity value for each
531
GlyT2 residue. We did this for each type of bioactive lipid and used the binding
532
propensity data as beta values to map the data onto the structure of GlyT2.
533 534
ASSOCIATED CONTENT
535
Supporting Information
536
Additional figures are given as Supporting Information, including: order parameters
537
for larger bilayer lacking GlyT2 (Figure S1); the root-mean-square deviation of
538
GlyT2 (Figure S2); and bioactive lipid binding patterns over time (Figure S3).
539 540
AUTHOR INFORMATION
541
Corresponding Author
542
* Dr. Megan O’Mara
ACS Paragon Plus Environment
18
Page 19 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
543
* E-mail:
[email protected] 544
Present Addresses
545
1Research
546
2601, Australia.
547
Author Contributions
548
Conceived and designed the experiments: MLO and AS. Performed the experiments:
549
AS. Analysed the data: AS. Contributed reagents/materials/analysis tools: MLO.
550
Wrote the manuscript: AS and MLO.
551
Funding Sources
552
This work was supported by grants from the National Health and Medical Research
553
Council (APP1082570, APP1144429). AS is a recipient of a Westpac Future Leaders
554
Scholarship.
555
Notes
556
The authors declare no competing financial interest.
School of Chemistry, The Australian National University, Canberra ACT
557 558
ACKNOWLEDGMENT
559
The research was undertaken with the assistance of resources and services from the
560
National Computational Infrastructure (NCI), which is supported by the Australian
561
Government. We thank Prof Robert Vandenberg and Dr Shannon Mostyn for their
562
discussions and feedback.
563
ACS Paragon Plus Environment
19
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
564
Page 20 of 37
References
565 566
(1) Lynch, J. W. (2004) Molecular structure and function of the glycine receptor
567
chloride channel. Physiol. Rev. 84, 1051–1095.
568
(2) Jiménez, E., Núñez, E., Ibáñez, I., Zafra, F., Aragón, C., and Giménez, C. (2015)
569
Glycine transporters GlyT1 and GlyT2 are differentially modulated by glycogen
570
synthase kinase 3β. Neuropharmacology 89, 245–254.
571
(3) De Juan-Sanz, J., Núñez, E., Zafra, F., Berrocal, M., Corbacho, I., Ibáñez, I.,
572
Arribas-González, E., Marcos, D., López-Corcuera, B., Mata, A. M., and Aragón, C.
573
(2014) Presynaptic control of glycine transporter 2 (GlyT2) by physical and
574
functional association with plasma membrane Ca2+-ATPase (PMCA) and Na+-Ca2+
575
exchanger (NCX). J. Biol. Chem. 289, 34308–34324.
576
(4) de Juan-Sanz, J., Núñez, E., Villarejo-López, L., Pérez-Hernández, D., Rodriguez-
577
Fraticelli, A. E., López-Corcuera, B., Vázquez, J., and Aragón, C. (2013) Na+/K+-
578
ATPase is a new interacting partner for the neuronal glycine transporter GlyT2 that
579
downregulates its expression in vitro and in vivo. J. Neurosci. 33, 14269–14281.
580
(5) Mostyn, S. N., Carland, J. E., Shimmon, S., Ryan, R. M., Rawling, T., and
581
Vandenberg, R. J. (2017) Synthesis and characterization of novel acyl-glycine
582
inhibitors of GlyT2. ACS Chem. Neurosci. 8, 1949–1959.
583
(6) Wiles, A. L., Pearlman, R. J., Rosvall, M., Aubrey, K. R., and Vandenberg, R. J.
584
(2006) N-Arachidonyl-glycine inhibits the glycine transporter, GLYT2a. J.
585
Neurochem. 99, 781–786.
586
(7) Edington, A. R., McKinzie, A. A., Reynolds, A. J., Kassiou, M., Ryan, R. M., and
587
Vandenberg, R. J. (2009) Extracellular loops 2 and 4 of GLYT2 are required for N-
588
arachidonylglycine inhibition of glycine transport. J. Biol. Chem. 284, 36424–36430.
589
(8) Huang, S. M., Bisogno, T., Petros, T. J., Chang, S. Y., Zavitsanos, P. A., Zipkin,
590
R. E., Sivakumar, R., Coop, A., Maeda, D. Y., De Petrocellis, L., Burstein, S., Di
591
Marzo, V., and Walker, J. M. (2001) Identification of a new class of molecules, the
592
arachidonyl amino acids, and characterization of one member that inhibits pain. J.
593
Biol. Chem. 276, 42639–42644.
594
(9) Carland, J. E., Mansfield, R. E., Ryan, R. M., and Vandenberg, R. J. (2013)
595
Oleoyl-L-carnitine inhibits glycine transport by GlyT2. Br. J. Pharmacol. 168, 891–
596
902.
597
(10) Succar, R., Mitchell, V. A., and Vaughan, C. W. (2007) Actions of N-
ACS Paragon Plus Environment
20
Page 21 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
598
arachidonyl-glycine in a rat inflammatory pain model. Mol. Pain 3, 24.
599
(11) Vuong, L. A. Q., Mitchell, V. A., and Vaughan, C. W. (2008) Actions of N-
600
arachidonyl-glycine in a rat neuropathic pain model. Neuropharmacology 54, 189–
601
193.
602
(12) LeCluyse, E. L., Appel, L. E., and Sutton, S. C. (1991) Relationship between
603
drug absorption enhancing activity and membrane perturbing effects of acylcarnitines.
604
Pharm. Res. An Off. J. Am. Assoc. Pharm. Sci. 8, 84–87.
605
(13) Schumann-Gillett, A., Blyth, M. T., and O’Mara, M. L. (2018) Is protein
606
structure enough? A review of the role of lipids in SLC6 transporter function.
607
Neurosci. Lett. DOI:10.1016/j.neulet.2018.05.020.
608
(14) Bradshaw, H. B., Rimmerman, N., Hu, S. J., Benton, V. M., Stuart, J. M.,
609
Masuda, K., Cravatt, B. F., O’Dell, D. K., and Walker, J. M. (2009) The
610
endocannabinoid anandamide is a precursor for the signaling lipid N-arachidonoyl
611
glycine by two distinct pathways. BMC Biochem. 10.
612
(15) Rimmerman, N., Bradshaw, H. B., Kozela, E., Levy, R., Juknat, A., and Vogel,
613
Z. (2012) Compartmentalization of endocannabinoids into lipid rafts in a microglial
614
cell line devoid of caveolin-1. Br. J. Pharmacol. 165, 2436–2449.
615
(16) Dainese, E., Oddi, S., Bari, M., and Maccarrone, M. (2007) Modulation of the
616
endocannabinoid system by lipid rafts. Curr. Med. Chem. 14, 2702–2715.
617
(17) Barnett-Norris, J., Lynch, D., and Reggio, P. H. (2005) Lipids, lipid rafts and
618
caveolae: their importance for GPCR signaling and their centrality to the
619
endocannabinoid system. Life Sci. 77, 1625–1639.
620
(18) Ibarguren, M., López, D. J., and Escribá, P. V. (2014) The effect of natural and
621
synthetic fatty acids on membrane structure, microdomain organization, cellular
622
functions and human health. Biochim. Biophys. Acta - Biomembr. 1838, 1518–1528.
623
(19) Tian, X., Guo, J., Yao, F., Yang, D. P., and Makriyannis, A. (2005) The
624
conformation, location, and dynamic properties of the endocannabinoid ligand
625
anandamide in a membrane bilayer. J. Biol. Chem. 280, 29788–29795.
626
(20) Reggio, P. H. (2010) Endocannabinoid binding to the cannabinoid receptors:
627
what is known and what remains unknown. Curr. Med. Chem. 17, 1468–1486.
628
(21) Pei, Y., Mercier, R. W., Anday, J. K., Thakur, G. A., Zvonok, A. M., Hurst, D.,
629
Reggio, P. H., Janero, D. R., and Makriyannis, A. (2008) Ligand-binding architecture
630
of human CB2 cannabinoid receptor: evidence for receptor subtype-specific binding
631
motif and modeling GPCR activation. Chem. Biol. 15, 1207–1219.
ACS Paragon Plus Environment
21
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 22 of 37
632
(22) Hurst, D. P., Grossfield, A., Lynch, D. L., Feller, S., Romo, T. D., Gawrisch, K.,
633
Pitman, M. C., and Reggio, P. H. (2010) A lipid pathway for ligand binding is
634
necessary for a cannabinoid G protein-coupled receptor. J. Biol. Chem. 285, 17954–
635
17964.
636
(23) Stanley, N., Pardo, L., and Fabritiis, G. De. (2016) The pathway of ligand entry
637
from the membrane bilayer to a lipid G protein-coupled receptor. Sci. Rep. 6, 22639.
638
(24) Hua, T., Vemuri, K., Pu, M., Qu, L., Han, G. W., Wu, Y., Zhao, S., Shui, W., Li,
639
S., Korde, A., Laprairie, R. B., Stahl, E. L., Ho, J.-H., Zvonok, N., Zhou, H.,
640
Kufareva, I., Wu, B., Zhao, Q., Hanson, M. A., Bohn, L. M., Makriyannis, A.,
641
Stevens, R. C., and Liu, Z.-J. (2016) Crystal structure of the human cannabinoid
642
receptor CB 1. Cell 167, 750–762.
643
(25) Subramanian, N., Scopelliti, A. J., Carland, J. E., Ryan, R. M., O’Mara, M. L.,
644
and Vandenberg, R. J. (2016) Identification of a 3rd Na+ binding site of the glycine
645
transporter, GlyT2. PLoS One 11, DOI: 10.1371/journal.pone.0159896.
646
(26) Olsen, B. N., Bielska, A. A., Lee, T., Daily, M. D., Covey, D. F., Schlesinger, P.
647
H., Baker, N. A., and Ory, D. S. (2013) The structural basis of cholesterol
648
accessibility in membranes. Biophys. J. 105, 1838–1847.
649
(27) Hung, W.-C., Lee, M.-T., Chen, F.-Y., and Huang, H. W. (2007) The condensing
650
effect of cholesterol in lipid bilayers. Biophys. J. 92, 3960–7.
651
(28) Hofsäß, C., Lindahl, E., and Edholm, O. (2003) Molecular dynamics simulations
652
of phospholipid bilayers with cholesterol. Biophys. J. 84, 2192–2206.
653
(29) Klauda, J. B., Brooks, B. R., and Pastor, R. W. (2006) Dynamical motions of
654
lipids and a finite size effect in simulations of bilayers. J. Chem. Phys. 125, 144710.
655
(30) Piotto, S., Trapani, A., Bianchino, E., Ibarguren, M., López, D. J., Busquets, X.,
656
and Concilio, S. (2014) The effect of hydroxylated fatty acid-containing
657
phospholipids in the remodeling of lipid membranes. Biochim. Biophys. Acta -
658
Biomembr. 1838, 1509–1517.
659
(31) Medeiros, D., Silva-Gonçalves, L. D. C., Da Silva, A. M. B., Dos Santos
660
Cabrera, M. P., and Arcisio-Miranda, M. (2017) Membrane-mediated action of the
661
endocannabinoid anandamide on membrane proteins: implications for understanding
662
the receptor-independent mechanism. Sci. Rep. 7, 41362.
663
(32) Lynch, D. L., and Reggio, P. H. (2005) Molecular dynamics simulations of the
664
endocannabinoid N-arachidonoylethanolamine (anandamide) in a phospholipid
665
bilayer: probing structure and dynamics. J. Med. Chem. 48, 4824–4833.
ACS Paragon Plus Environment
22
Page 23 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
666
(33) Reggio, P. H., and Traore, H. (2000) Conformational requirements for
667
endocannabinoid interaction with the cannabinoid receptors, the anandamide
668
transporter and fatty acid amidohydrolase. Chem. Phys. Lipids 108, 15–35.
669
(34) Núñez, E., Alonso-Torres, P., Fornés, A., Aragón, C., and López-Corcuera, B.
670
(2008) The neuronal glycine transporter GLYT2 associates with membrane rafts:
671
functional modulation by lipid environment. J. Neurochem. 105, 2080–2090.
672
(35) Oh, D. Y., Jung, M. Y., Mi, J. M., Jong-Ik, H., Han, C., Ju, Y. L., Jae, I. K.,
673
Sunoh, K., Hyewhon, R., David, K. O. D., Michael, J. W., Heung, S. N., Min, G. L.,
674
Hyuk, B. K., Kyungjin, K., and Jae, Y. S. (2008) Identification of farnesyl
675
pyrophosphate and N-arachidonylglycine as endogenous ligands for GPR92. J. Biol.
676
Chem. 283, 21054–21064.
677
(36) Ahn, K. H., Bertalovitz, A. C., Mierke, D. F., and Kendall, D. A. (2009) Dual
678
role of the second extracellular loop of the cannabinoid receptor 1: ligand binding and
679
receptor localization. Mol Pharmacol. 76, 833–842.
680
(37) Bai, J.-Y., Ding, W.-G., Kojima, A., Seto, T., and Matsuura, H. (2015) Putative
681
binding sites for arachidonic acid on the human cardiac Kv 1.5 channel. Br. J.
682
Pharmacol. 172, 5281–5292.
683
(38) Malde, A. K., Zuo, L., Breeze, M., Stroet, M., Poger, D., Nair, P. C.,
684
Oostenbrink, C., and Mark, A. E. (2011) An Automated force field Topology Builder
685
(ATB) and repository: version 1.0. J. Chem. Theory Comput. 7, 4026–4037.
686
(39) Koziara, K. B., Stroet, M., Malde, A. K., and Mark, A. E. (2014) Testing and
687
validation of the Automated Topology Builder (ATB) version 2.0: prediction of
688
hydration free enthalpies. J. Comput. Aided. Mol. Des. 28, 221–233.
689
(40) Canzar, S., El-Kebir, M., Pool, R., Elbassioni, K., Malde, A. K., Mark, A. E.,
690
Geerke, D. P., Stougie, L., and Klau, G. W. (2013) Charge group partitioning in
691
biomolecular simulation. J. Comput. Biol. 20, 188–198.
692
(41) Poger, D., and Mark, A. E. (2010) On the validation of molecular dynamics
693
simulations of saturated and cis-monounsaturated phosphatidylcholine lipid bilayers:
694
a comparison with experiment. J. Chem. Theory Comput. 6, 325–336.
695
(42) Szegezdi, J., and Csizmadia, F. (2007) A method for calculating the pKa values
696
of small and large molecules. ACS Spring Meet.
697
(43) Abraham, M. J., Murtola, T., Schulz, R., Páll, S., Smith, J. C., Hess, B., and
698
Lindahl, E. (2015) Gromacs: high performance molecular simulations through multi-
699
level parallelism from laptops to supercomputers. SoftwareX 1–2, 19–25.
ACS Paragon Plus Environment
23
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 24 of 37
700
(44) Schmid, N., Eichenberger, A. P., Choutko, A., Riniker, S., Winger, M., Mark, A.
701
E., and Van Gunsteren, W. F. (2011) Definition and testing of the GROMOS force-
702
field versions 54A7 and 54B7. Eur. Biophys. J. 40, 843–856.
703
(45) Hermans, J., Berendsen, H. J. C., Van Gunsteren, W. F., and Postma, J. P. M.
704
(1984) A consistent empirical potential for water–protein interactions. Biopolymers
705
23, 1513–1518.
706
(46) Carland, J. E., Thomas, M., Mostyn, S. N., Subramanian, N., O’Mara, M. L.,
707
Ryan, R. M., and Vandenberg, R. J. (2018) Molecular determinants for substrate
708
interactions with the glycine transporter GlyT2. ACS Chem. Neurosci. 9, 603–614.
709
(47) Bussi, G., Donadio, D., and Parrinello, M. (2007) Canonical sampling through
710
velocity rescaling. J. Chem. Phys. 126, 014101.
711
(48) Hess, B., Bekker, H., Berendsen, H. J. C., and Fraaije, J. G. E. M. (1997) LINCS:
712
a linear constraint solver for molecular simulations. J Comput Chem 18, 1463–1472.
713
(49) Miyamoto, S., and Kollman, P. A. (1992) Settle: an analytical version of the
714
SHAKE and RATTLE algorithm for rigid water models. J. Comput. Chem. 13, 952–
715
962.
716
(50) Humphrey, W., Dalke, A., and Schulten, K. (1996) VMD: visual molecular
717
dynamics. J. Mol. Graph. 14, 33–38.
718
(51) Buchoux, S. (2017) FATSLiM: a fast and robust software to analyze MD
719
simulations of membranes. Bioinformatics 33, 133–134.
720
(52) Pluhackova, K., Kirsch, S. A., Han, J., Sun, L., Jiang, Z., Unruh, T., and
721
Böckmann, R. A. (2016) A critical comparison of biomembrane force fields: structure
722
and dynamics of model DMPC, POPC, and POPE bilayers. J. Phys. Chem. B 120,
723
3888–3903.
724
(53) Piggot, T. J., Ángel, P., and Khalid, S. (2017) Correction to molecular dynamics
725
simulations of phosphatidylcholine membranes: a comparative force field study. J.
726
Chem. Theory Comput. 13, 1862–1865.
727
(54) Piggot, T. J., Allison, J. R., Sessions, R. B., and Essex, J. W. (2017) On the
728
calculation of acyl chain order parameters from lipid simulations. J. Chem. Theory
729
Comput. 13, 5683–5696.
730
(55) Schumann-Gillett, A., and O’Mara, M. L. (2018) The effects of oxidised
731
phospholipids and cholesterol on the biophysical properties of POPC bilayers.
732
Biochim. Biophys. Acta - Biomembr. DOI: 10.1016/j.bbamem.2018.07.012.
733
(56) Maiorov, V. N., and Crippen, G. M. (1995) Size-independent comparison of
ACS Paragon Plus Environment
24
Page 25 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
734
ACS Chemical Neuroscience
protein three-dimensional structures. Proteins 22, 273–283.
735 736
ACS Paragon Plus Environment
25
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
737
Page 26 of 37
Tables
738 Property
Area per POPC (Å2)
RMSD (Å)
System
No GlyT2
GlyT2
No GlyT2
GlyT2
GlyT2
POPC + 20 % CLR
42.8 ± 0.5
42.8 ± 0.2
53.0 ± 0.7
55.2 ± 0.9
3.1 ± 0.2
43.0 ± 0.3
42.8 ± 0.2
50.5 ± 0.6
53.8 ± 0.8
2.8 ± 0.4
43.4 ± 0.4
42.9 ± 0.2
49.2 ± 0.8
53.8 ± 0.9
3.2 ± 0.3
43.0 ± 0.3
42.8 ± 0.3
51.2 ± 0.6
55.3 ± 0.8
2.9 ± 0.4
43.1 ± 0.4
42.8 ± 0.2
51.0 ± 0.7
54.5 ± 0.9
3.0 ± 0.3
POPC + 20 % CLR + 5 % ANAN POPC + 20 % CLR + 5 % NAGL POPC + 20 % CLR + 5 % CARN Mean
739 740 741 742
P-P distance (Å)
Table 1. Mean properties of the bilayer and mean root-mean-square deviation (RMSD) of GlyT2 during the simulations (100 ns–500 ns). The data are averaged over the three simulation replicates. The error is the standard deviation. CLR, cholesterol; ANAN, anandamide; NAGL, N-arachidonylglycine; CARN, oleoyl-L-carnitine.
743
ACS Paragon Plus Environment
26
Page 27 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
Number of lipids Bilayer
No GlyT2 BIOA
POPC + 20 % CLR POPC + 20 % CLR + 5 % ANAN POPC + 20 % CLR + 5 % ANAN (large bilayer) POPC + 20 % CLR + 5 % NAGL POPC + 20 % CLR + 5 % CARN
744 745
CLR
POPC
-
32
16
GlyT2 TOTAL
BIOA
CLR
POPC
TOTAL
160
192
-
223
851
1074
64
240
320
55
223
851
1129
64
256
960
1280
-
-
-
-
16
64
240
320
57
218
851
1126
16
64
240
320
57
223
871
1151
Table 2. The total number of lipids in each bilayer. BIOA, bioactive lipid; CLR, cholesterol; ANAN, anandamide; NAGL, N-arachidonylglycine; CARN, oleoyl-L-carnitine.
746 747 748 749 750 751 752
Figure Legends
753 754
Figure 1. The chemical structure of anandamide, N-arachidonylglycine and oleoyl-L-
755
carnitine. The carbon-carbon bonds are coloured pink for anandamide, blue for N-
756
arachidonylglycine and maroon for oleoyl-L-carnitine. The carbon numbers that have
757
cis double bonds are labelled, as well as the number of carbon atoms in each acyl
758
chain.
759 760 761
ACS Paragon Plus Environment
27
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 28 of 37
762
Figure 2. The systems studied in this work. The composition of each bilayer is
763
labelled. A–D) The bilayers lacking GlyT2. E–H) the bilayers containing GlyT2.
764
POPC and the bioactive lipids are shown in space-fill representation, with POPC
765
carbon atoms coloured tan and the carbon atoms for anandamide (ANAN), N-
766
arachidonylglycine (NAGL) and oleoyl-L-carnitine (CARN), coloured pink, blue and
767
maroon respectively. Cholesterol is shown as yellow sticks; water is shown as a
768
transparent blue surface; and GlyT2 is represented as cartoon, with its molecular
769
surface shown in transparent grey, its transmembrane helices shown in multiple
770
colours and extracellular loop 4 (EL4) coloured black.
771 772 773 774
Figure 3. Biophysical properties of the bilayers calculated over the last 400 ns of
775
simulations. (A) Interleaflet phosphorus–phosphorus distance for each bilayer, with
776
(solid) and without (transparent) GlyT2. The data are expressed as box plots. (B) Area
777
per POPC lipid for each bilayer, with (solid) and without (transparent) GlyT2. The
778
data are expressed as box plots. (C) and (D) Order parameters for each carbon atom
779
on the POPC acyl chains for the bilayers lacking (C) and containing (D) GlyT2. The
780
plots corresponding to each of the two acyl chains of POPC are labelled. The data are
781
averaged over the simulation time and the three replicates for each system. Note that
782
all data are included in panel B; the data are simply overlapping. The systems were
783
CONT., control bilayer (20 % cholesterol + 80 % POPC); ANAN, anandamide
784
bilayer (5 % anandamide + 20 % cholesterol + 75 % POPC); NAGL, N-
785
arachidonylglycine bilayer (5 % N-arachidonylglycine + 20 % cholesterol + 75 %
786
POPC); and CARN, oleoyl-L-carnitine bilayer (5 % oleoyl-L-carnitine + 20 %
787
cholesterol + 75 % POPC).
788 789 790 791
Figure 4. Locations of the lipids in the bilayers. (A–B) Lipid head group atom
792
number density across the bilayer. The normalised number of atoms at each location
793
across the bilayer is shown, averaged over time (100–500 ns) and the three replicates
794
for each system. Water extends from 0 nm to 5 nm. The atoms chosen to represent the
795
head group of each lipid were: POPC, phosphorus; cholesterol (CLR), oxygen; and
ACS Paragon Plus Environment
28
Page 29 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
796
bioactive lipid (BIOA), nitrogen. The data for each lipid are labelled and coloured
797
according to the type of bioactive lipid present: CONT., control bilayer (20 %
798
cholesterol + 80 % POPC); ANAN, anandamide bilayer (5 % anandamide + 20 %
799
cholesterol + 75 % POPC); NAGL, N-arachidonylglycine bilayer (5 % N-
800
arachidonylglycine + 20 % cholesterol + 75 % POPC); and CARN, oleoyl-L-carnitine
801
bilayer (5 % oleoyl-L-carnitine + 20 % cholesterol + 75 % POPC). A) Bilayer without
802
GlyT2, with a transparent snapshot from an ANAN bilayer simulation overlaid.
803
Anandamide is shown in pink space fill, POPC is shown as tan space fill and
804
cholesterol is shown as yellow sticks. Water is excluded for clarity. B) Bilayer with
805
GlyT2, with the atom density of GlyT2 shaded grey. C) Anandamide “snorkelled”
806
from the lipid/water interface and moved into the centre of the bilayer (two examples
807
circled in black). GlyT2 is shown in cartoon representation, coloured by helix. A
808
slice-through of the surface of GlyT2 is shown in transparent grey. Anandamide is
809
shown as space-fill in CPK colouring, with magenta carbon atoms. The phosphorus
810
atoms of POPC are shown as purple spheres. POPC, cholesterol, water and ions are
811
hidden for clarity.
812 813 814 815
Figure 5. Average intra-molecular distance for the ends of the bioactive lipids. The
816
data are shown as violin, or “jitter” plots, coloured by lipid for the membranes
817
without (A) and with (B) GlyT2. Averages were taken over 100–500 ns. The number
818
of data points corresponds to the number of lipids, e.g. each anandamide molecule, in
819
each system, for the three replicates. The systems are labelled ANAN, anandamide
820
bilayer (5 % anandamide + 20 % cholesterol + 75 % POPC); NAGL, N-
821
arachidonylglycine bilayer (5 % N-arachidonylglycine + 20 % cholesterol + 75 %
822
POPC); and CARN, oleoyl-L-carnitine bilayer (5 % oleoyl-L-carnitine + 20 %
823
cholesterol + 75 % POPC). The united-atom structures, with the distance vectors and
824
example values for the bioactive lipids are shown in (C). The colours are as in (A) and
825
(B), with the first and last carbon in the distance pairing shown as black spheres.
826 827 828
Figure 6. Binding locations and binding propensity of anandamide (A), N-
829
arachidonylglycine (B) and oleoyl-L-carnitine (C) to GlyT2. GlyT2 is shown in
ACS Paragon Plus Environment
29
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 30 of 37
830
cartoon, coloured by transmembrane helix (TM). The binding propensity is shown as
831
a greyscale colour band, with dark bands indicating low binding propensity (greater
832
than 0.001) and white bands indicating 0.1 (i.e. 100 %) binding propensity (i.e. most
833
of the simulation, for all three replicates). Solid rainbow colour on the transmembrane
834
helices and light salmon on the extra- and intracellular regions indicate areas where
835
bioactive lipids never bound, i.e. the binding propensity was 0 %.
836 837 838
Graphical Table of Contents
839
840
ACS Paragon Plus Environment
30
Page 31 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
Graphical Table of Contents 80x42mm (300 x 300 DPI)
ACS Paragon Plus Environment
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Figure 1. The chemical structure of anandamide, N-arachidonylglycine and oleoyl-L-carnitine. The carboncarbon bonds are coloured pink for anandamide, blue for N-arachidonylglycine and maroon for oleoyl-Lcarnitine. The carbon numbers that have cis double bonds are labelled, as well as the number of carbon atoms in each acyl chain. 84x66mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 32 of 37
Page 33 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
Figure 2. The systems studied in this work. The composition of each bilayer is labelled. A–D) The bilayers lacking GlyT2. E–H) the bilayers containing GlyT2. POPC and the bioactive lipids are shown in space-fill representation, with POPC carbon atoms coloured tan and the carbon atoms for anandamide (ANAN), Narachidonylglycine (NAGL) and oleoyl-L-carnitine (CARN), coloured pink, blue and maroon respectively. Cholesterol is shown as yellow sticks; water is shown as a transparent blue surface; and GlyT2 is represented as cartoon, with its molecular surface shown in transparent grey, its transmembrane helices shown in multiple colours and extracellular loop 4 (EL4) coloured black. 84x133mm (300 x 300 DPI)
ACS Paragon Plus Environment
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Figure 3. Biophysical properties of the bilayers calculated over the last 400 ns of simulations. (A) Interleaflet phosphorous–phosphorous distance for each bilayer, with (solid) and without (transparent) GlyT2. The data are expressed as box plots. (B) Area per POPC lipid for each bilayer, with (solid) and without (transparent) GlyT2. The data are expressed as box plots. (C) and (D) Order parameters for each carbon atom on the POPC acyl chains for the bilayers lacking (C) and containing (D) GlyT2. The plots corresponding to each of the two acyl chains of POPC are labelled. The data are averaged over the simulation time and the three replicates for each system. Note that all data are included in panel B; the data are simply overlapping. The systems were CONT., control bilayer (20 % cholesterol + 80 % POPC); ANAN, anandamide bilayer (5 % anandamide + 20 % cholesterol + 75 % POPC); NAGL, N-arachidonylglycine bilayer (5 % Narachidonylglycine + 20 % cholesterol + 75 % POPC); and CARN, oleoyl-L-carnitine bilayer (5 % oleoyl-Lcarnitine + 20 % cholesterol + 75 % POPC).
ACS Paragon Plus Environment
Page 34 of 37
Page 35 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
Figure 4. Locations of the lipids in the bilayers. (A–B) Lipid head group atom number density across the bilayer. The normalised number of atoms at each location across the bilayer is shown, averaged over time (100–500 ns) and the three replicates for each system. Water extends from 0 nm to 5 nm. The atoms chosen to represent the head group of each lipid were: POPC, phosphorus; cholesterol (CLR), oxygen; and bioactive lipid (BIOA), nitrogen. The data for each lipid are labelled and coloured according to the type of bioactive lipid present: CONT., control bilayer (20 % cholesterol + 80 % POPC); ANAN, anandamide bilayer (5 % anandamide + 20 % cholesterol + 75 % POPC); NAGL, N-arachidonylglycine bilayer (5 % Narachidonylglycine + 20 % cholesterol + 75 % POPC); and CARN, oleoyl-L-carnitine bilayer (5 % oleoyl-Lcarnitine + 20 % cholesterol + 75 % POPC). A) Bilayer without GlyT2, with a transparent snapshot from an ANAN bilayer simulation overlaid. Anandamide is shown in pink space fill, POPC is shown as tan space fill and cholesterol is shown as yellow sticks. Water is excluded for clarity. B) Bilayer with GlyT2, with the atom density of GlyT2 shaded grey. C) Anandamide “snorkelled” from the lipid/water interface and moved into the centre of the bilayer (two examples circled in black). GlyT2 is shown in cartoon representation, coloured by helix. A slice-through of the surface of GlyT2 is shown in transparent grey. Anandamide is shown as spacefill in CPK colouring, with magenta carbon atoms. The phosphorus atoms of POPC are shown as purple spheres. POPC, cholesterol, water and ions are hidden for clarity. 2799x2481mm (30 x 30 DPI)
ACS Paragon Plus Environment
ACS Chemical Neuroscience 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Figure 5. Average intra-molecular distance for the ends of the bioactive lipids. The data are shown as violin, or “jitter” plots, coloured by lipid for the membranes without (A) and with (B) GlyT2. Averages were taken over 100–500 ns. The number of data points corresponds to the number of lipids, e.g. each anandamide molecule, in each system, for the three replicates. The systems are labelled ANAN, anandamide bilayer (5 % anandamide + 20 % cholesterol + 75 % POPC); NAGL, N-arachidonylglycine bilayer (5 % Narachidonylglycine + 20 % cholesterol + 75 % POPC); and CARN, oleoyl-L-carnitine bilayer (5 % oleoyl-Lcarnitine + 20 % cholesterol + 75 % POPC). The united-atom structures, with the distance vectors and example values for the bioactive lipids are shown in (C). The colours are as in (A) and (B), with the first and last carbon in the distance pairing shown as black spheres. 84x88mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 36 of 37
Page 37 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Chemical Neuroscience
Figure 6. Binding locations and binding propensity of anandamide (A), N-arachidonylglycine (B) and oleoylL-carnitine (C) to GlyT2. GlyT2 is shown in cartoon, coloured by transmembrane helix (TM). The binding propensity is shown as a greyscale colour band, with dark bands indicating low binding propensity (greater than 0.001) and white bands indicating 0.1 (i.e. 100 %) binding propensity (i.e. most of the simulation, for all three replicates). Solid rainbow colour on the transmembrane helices and light salmon on the extra- and intracellular regions indicate areas where bioactive lipids never bound, i.e. the binding propensity was 0 %.
ACS Paragon Plus Environment