background image

CPT1a regulates the delivery of extracellular fatty acids for cardiolipin turnover in 

prostate cancer cells. 

Nancy T Santiappillai

1

, Mariam F Hakeem-Sanni

1

, Anabel Withy

1

, Lisa M Butler

2,3

, Lake-Ee 

Quek

4

, Andrew J Hoy

1,5 

1

The University of Sydney, Charles Perkins Centre, School of Medical Sciences, Sydney, New South Wales, 2006, 

Australia. 

2

South Australian Immunogenomics Cancer Institute and Freemasons Centre for Male Health and Wellbeing, 

University of Adelaide, Adelaide, SA 5005, Australia 

3

South Australian Health and Medical Research Institute, Adelaide, SA 5000, Australia 

4

The University of Sydney, Charles Perkins Centre, School of Mathematics and Statistics, Sydney, New South 

10 

Wales, 2006, Australia. 

11 

5

Lead contact. Tel: +61 2 9351 2514; Email: andrew.hoy@sydney.edu.au 

 

12 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

Abstract

 

13 

Mitochondrial fatty acid oxidation (FAO) has been proposed to be a major bioenergetic 

14 

pathway in prostate cancer. However, this concept fails to consider FAO relative to other 

15 

mitochondrial substrates. Here, we found extracellular long-chain fatty acids (LCFAs), 

16 

including palmitate, stearate, oleate, linoleate, linolenate, are minor sources of carbon entering 

17 

the TCA cycle compared to glucose and glutamine in prostate cancer cells, despite being 

18 

assimilated in the mitochondria as acyl-carnitines. In contrast, cardiolipins were a prominent 

19 

LCFAs sink, with some species achieving greater than 50% 

13

C-labelling within 6 hours, 

20 

suggesting high cardiolipin turnover using extracellular LCFAs. Knockdown of CPT1a, the 

21 

rate-limiting enzyme of LCFA entry into mitochondria, reduced the incorporation of 

22 

extracellular linoleate into cardiolipins. These results demonstrate that FAO is not a major input 

23 

for the TCA cycle and provide evidence for an underappreciated role for CPT1a in regulating 

24 

LCFAs entry into mitochondria for cardiolipin remodelling. 

 

25 

Keywords 

26 

Fatty acid oxidation, parallel stable isotope tracing, TCA cycle, cardiolipin, prostate cancer, 

27 

metabolomics

 

 

28 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

Introduction 

29 

Prostate cancer (PCa) display unique metabolic features compared to many other solid 

30 

tumours, as it typically does not exhibit the “Warburg effect”. A hallmark of PCa is alterations 

31 

in lipid metabolism, which arise from reports that PCa overexpress lipid metabolism genes

1-3

32 

have enhanced fatty acid (FA) oxidation (FAO) rates

4-6

, increased sensitivity to inhibition of 

33 

FAO and FA synthesis pathways

7,8

, lipid profile alterations that associate with PCa 

34 

progression

8

, and increased citrate oxidation for lipogenesis

9,10

 compared to benign epithelial 

35 

cells and tissues. Of these features, it is the broader interpretations of the higher FAO rates that 

36 

are poorly supported. In that, it has been long proposed that FAO is the predominant 

37 

bioenergetic source in PCa

11,12

, yet there have been no quantifications of FAO rates in relation 

38 

to other major mitochondria substrates, such as glucose and glutamine. PCa cells consume non-

39 

trivial amounts of glucose and glutamine

6,13

 and this shifts along a spectrum of glucose and 

40 

lipid utilisation from benign to castration resistant PCa

14,15

. However, it is unclear whether 

41 

FAO is a major carbon source for the TCA cycle, and thereby support the widely held belief 

42 

that it is the predominant bioenergetic source, compared to glucose and glutamine in PCa cells.  

43 

Cytosolic LCFA-CoAs enter the mitochondria through the carnitine shuttle system

16

, with the 

44 

contemporary view being that mitochondria LCFA-CoAs are shunted into 

β

-oxidation to 

45 

produce acetyl-CoA for the TCA cycle, and NADH and FADH

2

 for the electron transport chain 

46 

(ETC)

17

. As such, CPT1 is viewed as the rate-limiting step of FAO, which is supported by 

47 

evidence that loss-of-function decreases CO

2

 and ATP production and causes cell death in 

48 

cancer cells

5,18-20

. This CPT1a-FAO-ATP-cell death cascade requires FAO to be the 

49 

predominant bioenergetic process, and that glucose and glutamine metabolism are unable to 

50 

sustain ATP levels in the setting of impaired FAO. Significant doubts about this cascade arose 

51 

from Yao 

et al

21

 

who reported that pharmacological inhibition of FAO with low concentrations 

52 

of the CPT1 inhibitor etomoxir did not affect the proliferation rates of various cancer cells. 

53 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

Secondly, they showed that CPT1a loss-of-function in BT549 breast cancer cells decreased 

54 

mitochondrial lipid levels and altered mitochondrial morphology

21

. Similar observations have 

55 

been reported in T cells

22

 and mouse heart

23

. Together, these results raise the possibility that 

56 

intramitochondrial LCFA-CoAs are not a major TCA cycle source of carbons, and that CPT1 

57 

is essential for maintaining intramitochondrial lipid homeostasis and ultimately questions the 

58 

sole commitment of mitochondria LCFA-CoAs into beta-oxidation. In this study, we aimed to 

59 

determine whether FAO provides a significant carbon source to the TCA cycle in PCa cells and 

60 

whether extracellular LCFAs contribute to other significant mitochondrial fates. 

61 

62 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

Results 

63 

Glucose and glutamine are the major TCA cycle carbon sources in prostate cancer cells. 

64 

PCa is characterised by higher rates of FAO compared to prostate epithelial cells

4-6

, however, 

65 

it remains to be determined how these rates relate to the use of other sources. We aimed to 

66 

quantify, in parallel, the conversion of extracellular glucose, glutamine and LCFAs into TCA 

67 

cycle metabolites in a panel of PCa cells spanning the spectrum of disease, from benign PNT1 

68 

cells to late-stage PC-3 cells. Consistent with previous reports

4-6

, PCa cells had 2-4- fold greater 

69 

FAO rates compared to PNT1 benign cells (Fig 1a). These differences in baseline FAO rates 

70 

could not be explained by active mitochondria amount (Fig. 1b, c), nor by CPT1a protein levels 

71 

(Fig. 1d, e). Next, we performed parallel [U-

13

C] isotope tracing experiments to quantify the 

72 

relative incorporation of uniformly labelled palmitate (FA 16:0), glucose, and glutamine 

73 

carbons into TCA cycle metabolites

Palmitate was chosen as the representative LCFA as it is 

74 

the most abundant FA circulating in humans

24

. We determined that 6 hours was the optimal 

75 

endpoint for our tracing experiments as isotopic steady state was reached, and media substrates 

76 

were not depleted (>50% of starting levels; Extended Data Fig. 1a). Similar studies using 

13

C-

77 

glucose and -glutamine in cultured cells have shown that 3 to 6 hours of label exposure is 

78 

sufficient for isotopic steady state

25,26

. We separately noted that doubling of media volume to 

79 

2 mL was necessary for a 12 to 24 hour experiment, in order to maintain media substrates above 

80 

50% (Extended Data Fig. 1b). A notable pattern was the consumption of media lactate by cells 

81 

that had depleted glucose (Extended Data Fig. 1a), thereby, demonstrating metabolic plasticity 

82 

of PCa cells. 

83 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

 

84 

Fig. 1: Palmitate is a minor TCA cycle substrate source compared to glucose and 

85 

glutamine. 

86 

(A)

 

Fatty acid [1-

14

C]-FA 16:0 oxidation rates. N=3-5 per cell line. FAO rates were compared to PNT1 

87 

cells. 

88 

(B)

 

Active mitochondria as a fraction of mitochondrial potential to mitochondrial content normalised 

89 

to PNT1 cells. N=3-4 per cell line. 

90 

(C)

 

Correlation of active mitochondria content to FAO excluding PNT1 cells. Error bars represent 

 

91 

SD. P and r

2

 values by linear regression. 

92 

(D)

 

Quantification of CPT1a protein expression normalised to PNT1 cells. N=3 per cell line. A 

93 

representative immunoblot is shown. 

94 

(E)

 

Correlation of CPT1a protein levels to FAO. Error bars represent 

 SD. P and r

2

 values by linear 

95 

regression. 

96 

(F)

 

Fractional contribution of [U-

13

C]-glucose to pyruvate (red), [U-

13

C]-glutamine to intracellular 

97 

glutamine (blue), [U-

13

C]-FA 16:0 to palmitoyl-carnitine (orange) after 6 hours of labelling. [U-

98 

13

C]-glucose, glutamine, or 16:0 to citrate, oxoglutarate and malate. N=9 per cell line. Level of 

99 

natural enrichment of metabolite standards indicated. 

100 

(G)

 

Fatty acid [1-

14

C]-16:0 oxidation rates of MR42D and MR49F treatment resistant cell lines 

101 

compared to treatment sensitive LNCaP cells. N=3 per cell line. 

102 

(H)

 

Fractional contribution of [U-

13

C]-glucose, glutamine, and FA 16:0 to citrate in MR42D, MR49F 

103 

and LNCaP cells. N=9 per cell line. 

104 

(I)

 

Fractional contribution of [U-

13

C]-FA 16:0 to citrate in LNCaP, C4-2B and 22RV1 3D spheroid 

105 

models. N=3 per cell line. 

106 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

Graphs show mean 

 SEM unless stated otherwise. *p<0.05 by One-way ANOVA with Dunnett’s 

107 

multiple comparisons test, unless stated otherwise.

 

108 

We achieved sufficient labelling of the respective intracellular TCA cycle precursor pools after 

109 

6 hours (pyruvate >65%, glutamine >85%, and palmitoyl-carnitine >60%; Fig. 1f) and 

110 

confirmed the uptake of extracellular labelled substrates. The fractional contribution of [U-

111 

13

C]-glucose and [U-

13

C]-glutamine, calculated as the weighted sum by the number of labelled 

112 

carbons

27

, combined ranged from 60 to 80% for all measured TCA cycle metabolites (Fig. 1f). 

113 

In contrast, extracellular [U-

13

C]-palmitate contributed less than 15% of carbons to TCA cycle 

114 

metabolites (Fig, 1f and Extended Data Fig. 1c-d). 22Rv1 and AD1 cells had the greatest 

115 

enrichment of 

13

C-palmitate, which was consistent with their higher FAO rates (Extended Data 

116 

Fig. 1c). In general, the fractional contribution of palmitate to TCA cycle metabolites weakly 

117 

correlated with FAO and, somewhat surprisingly, inversely correlated with active mitochondria 

118 

content (Extended Data Fig. 1d-e). Overall, non-trivial amounts of carbon from extracellular 

119 

palmitate oxidation entered the TCA cycle, but this influx was dwarfed by the contribution of 

120 

extracellular glucose and glutamine. 

121 

The treatment resistant PCa cells MR42D and MR49F

28

 have been reported to exhibit enhanced 

122 

FA uptake and oxidation

29,30

 and so we next assessed the contribution of extracellular glucose, 

123 

glutamine and palmitate to TCA cycle intermediates in these models. In our hands, FAO rates 

124 

for both MR42D and MR49F cells were not different to treatment sensitive LNCaP cells (Fig. 

125 

1g). Likewise, the fractional contribution of palmitate to citrate in MR42D and MR49F cells 

126 

was similar to LNCaP cells (4-8.5%), and consistent with our other results, citrate was largely 

127 

derived from glucose (35%) and glutamine (20%; Fig. 1h and Extended Data Fig. 1g). We 

128 

repeated our parallel tracing approach using 3D PCa spheroids as they exhibit a more oxidative 

129 

phenotype compared to 2D cell cultures

31-33

. Again, we observed minimal contribution of 

130 

palmitate into TCA cycle metabolites in LNCaP, C4-2B and 22RV1 3D spheroids (<6%; Fig. 

131 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

1i and Extended Data Fig. 1j). Importantly, we confirmed that media glucose, glutamine, and 

132 

palmitate were sufficiently plentiful prior to harvesting cells at 6 hours (Extended Data Fig. 1a, 

133 

h), and that substantial 

13

C labelling of precursor pools were achieved in both treatment 

134 

resistant cells and 3D spheroids (Extended Data Fig. 1f, i). Altogether, we show that 

135 

extracellular glucose and glutamine are the major TCA cycle substrates compared to exogenous 

136 

palmitate in a range of prostate cells, regardless of disease stage pathophysiology, treatment 

137 

responsiveness, or multicellular microenvironment. 

138 

Palmitate oxidation increases in the absence of glucose and glutamine but is insufficient 

139 

to maintain TCA cycle function. 

140 

One concern we had was that the low enrichment values (Fig. 1) were, in part, due to a technical 

141 

or biological issue. As such, we set out to meaningfully increase the enrichment of extracellular 

142 

palmitate into TCA cycle intermediates. Others have shown that limiting glucose availability 

143 

increased FAO rates in MCF7 and T47D breast cancer cells

34

 and so we next determined 

144 

whether FAO in PCa cells is increased by withdrawing glucose from the media, and hence lead 

145 

to greater palmitate incorporation into the TCA cycle. Compared to basal FAO rates measured 

146 

in the presence of 5 mM glucose, FAO in glucose-free conditions was 1- to 5- fold greater 

147 

across the panel of prostate cells (Fig. 2a). We also measured FAO in a range of glucose 

148 

concentrations (0 mM to 5 mM) and observed a dose-dependent suppression of FAO (Extended 

149 

Data Fig. 2a). These data demonstrate that glucose levels, and glycolytic activity, impacts FAO 

150 

rates. 

151 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

 

152 

Fig. 2: Palmitate alone is insufficient at maintaining TCA cycle activity. 

153 

(A)

 

[1-

14

C]-FA 16:0 oxidation rates in the presence (blue) or absence (red) of glucose. N=3-5 per cell line. 

154 

Multiple unpaired t-tests. 

155 

(B)

 

Fractional contribution of [U-

13

C]-FA 16:0 to citrate with or without glucose or glucose and glutamine. N=9 

156 

per cell line.  

157 

(C)

 

Schematic tracing conversion of [U-

13

C]-FA 16:0 into citrate m+2 and m+4 isotopologues.  

158 

(D)

 

Enrichment fraction (%) of [U-

13

C]-FA 16:0 to m+2 and m+4 citrate isotopologues in +glucose/+glutamine, 

159 

- glucose/+glutamine, and -glucose/-glutamine media conditions. N=6 per cell line. 

160 

(E)

 

Fractional contribution of [U-

13

C]-FA 16:0 to glutamate and aspartate +glucose/+glutamine, -

161 

glucose/+glutamine, or -glucose/-glutamine. N=5-6 per cell line.  

162 

(F)

 

Relative abundance (normalised to +glucose/+glutamine) of [U-

13

C]-FA 16:0 to citrate. N=4-9 per cell line.  

163 

Graphs show mean 

 SEM. P<0.05 by One-way ANOVA with Dunnett’s multiple comparisons test, unless stated 

164 

otherwise. *compared to +glucose/+glutamine, ^compared to -glucose/+glutamine. 

165 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

10 

We next quantified the extent that extracellular palmitate provides carbons to TCA cycle 

166 

metabolites in the absence of glucose or glucose and glutamine. There was no significant 

167 

impact on cellular metabolic activity (MTT) or cell viability in glucose-free and glucose-/ 

168 

glutamine-free media conditions at 6 hours (Extended Data Fig. 2b, c). Consistent with the 

169 

increase in FAO rates (Fig. 2a), the fractional contribution of palmitate to citrate was increased 

170 

in glucose-free and glucose-/ glutamine-free conditions, but it remained remarkably minimal 

171 

(~4% to ~6-8%; Fig. 2b). Notably, PNT1 benign and AD1 PCa cells had increased palmitate 

172 

labelling of citrate only in the absence of both glutamine and glucose, but not glucose alone 

173 

(Fig. 2b), suggesting that these cells mobilised glutamine first instead of palmitate to 

174 

compensate for glucose deprivation. In the absence of glucose and glutamine, palmitate 

175 

produced markedly more m+2 and m+4 citrate isotopologues, which represent palmitate 

176 

utilisation in the first and second turns of the TCA cycle, respectively (Fig. 2c, d). This suggests 

177 

that palmitate carbons were retained in the TCA cycle for longer under glucose and glutamine 

178 

deprivation. We also observed other dynamic changes in the usage of palmitate-derived carbons 

179 

during glucose and glutamine deprivation, especially into cataplerosis-derived metabolites 

180 

glutamate and aspartate (Fig. 2e), which are produced from oxoglutarate and oxaloacetate 

181 

respectively (Fig. 2c). Firstly, we observed essentially no incorporation of 

13

C-palmitate into 

182 

glutamate and aspartate in cells cultured in media containing physiological levels of glucose 

183 

and glutamine (Fig. 2e). In glucose-free, and most strikingly in glucose-/glutamine-free 

184 

conditions, we saw significant 

13

C-palmitate labelling into glutamate and aspartate (Fig. 2c, e). 

185 

As aspartate synthesis is essential for cell proliferation

35

 and glutamate for glutathione 

186 

synthesis and redox homeostasis

36

, our results suggest that prostate cells, in the absence of 

187 

glucose or both glucose and glutamine, upregulate cataplerotic reactions in an attempt to 

188 

maintain cellular function and viability. Together, these data demonstrate the dynamic 

189 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

11 

adaptability of prostate cell palmitate metabolism to supply carbons to the TCA cycle when 

190 

deprived of glucose only or glucose and glutamine. 

191 

FAO has been proposed to be the dominant bioenergetic source for PCa cells

11,12

. As such, we 

192 

reasoned that FAO would be capable of sustaining TCA cycle activity in glucose or glucose-

193 

glutamine deprived conditions. In striking contrast to 

13

C-palmitate enrichment patterns, citrate, 

194 

malate and oxoglutarate abundance were reduced in cells cultured in glucose-free and glucose-

195 

/ glutamine-free settings for 6 hours compared to glucose and glutamine replete settings (Fig. 

196 

2f and Extended Data Fig. 2d). Specifically, citrate levels were decreased by 43-70% in most 

197 

cells, except for C4-2B and 22Rv1 cells, which were not impacted by glucose- and 

198 

glucose/glutamine-free conditions (Fig. 2f). We did observe striking reductions in malate and 

199 

oxoglutarate levels (between 53-95%) in all cells in response to glucose and glutamine 

200 

depletion (Extended Data Fig. 2d). Importantly, these results preceded the cell death which 

201 

occurred between 6 and 24 hours of culturing (Extended Data Fig. 2b, c). Combined, these data 

202 

show that even with the increased FAO rates to incorporate more palmitate carbon into the TCA 

203 

cycle as well as into aspartate and glutamate, these changes were insufficient to maintain 

204 

metabolite abundance and so TCA cycle activity that were followed by cell death. 

205 

Extracellular LCFAs are minor TCA cycle substrates.

 

206 

Since palmitate is a minor carbon source for the TCA cycle (Fig. 1 and 2), we questioned 

207 

whether this was a palmitate-specific phenotype.

 

To address this, we repeated our experiments 

208 

with a panel of non-essential (16:0, 18:0, 18:1) and essential (18:2, 18:3) LCFAs of varying 

209 

saturation and carbon chain lengths. There was a similar trend of minimal contribution (2-18%) 

210 

to TCA cycle metabolites observed in our panel of prostate cells, despite the diversity of LCFAs 

211 

(Fig. 3a). 22Rv1 and AD1 cells had the greatest palmitate labelling and FAO rates amongst 

212 

prostate cells (Fig. 1a and Extended Data Fig. 1c), and consistently exhibited greater utilisation 

213 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

12 

of LCFAs by the TCA cycle compared to an unlabelled control (Fig. 3a). In general, there was 

214 

no clear substrate preference between palmitate and other LCFAs to feed carbons to the TCA 

215 

cycle (Fig. 1c), despite substantial labelling of respective acyl-carnitine species (0.8-1, ratio 

216 

m

n

/m0+m

n

; Fig. 3b). Collectively, we show that in the presence of physiological glucose and 

217 

glutamine, diverse extracellular LCFAs were not major carbon sources for the TCA cycle, 

218 

despite substantial acyl-carnitine precursor labelling. 

219 

 

220 

Fig. 3: LCFAs of varying saturations are minor contributing substrate sources to the TCA 

221 

cycle.

 

222 

(A)

 

Fractional contribution of [U-

13

C]-FA 16:0, 18:0, 18:1, 18:2, 18:3 to citrate, oxoglutarate, malate. N=9 per 

223 

cell line.  

224 

(B)

 

Fractional contribution of acyl-carnitines associated with the [U-

13

C]-FA substrates. N=9 per cell line. 

225 

Graphs show mean 

 SEM unless stated otherwise. P<0.05 by Two-way ANOVA with Tukey’s multiple 

226 

comparisons test, *compared to unlabelled control (dotted line), #compared to [U-

13

C]-FA 16:0. 

227 

One possible reason for the disconnect between the substantial conversion of LCFA to acyl-

228 

carnitines and the lack of conversion to TCA cycle intermediates could be incomplete FAO, 

229 

which generates medium chain acyl-carnitine species instead of fully oxidising LCFAs to 

230 

acetyl-CoA

37,38

. In agreement with this hypothesis, FA 16:0, 18:0, and 18:1 greatly enriched 

231 

shortened acyl-carnitine species (Extended Data Fig. 3). Specifically, FA 16:0 significantly 

232 

enriched shortened CAR 14:0 (0.04-0.7 fraction) to CAR 10:0 (0-0.45 fraction), FA 18:0 

233 

enriched CAR 16:0 (0.01-0.38 fraction) to CAR 12:0 (0.04-0.4 fraction), and FA 18:1 enriched 

234 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

13 

CAR 16:1 (0.03-0.79 fraction) to CAR 12:1 (0-0.1 fraction). The greatest amounts of 

235 

exogenous FA enrichment into shortened acyl-carnitines were observed in PNT1, 22Rv1, AD1 

236 

and PC-3 cells (Extended Data Fig. 3), which were cells that also had the greatest incorporation 

237 

of LCFAs into precursor acyl-carnitines (Fig. 3b). These data show that labelled extracellular 

238 

LCFAs are not completely converted to acetyl-CoA, and thereby provides new insights into the 

239 

relationships between acyl-carnitine synthesis, LCFA entry into mitochondria, and FAO. 

240 

Extracellular LCFAs supports the rapid turnover of cardiolipin pools. 

241 

Our observations showed that extracellular LCFAs were not completely oxidised to acetyl-CoA 

242 

and minimally contributed carbons to the TCA cycle, despite being substantially present as 

243 

acyl-carnitines (Fig. 3 and Extended Data Fig. 3). As such, we asked whether there are other 

244 

pathways that could consume extracellular LCFAs after entering the mitochondria via CPT1a. 

245 

This question arises from the work of Yao and colleagues

21

 as they were the first to report a 

246 

reduction of mitochondrial complex lipids in BT549 breast cancer cells with CPT1a knocked 

247 

down. Of their observed lipidome results, cardiolipins (CLs) piqued our interest as they are 

248 

exclusively located within the inner mitochondrial membrane (IMM) and account for 10-20% 

249 

of mitochondria phospholipid content

39

. CLs are composed of four fatty acyl chains of different 

250 

saturation and chain lengths

40

 and provide essential structural functions to mitochondria 

251 

membrane and cristae

41

, and ETC complexes I-IV for ATP production

42-44

. As CL remodelling 

252 

functions to mature nascent CLs by the reacylation from donated acyl-CoAs

45

, it was 

253 

conceivable that this process could be a significant consumer of mitochondria LCFA-CoAs. 

254 

Therefore, we next investigated the contribution of extracellular LCFAs to CL synthesis and 

255 

remodelling (Fig. 4a). First, we examined the enrichments of CL pools from extracellular [U-

256 

13

C]-16:0, 18:0, 18:1, and 18:2 LCFAs. Surprisingly, cells cultured in media containing labelled 

257 

LCFAs for 6 hours had >50% enrichment of many CL species, indicating substantial LCFA 

258 

turnover of CL pools (Fig. 4b), in stark contrast to the marginal arrival of LCFA label into the 

259 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

14 

TCA cycle (Fig. 1-3). CL has been shown to be a very slow turnover lipid class with CL half-

260 

life reported to be approximately two days in isolated rat cardiomyocytes

46

 and cultured rat 

261 

H9c2 cardiomyoblast cells

47

. These turnover rates are much slower than other 

262 

glycerophospholipids, including PC

46-49

. Here, we saw time-dependent enrichment of [U-

13

C]-

263 

FA 16:0 into PC 32:1 and 34:1 of C4-2B and PC-3 cells (Extended Data Fig. 4a) and that this 

264 

enrichment was greater than related FA 16:0 containing CL species (CL 68:2 and CL 70:2) at 

265 

6 hours (Extended Data Fig. 4b). As such, C4-2B and PC-3 cells incorporate extracellular FAs 

266 

into PCs at a higher proportion relative to CLs. 

267 

FAs, following activation into FA-CoAs, are substrates for elongation and desaturation 

268 

reactions

50

. In our experiments, we saw that exogenous FAs 16:0, 18:0, and 18:2 were directly 

269 

incorporated into the acyl-chains of CLs (i.e. unmodified via elongation or desaturation; 

270 

Extended Data Fig. 4c). For instance, CL 68:2 is composed of 16:0 and 18:1 acyl-chains and 

271 

was predominantly labelled by FA 16:0, while CL 70:3 is also composed of 16:0 and 18:1 acyl-

272 

chains but was mainly labelled by FA 18:1 (Fig. 4c). Similarly, CL 72:5 is composed of 18:1 

273 

and 18:2 chains and was almost entirely labelled by FA 18:2 within 6 hours (Fig. 4c). We 

274 

observed some differences between PCa cells and benign PNT1 cells; however, no obvious 

275 

pattern could be discerned (Fig. 4c). Finally, to compare the relative incorporation of the four 

276 

FAs into CLs, we calculated an overall percentage of labelled fatty acyl chains among the 15 

277 

quantified CL species. In general, a greater proportion of FA 18:2 was labelled compared to 

278 

FAs 16:0, 18:0, and 18:1 in our panel of prostate cells at 6 hours exposure to the respective 

279 

substrates (Extended Data Fig. 4d). This suggests 18:2-containing CLs have the greatest 

280 

turnover, which was also reported for cardiomyocytes and HeLa cells

51,52

. This may associate 

281 

with the fact a majority of CL acyl chains is 18:2. The exceptions were LNCaP and C4-2B 

282 

cells, which showed greater FA 16:0 13C enrichment among the CLs (32% and 14.2%, 

283 

respectively) compared to FA 18:2 (9.5% and 8.2%, respectively; Extended Data Fig. 4d). 

284 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

15 

Collectively, these data demonstrate that PNT1 prostate epithelial and a range of PCa cells 

285 

rapidly incorporate extracellular LCFAs into CL acyl-chains. 

 

286 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

16 

 

287 

Fig. 4: Extracellular LCFAs are directly and rapidly incorporated into cardiolipin pools. 

 

288 

(A)

 

Schematic depicting fates of FAs entering the mitochondria including FAO and CL pathways. 

289 

(B)

 

Fractional enrichment of [U-

13

C]-FA 16:0, 18:0, 18:1, 18:2 (150 

M) tracing to CL species for 6 hours. N=9 

290 

per cell line. 

291 

(C)

 

Fractional enrichment of [U-

13

C]-FA 16:0, 18:0, 18:1, 18:2 to CL 68:2, CL 70:3, CL 72:5 species. N=9 per 

292 

cell line.  

293 

Graphs show mean 

 SEM. P<0.05 by One-way ANOVA with Tukey’s multiple comparisons test, *PCa cells vs 

294 

PNT1 cells for each FA. 

295 

Reduced CPT1a protein levels decreased FAO and incorporation of LCFAs into CL. 

296 

The high incorporation of extracellular LCFAs into CLs in both PCa cells and PNT1 cells was 

297 

in stark contrast to the magnitude of enrichment into TCA cycle metabolites (Fig. 3 and 4). 

298 

LCFAs are incorporated into CLs via multiple starting points, not exclusively competing with 

299 

FAO for mitochondria FAs that are supplied via CPT1a. While these intramitochondrial acyl-

300 

CoAs are substrates for CL remodelling acyltransferases MLCAT1

45

 and ALCAT1

53

, CLs are 

301 

also assembled by the 

de novo

 pathway producing nascent CLs from phosphatidylglycerol 

302 

(PG)

54

, and then remodelled via acyl-chain exchange with PC and phosphatidylethanolamine 

303 

by the actions of Tafazzin

55

. To tease out the association between CPT1a and CL remodelling, 

304 

we quantified the incorporation of extracellular FA 18:2 in cells with CPT1a knocked down, 

305 

since 18:2 has been reported to be predominantly incorporated into CLs via remodelling 

306 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

17 

pathways

45,51

 (Fig. 5a). We included FA 16:0 in our experiments as it can be incorporated via 

307 

both 

de novo

 synthesis and remodelling

56,45

. Further, we used C4-2B cells for this experiment 

308 

as they had the greatest level of CPT1a protein in our panel of prostate cells (Fig. 1d). As 

309 

expected, CPT1a knockdown (CPT1a

KD

; reduced by 89%) in C4-2B cells (Fig. 5b) decreased 

310 

palmitate oxidation (Fig. 5c) and the abundance of 

13

C-labelled 18:2 and 16:0 acyl-carnitines 

311 

compared to scrambled non-targeting control cells (Fig. 5d). CPT1a

KD

 decreased FA 16:0 

312 

enrichment of citrate from 3.5% to 3% (delta 0.5%) but did not impact the contribution from 

313 

FA 18:2 (Fig. 5e). 

314 

In contrast, we found CPT1a

KD

 decreased FA 18:2 enrichment of CLs. The overall proportion 

315 

of labelled FA 18:2 among the CL species was decreased from 19% to 15%, but this reduction 

316 

was not reciprocated by FA 16:0 (Fig. 5f, Extended Data Fig. 5a). Namely, we saw all 

317 

individual CL species had less FA 18:2 enrichment, but the effects were mixed for FA 16:0 

318 

enrichment (Fig. 5g). The latter suggests FA 16:0 is incorporated into CLs largely via the 

de 

319 

novo

 pathway, which is independent of CPT1a and mitochondrial acyl-CoAs

54

. As expected, 

320 

there was no difference in 

13

C-FA 18:2 and 

3

C-FA 16:0 PG enrichment between CPT1a

KD 

and 

321 

scrambled control cells (Fig. 5h, Extended Data Fig. 5b, c), which suggests that the reduction 

322 

in CPT1 activity did not divert LCFAs towards the endoplasmic reticulum to increase PG and 

323 

CL synthesis. Altogether, we showed that that CPT1a regulates the incorporation of FA 18:2 

324 

into CL, thereby demonstrating that CL remodelling is a pathway downstream of CPT1a 

325 

alongside FAO and the TCA cycle (Fig. 5i). 

326 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

18 

 

327 

Fig. 5: CPT1a knockdown reduced FA 18:2 derived CL remodelling. 

328 

(A)

 

Schematic depicting expected FA 18:2 to CL remodelling downstream of CPT1a and FA 16:0 to CL synthesis 

329 

pathways derived from the endoplasmic reticulum. 

330 

(B)

 

CPT1a knockdown by reverse siRNA transfection in C4-2B cells for 72 hours quantified by immunoblot. 

331 

Representative immunoblots shown. N=3 per group.  

332 

(C)

 

[1-

14

C]-FA 16:0 oxidation rates in non-targeting scrambled (scr) and siCPT1a C4-2B cells. N=3 per group. 

333 

(D)

 

Intensity of [U-

13

C]-FA 18:2 or 16:0 to carnitine species following siCPT1a treatment. N=9 per group. 

334 

Statistically significant differences between labelled (m16 orange or m18 purple) intensities (*) or unlabelled 

335 

(m0 grey) intensities (#). 

336 

(E)

 

Fractional contribution of unlabelled or [U-

13

C]-FA 18:2 or 16:0 to citrate in scr or siCPT1a C4-2B cells. N=9 

337 

per group.  

338 

(F)

 

Percentage of 18:2 or 16:0 acyl chain contained in CLs labelled by exogenous [U-

13

C]-FA. N=9 per group.  

339 

(G)

 

Fractional contribution of [U-

13

C]-FA 18:2 or 16:0 to PG and CL species. Annotated n values on the x-axis 

340 

indicate the number of 18:2 or 16:0 containing acyl-chains. N=9 per group. Unpaired t-tests. 

341 

(H)

 

Delta differences between fractional contribution of [U-

13

C]-FA 18:2 or 16:0 to PG (N=6) and CL species 

342 

(N=17). Statistically significant differences within [U-

13

C]-FA 18:2 or 16:0 groups to 0 difference determined 

343 

by Nonparametric Wilcoxon signed rank test. Error bars represent 

 min to max values. 

344 

(I)

 

Schematic depicting major carbon contributions from glucose and glutamine into the TCA cycle in 

345 

comparison to the substantial contributions of LCFA-derived acyl-CoAs for CLs either by CL remodelling 

346 

downstream of CPT1a or CL synthesis. 

347 

Graphs show mean 

 SEM unless stated otherwise. ns not significant, *p<0.05 between siCPT1a and scr control 

348 

groups by student’s t-test unless stated otherwise.

 

 

349 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

19 

Discussion 

350 

PCa do not exhibit classical Warburg metabolism but displays a more lipid dependent 

351 

phenotype

8,57

. One aspect of this lipid phenotype that is widely accepted is the assumption that 

352 

FAO is the predominant bioenergetic pathway in PCa, which likely arises from reports of 

353 

enhanced FAO rates in PCa cells compared to benign cells

4-6

, low glycolytic activity at early 

354 

disease stages

14,15

, and increased sensitivity to FAO targeting

7

. However, this belief that FAO 

355 

is a dominant bioenergetic pathway in PCa, first proposed by Liu

11

 in 2006, failed to consider 

356 

the contribution of other major respiration substrates. CPT1a has been implicated as the rate-

357 

limiting step of FAO because its inhibition results in decreased ATP production and cancer cell 

358 

death

5,18-20

, and assumes that LCFAs are entirely diverted to FAO and the TCA cycle. Here, we 

359 

generated evidence disputing this dogmatic view by showing extracellular LCFAs are minor 

360 

carbon sources to the TCA cycle compared to glucose and glutamine in PCa cells, despite their 

361 

increased FAO rates compared to benign PNT1 cells. We also showed that LCFAs are not 

362 

completely oxidised to acetyl-CoA and produce shorter acyl-carnitine species. Most strikingly, 

363 

there was substantial incorporation of LCFAs into CLs within 6 hours, and that CPT1a 

364 

regulated the assimilation of FA 18:2 into CLs. Together, our data show a substantial role for 

365 

CPT1a to regulate CL turnover in PCa cells and propose that observations of a CPT1a-

366 

dependent reduction of FAO and mitochondrial function can be explained, in part, by perturbed 

367 

CL maturation and reduced oxidative phosphorylation efficiency. 

368 

We and others have shown that FAO in PCa cells is faster than benign prostate epithelial cells

4-

369 

6

, and that inhibiting CPT1 impacts cell viability

18-20

. These observations, combined with 

370 

similar studies and the hypothesis article by Liu

11

, has resulted in the widely held view that 

371 

FAO is a dominant bioenergetic pathway in PCa. Here, we present comprehensive evidence 

372 

that challenges this. In a series of experiments, we show that extracellular LCFAs are a minor 

373 

contributor of carbons to TCA cycle metabolites, compared to glucose and glutamine, which 

374 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

20 

may, in part, be explained by incomplete oxidation. Further, FAO could not meet the 

375 

anaplerotic demand to maintain TCA cycle activity in the absence of glucose alone or glucose 

376 

and glutamine, leading to cell death. We also determined that extracellular FAs essentially do 

377 

not contribute to glutamate and aspartate synthesis in the presence of extracellular glucose and 

378 

glutamine, but do so in their absence. These results were seen across the spectrum of PCa 

379 

progression, from benign to late-stage, 2D and 3D models. Importantly, the magnitude of 

380 

exogenous FA enrichment into citrate was comparable to similar studies, including ~15% in 

ex 

381 

vivo

 cultured malignant Patient-Derived Xenografts (500 

μ

M [U-

13

C]palmitate, 4 hours)

6

 and 

382 

BT549 cells (100 

μ

M [U-

13

C]palmitate, 24 hours)

21

, <15% in HCC1954 and SKBR3 breast 

383 

cancer cells (100

 μ

M [U-

13

C]palmitate

 

, 24

 hours)

58

, and ~10% in MCF-7 and MDA-MB-231 

384 

breast cancer cells cultured in glucose-free media (100 

μ

M [U-

13

C]palmitate, 4 hours)

59

385 

Therefore, these observations raise questions regarding the many studies demonstrating that 

386 

CPT1 loss-of-function leads to cell death 

19,60

. One possible explanation is that many studies 

387 

used etomoxir, an irreversible inhibitor of CPT1, at concentrations that have been reported to 

388 

have significant off-target effects

21

, and often failed to identify the minimum concentration of 

389 

etomoxir to maximally suppress FAO. Another reason could be due to the focus on 

390 

demonstrating a dose-dependent reduction in cell viability and thereby determining IC

50

 in the 

391 

absence of measures of FAO or CPT1 activity. Finally, the other technical consideration when 

392 

comparing our results to others is the use of the Seahorse XF Analyser Palmitate Oxidation 

393 

Kit, which measures oxygen consumption in saturating conditions of palmitate and carnitine 

394 

but in media that is free of glucose, pyruvate, and glutamine. Further, the recommendation to 

395 

culture cells in substrate limited conditions overnight will impact cellular metabolism. We 

396 

observed striking increases in FAO rates, ranging from a doubling to a 5-fold increase, as well 

397 

as substantial changes in intracellular partitioning of LCFA carbons in glucose-free conditions. 

398 

As such, any observed changes in FAO performed in conditions free of major TCA cycle 

399 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

21 

substrates must be interpreted with caution and likely explains the differences between our 

400 

observations and other published studies. Unlike other studies, we showed that CPT1a

KD

 

401 

substantially decreased acyl-carnitine levels and the synthesis from extracellular LCFAs, and 

402 

had a minor impact the contribution of LCFAs to TCA cycle metabolites in media containing 

403 

5 mM glucose and 1 mM glutamine. Overall, we conclude that extracellular LCFAs are not 

404 

completely oxidised, and are a minor source of carbons for TCA cycle metabolites in a panel 

405 

of PCa cells and benign PNT1 cells, therefore that FAO is not a major bioenergetic pathway. 

406 

In general, much of the literature discussing CPT1 function and mitochondrial FA metabolism 

407 

state that CPT1 is rate-limiting for FAO and that FAO is the sole destination for 

408 

intramitochondrial LCFA-CoAs following facilitated transport by the CPT1 –

 

carnitine-

409 

acylcarnitine translocase – CPT2 system. This perception fails to consider that there are other 

410 

destinations, especially given the evidence that CPT1 inhibition reduces mitochondrial lipid 

411 

levels in cancer cells

21,61

 and alters mitochondrial morphology in BT549 cells

21

, T cells

22

, and 

412 

rat heart muscle

23

. In our studies, we focused on the mitochondria exclusive CL as it is essential 

413 

for mitochondria function and bioenergetics

62

. To our surprise, we saw rapid incorporation of 

414 

extracellular LCFAs into CLs in PCa cells (>50% after 6 hours), which was much faster than 

415 

what has been reported in isolate rat cardiomyocytes (2 days)

46

 and cultured rat H9c2 

416 

cardiomyoblast cells (~54 hours)

47

. That said, we did determine that FA 18:2 was 

417 

predominantly incorporated into the acyl-chains of CLs compared to other LCFAs, which is 

418 

similar to previous reports

51,52

 and that the rate of incorporation of LCFAs into other lipid 

419 

classes was faster than CL

46

. These results were especially important as cancer cells that 

420 

substantially incorporate exogenous 18:2 into CL acyl-chains have enhanced stimulation of 

421 

ETC complex I activity

51

 and reduced susceptibility to apoptosis

63

. These results showing 

422 

significant enrichment of CLs with extracellular LCFAs provided a platform to determine 

423 

whether CPT1 regulates LCFA incorporation into CLs, especially given that CPT1 loss-of-

424 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

22 

function reduces CL and mitochondrial lipids

21,61

. Consistent with our hypothesis, depletion of 

425 

CPT1a reduced 

13

C-labelled FA 18:2 incorporation into CL in C4-2B PCa cells, and so suggest 

426 

that CPT1 influences CL remodelling which may influence mitochondrial respiration and 

427 

apoptosis

51,63

. Taken together, we propose that CLs, alongside FAO, are a prominent fate for 

428 

extracellular LCFAs entering the mitochondria via CPT1a. 

429 

A limitation of our study is that we only quantified the contribution of extracellular substrates 

430 

to mitochondrial pathways, which accounted for less than 65% of carbons to citrate. We 

431 

focused on the relationship between FAO and TCA cycle metabolites, in part, to complement 

432 

our previous reports that extracellular LCFAs as the major precursors for PCa lipids

5

. It is likely 

433 

that the unaccounted for sources of carbons for TCA cycle metabolites include lactate

64

 or 

434 

branched chain amino acids

65

, as well as intracellular sources of FAs and amino acids. 

435 

Quantifying the totality could make new insights into carbon source for the TCA cycle and 

436 

how differs in other cancer types, disease stage and impacted by the tumour microenvironment. 

437 

We also provide evidence for incomplete FAO in PCa cells; however, further investigation into 

438 

the roles of shortened acyl-carnitines may provide insights into other fates of LCFAs, alongside 

439 

our new insights into CLs and TCA cycle biology. Finally, we showed for the first time that 

440 

CPT1a impacts CL remodelling in PCa cells. Future experiments are required to link CPT1a 

441 

function, CL homeostasis and ETC complex structure

62

 and other aspects of mitochondrial 

442 

function, beyond FAO. This is especially true given that altered CL function has similar effects 

443 

on mitochondrial morphology, function, and cell viability

66

 as CPT1a loss-of-function 

444 

experiments

5,18-22

445 

This study provides evidence that extracellular LCFAs are rapidly incorporated into CLs but 

446 

are not the predominant carbon source to the TCA cycle in prostate cells compared to glucose 

447 

and glutamine. Further, we conclude that CPT1a influences CL remodelling alongside FAO 

448 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

23 

and so intramitochondrial LCFA-CoAs are not exclusively used for oxidation. Therefore, this 

449 

work puts forward a new view of the role of CPT1 and its relationship to energy homeostasis 

450 

and cell viability in PCa.

 

 

451 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

24 

Methods 

452 

Key Resources Table 

453 

REAGENT or RESOURCE 

SOURCE 

IDENTIFIER 

Antibodies 

Mouse monoclonal anti-CPT1A 

Abcam 

Cat#Ab128568; RRID: 
AB_11141632 

Rabbit monoclonal anti-GAPDH 

Cell Signalling 
Technology 

Cat#2118; RRID: 
AB_561053 

Chemicals, peptides, and recombinant proteins 

Enzalutamide MDV3100 

Selleckchem 

Cat#S1250 

Lipofectamine RNAiMAX transfection reagent 

Thermo Fisher Scientific 

Cat#13778075 

[1-

14

C] Palmitic acid 

Perkin Elmer 

Cat#NEC075H250UC 

Palmitic acid 

Sigma-Aldrich 

Cat#P0500 

D-Glucose Sigma-Aldrich 

Cat#273053 

L-Glutamine Gibco 

Cat#25030081 

Fatty acid free BSA 

Bovogen 

Cat#BSAS0.10 

L-carnitine hydrochloride 

Sigma-Aldrich 

Cat#C0283 

Mitotracker Green 

Thermo Fisher Scientific 

Cat#M7514 

Mitotracker Red 

Thermo Fisher Scientific 

Cat#M7512 

FCCP Sigma-Aldrich 

Cat#SML2959 

DTT Sigma-Aldrich 

Cat#D0632 

Protease and Phosphatase Inhibitor Cocktail, EDTA free 

Astral Scientific 

Cat#T-2496 

Thiazolyl blue tetrazolium bromide MTT reagent 

Sigma-Aldrich 

Cat#M2128 

Linoleic acid 

Sigma Aldrich 

Cat#L1012 

D-Glucose (U-

13

C

6

, 99%) 

Cambridge Isotope 
Laboratories, Inc. 

Cat#CLM-1396 

L-Glutamine (U-

13

C

5

, 99%) 

Cambridge Isotope 
Laboratories, Inc. 

Cat#CLM-1822-H 

Palmitic acid (U-

13

C

16

, 98%) 

Cambridge Isotope 
Laboratories, Inc. 

Cat#CLM-409 

Stearic acid (U-

13

C

18

, 98%) 

Cambridge Isotope 
Laboratories, Inc. 

Cat#CLM-6990 

Linoleic acid (U-

13

C

18

, 98%) 

Cambridge Isotope 
Laboratories, Inc. 

Cat#CLM-10487 

Linolenic acid (U-

13

C

18

, 98%) 

Cambridge Isotope 
Laboratories, Inc. 

Cat#8386 

Critical commercial assays 

Pierce BCA Protein assay kit 

Thermo Fisher Scientific 

Cat#23225 

Immobilon Crescendo Western HRP substrate Merck 

Millipore Cat#WBLUR0500 

CellCarrier Spheroid ULA 96-well microplates 

Perkin Elmer 

Cat#6055330 

Experimental models: Cell lines 

PNT1 ECACC 

95012614 

LNCaP ATCC 

CRL-1740 

C4-2B ATCC 

CRL-3315 

22Rv1 ATCC 

CRL-2505 

CWR-R1-AD1 

A/Prof. Luke Selth, 
Flinders University 

Nyquist 2013

67

, Li 2013

68

 

PC-3 ATCC 

CRL-1435 

MR42D 

Prof. Lisa Butler, 
SAHMRI 

Bishop 2017

28

 

MR49F 

Prof. Lisa Butler, 
SAHMRI 

Bishop 2017

28

 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

25 

Oligonucleotides 

ON-TARGETplus Human CPT1A siRNA SMARTpool 

Horizon Discovery 

Cat#L-009749-00-0005 

ON-TARGETplus Non-targeting Control pool 

Horizon Discovery 

Cat#

 

D-001810-10-05 

Software and algorithms 

GraphPad Prism V9 

GraphPad Software 

https://www.graphpad.co
m/

 

Image Lab 6.1 

Bio-Rad 

https://www.bio-
rad.com/en-
au/product/image-lab-
software

 

FlowJo 2 

FlowJo LLC 

https://www.flowjo.com/

 

MATLAB Mathworks 

https://au.mathworks.com
/products/matlab.html

 

MSConvert N/A 

Chambers 

2012

69

 

MSDial Riken 

http://prime.psc.riken.jp/c
ompms/msdial/main.html

 

Biorender Biorender 

https://biorender.com/

 

 

454 

Method Details 

455 

Cell lines and culture conditions 

456 

Human benign prostate epithelial cell line PNT1 and human prostate carcinoma cells lines 

457 

LNCaP (androgen receptor (AR) +, hormone sensitive), C4-2B, 22Rv1, AD1 (AR+, androgen 

458 

independent), and PC-3 (AR-, hormone resistant) cells were cultured in RPMI1640 medium 

459 

(Life Technologies), supplemented with 10% FBS (Cytiva Hyclone), and 1% 

460 

penicillin/streptomycin (Gibco). MR42D and MR49F (hormone resistant, treatment resistant) 

461 

cells were cultured in RPMI 1640 medium (Life Technologies), supplemented with 10% FBS, 

462 

1% penicillin/streptomycin, and 10

M enzalutamide. All cells were incubated at 37

C and 5% 

463 

CO

2

.  

464 

Spheroid culturing 

465 

Cells were seeded in 96-well Spheroid microplates, at a density of 5 000 cells/well (LNCaP), 

466 

8 000 cells/well (22Rv1), and 2 000 cells/well (C4-2B), in RPMI medium in 100 

L per well. 

467 

An additional 100 

L of medium was added to cells at 24h and left un-agitated for a further 

468 

24h until spheroids formed.  

469 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

26 

Cell Transfection 

470 

C4-2B cells were transfected for 72 hours using RNAiMAX transfection reagent and 25 pmol 

471 

pooled CPT1a siRNA or non-targeting scrambled control, according to the manufacturer’s 

472 

instructions. After 72 hours, cells were used for subsequent radiolabelling, western 

473 

immunoblotting, and 

13

C-labelling experiments. 

474 

14

C-labelling 

475 

Cells were washed in warm PBS and incubated in 600 

L media containing 0.1 mmol/L cold 

476 

palmitate, [1-

14

C]-palmitate (0.1

Ci/mL) conjugated to 2% (wt/vol) FA-free BSA and 1 

477 

mmol/L -carnitine in RPMI glucose-free medium (Gibco) for 4 hours. Fatty acid oxidation 

478 

rates were determined from 

14

CO

2

 production as previously described

1

 and supplemented with 

479 

0 mM or 5 mM glucose to measure metabolic flexibility. Medium for dose dependent glucose 

480 

deprivation results were supplemented with 0 mM, 1 mM, 3 mM, 5 mM glucose. Protein 

481 

content was measured by BCA protein assay and absorbances read at 540 nm using a TECAN 

482 

infinite M1000 pro plate reader. 

483 

Mitotracker

 

484 

Cells were seeded in 12-well plates at a density of 2x10

5

 cells/well in RPMI medium. After 24 

485 

hours, cells were washed in warm PBS and obtained by trypsinisation and collected in serum-

486 

free RPMI medium. The cell suspension was centrifuged at 1500 rpm for 5 min in fluorescence-

487 

activated cell sorting (FACS) tubes. Cells were subsequently stained with 200 nM Mitotracker 

488 

Red and Mitotracker Green at 37

C for 30mins. Additionally, unstained, red only, green only, 

489 

and 2.5 

M FCCP treated cells were included as experimental controls. Cells were centrifuged 

490 

at 1500 rpm for 5min and fixed with 4% paraformaldehyde at 37

C for 10min. Cells were 

491 

washed with warm PBS and centrifuged at 1500 rpm for 5min and resuspended in 150 

L PBS 

492 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

27 

before the samples were read on a BD LSR Fortessa flow cytometer. Data analysis was 

493 

performed using FlowJo software. 

494 

Immunoblotting 

495 

Cells were seeded in 6-well plates (5x10

5

). After 24 hours, cells were lysed in RIPA buffer 

496 

containing 1% DTT and 0.1% protease and phosphatase inhibitor cocktail. Protein content was 

497 

measured by BCA protein assay and absorbances read at 540 nm using a TECAN infinite 

498 

M1000 pro plate reader. Cell lysates were subjected to SDS-PAGE, transferred to 

499 

polyvinylidene difluoride (PVDF) membranes (Merck Millipore). Ponceau staining or 

500 

GAPDH was used as a loading control, and then membranes were immunoblotted for CPT1a. 

501 

Chemiluminescence was performed using Western HRP Substrate and imaged using the Bio-

502 

Rad ChemiDoc MP Imaging System (Bio-Rad Laboratories) and Image Lab software. 

503 

Cell viability 

504 

Cells were seeded in 96-well plates, in 6 well replicates (3x10

3

 cells/well). After 24 hours the 

505 

media was removed and replaced with glucose- and glutamine- free media (Gibco) 

506 

supplemented with 5 mM glucose, 1 mM glutamine and/or 150 

M palmitate, conjugated to 

507 

2% (wt/vol) FA-free BSA. For glucose-deprived conditions DMEM no glucose, no glutamine 

508 

media was supplemented with 1mM glutamine and 150 

M palmitate; for glucose and 

509 

glutamine-deprived conditions media was supplemented with 150 

M palmitate only. Media 

510 

(5 mM glucose, 1 mM glutamine, 150 

M palmitate) supplemented with 10% FBS was used 

511 

as a control. At 0, 6, 12 and 24 hours, 50 

L of MTT reagent was added to wells. MTT reactions 

512 

were incubated at 37°C and 5% CO

2

 for 4 hours, and then stopped by replacing media with 

513 

100 

L DMSO (Sigma–Aldrich) per well. Absorbances were read at 540 nm using a TECAN 

514 

infinite M1000 pro plate reader. For live cell counts, cells were seeded in 6-well plates at 5x10

5

 

515 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

28 

cells/well. After 24 hours, media was replaced as described for the MTT assay, and cells 

516 

counted by Trypan blue (Gibco) exclusion at 6 and 24 hours.

 

517 

U-

13

C stable isotope tracing  

518 

Cells were seeded in 6-well plates (5x10

5

) for 2D culture, or as above for 3D spheroid models. 

519 

After 24 hours, wells were washed with warm PBS and media replaced with 600 

L of DMEM 

520 

no glucose (0 mM), no glutamine (0 mM) media (Gibco), supplemented with 5 mM glucose, 1 

521 

mM glutamine and 150 

M FAs replaced with their respective [U-

13

C] forms (Palmitate (16:0), 

522 

Oleate (18:0), Stearate (18:1), Linoleate (18:2), or Linolenate (18:3)). For glucose-deprived 

523 

conditions, DMEM no glucose, no glutamine media was supplemented with 1mM glutamine 

524 

and [U-

13

C]-150 

M palmitate; for glucose and glutamine-deprived conditions, media was 

525 

supplemented with [U-

13

C]-150 

M palmitate only. FAs were conjugated to 2% (wt/vol) FA-

526 

free BSA containing media for 24 hours at 37

C prior to treating cells. Cells were incubated in 

527 

13

C containing medium for 6 hours, or up to 48 hours for time course experiments. For spheroid 

528 

experiments, 200 

L of spheroids was pooled and collected and centrifuged at 1400 rpm for 5 

529 

min at room temperature. The supernatant medium was collected and used for further 

530 

extracellular substrate experiments. The remaining cell pellet was washed with 2 mL cold PBS 

531 

and centrifuged again. The cell pellet was collected for subsequent LC-MS analysis.  

532 

Extracellular substrate sampling 

533 

Cells were seeded in 6-well plates (5x10

5

) for 2D culture of as above for 3D spheroid models. 

534 

After 24 hours, wells were washed with warm PBS and media replaced with 1 mL or 2 mL of 

535 

DMEM no glucose, no glutamine media (Gibco) supplemented with 5 mM glucose, 1 mM 

536 

glutamine and 150 

M FAs. 100 

L of extracellular media were collected from wells at 3, 6, 

537 

12, 24 hour timepoints for cells incubated in 1 mL of media, and 12, 24 hours for cells incubated 

538 

in 2mL of media. For spheroid experiments, 

13

C labelled medium was collected from cells at 0 

539 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

29 

and 6 hours timepoints. Media samples were centrifuged at 16 000xg for 5 min at 4

C and the 

540 

supernatant collected for subsequent LC-MS analysis. 

541 

Cell lysate extraction for LC-MS 

542 

Media was aspirated from plates and the cells were washed with 2 mL ice-cold 0.9% (w/v) 

543 

NaCl. Cells were scraped in 300 

L of extraction buffer, EB, (1:1 LC/MS methanol:water 

544 

(Optima) + 0.1x internal standards (non-endogenous polar metabolites) and transferred to a 1.5 

545 

mL microcentrifuge tube. A further 300 

L of EB was added to the cells and combined in the 

546 

tube. 600 

L of chloroform (Honeywell) was added before vortexing and incubating on ice for 

547 

10 min. Tubes were vortexed briefly and centrifuged at 15 000g for 10min at 4

C. The aqueous 

548 

upper layer was collected and dried without heat, using a Savant SpeedVac (Thermo Fisher) 

549 

for metabolomics LC-MS analysis. The lower layer was collected and dried under nitrogen 

550 

flow for lipidomics LC-MS analysis. For time course lipidomics experiments, extractions were 

551 

carried out as previously described

70

 by quenching with 0.1M formic acid, neutralising with 

552 

ammonium bicarbonate, and using 40:40:20 acetonitrile:methanol:water with 0.1 M formic 

553 

acid as the solvent system. 

554 

Measuring stable-isotope labelled metabolites by LC-MS 

555 

Dried aqueous upper phase samples were resuspended in 40 

L Amide buffer A (20 mM 

556 

ammonium acetate, 20 mM ammonium hydroxide, 95:5 HPLC H

2

O: Acetonitrile (v/v)) and 

557 

vortexed and centrifuged at 15, 000g for 5min at 4

C. 20

L of supernatant was transferred to 

558 

HPLC vials containing 20 

L acetonitrile for LCMS analysis of amino acids and glutamine 

559 

metabolites. The remaining 20 

L of resuspended sample was transferred to HPLC vials 

560 

containing 20 

L LC-MS H

2

O for LCMS analysis of glycolytic, pentose phosphate pathway 

561 

(PPP), and TCA cycle metabolites. Amino acids and glutamine metabolites were measured 

562 

using Vanquish-TSQ Altis (Thermo) LC-MS/MS system. Analyte separation was achieved 

563 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

30 

using a Poroshell 120 HILIC-Z Column (2.1x150 mm, 2.7 

m) (Agilent) at ambient 

564 

temperature. The pair of buffers used were Amide buffer A and 100% acetonitrile (Buffer B), 

565 

flowed at 200 

L/min; injection volume of 5 

L. Glycolytic, PPP and TCA cycle metabolites 

566 

were measured using 1260 Infinity (Agilent)-QTRAP6500+ (AB Sciex) LC-MS/MS system. 

567 

Analyte separation was achieved using a Synergi 2.5 

m Hydro-RP 100A LC Column (100x2 

568 

mm) at ambient temperature. The pair of buffers used were 95:5 (v/v) water:acetonitrile 

569 

containing 10 mM tributylamine and 15 mM acetic acid (Buffer A) and 100% acetonitrile 

570 

(Buffer B), flowed at 200 

L/min; injection volume of 5 

L. Raw data from both LC-MS/MS 

571 

systems were extracted using MSConvert

69

 and in-house MATLAB scripts. 

572 

Measuring extracellular substrates by LC-MS 

573 

20 

L of collected extracellular media was mixed with 80 

L water, vortexed. 10 

L of diluted 

574 

media was mixed with 90 

L of extraction buffer containing 1:1 (v/v) acetonitrile and methanol 

575 

+ 1x internal standards (non-endogenous standards) at -30

C. The mixture was centrifuged at 

576 

12 000xg for 5 min at 4

C and transferred into HPLC vials for LC-MS analysis measured using 

577 

the Vanquish-TSQ Altis as described above. Raw data from both LC-MS/MS systems were 

578 

extracted using MSConvert

69

 and in-house MATLAB scripts. 

579 

Measuring stable isotope labelled lipids by LC-MS 

580 

Dried lower phase samples were resuspended in 100

L of 4:2:1 (v/v) isopropanol: methanol: 

581 

chloroform containing 7.5 mM ammonium formate. Cardiolipin and acyl-carnitine lipids were 

582 

measured using a Thermo Scientific Q-Exactive-HF-X Hybrid Quadrupole Orbitrap LC-

583 

MS/MS system. 

584 

For cardiolipins, analyte separation was achieved using an Agilent Poroshell 120, EC-C18, 

585 

2.1x150mm, 2.7

m column. The pair of buffers used were 60:40 acetonitrile: water (v/v) with 

586 

10 mM ammonium formate and 0.1% formic acid (mobile phase A), and 90:10 isopropanol: 

587 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

31 

acetonitrile (v/v) with 10 mM ammonium formate and 0.1% formic acid (mobile phase B), 

588 

flowed at 200 

L/min on negative mode. MS1 data was acquired with the following settings: 

589 

3.5kV, capillarity temperature 300

C, 120 000, injection time 100 ms, AGC 1x10

6

, scan range 

590 

1100-1650. For ddMS2 data acquirement, the following settings were used: top 20, resolution 

591 

30,000, 200-2000, isolation 1.0m/z, nce 30, AGC target 1x10

3

, intensity threshold 5x10

3

592 

dynamic exclusion 20 s. Raw data from both LC-MS/MS systems were extracted using 

593 

MSConvert

69

, in-house MATLAB scripts, and MSDial. 

594 

For acyl-carnitines, analyte separation was achieved using an Agilent Poroshell 120, HILIC, 

595 

2.1x100 mm, 2.7 

m column. The pair of buffers used were 0.1% (v/v) formic acid and 10 mM 

596 

ammonium formate in H

2

O and 0.1% (v/v) formic acid in acetonitrile, flowed at 200 

L/min 

597 

on positive mode. MS1 data were acquired with the following settings: 3.5kV, capillarity 

598 

temperature 300

, 120 000, injection time 100ms, AGC 1x10

6

, scan range 200-1000. For 

599 

ddMS2 data acquirement, the following settings were used: top 5, resolution 30,000, 200-2000, 

600 

isolation 1.0m/z, nce 30, AGC target 1x10

3

, intensity threshold 5x10

3

, dynamic exclusion 20s. 

601 

Raw data from both LC-MS/MS systems were extracted using MSConvert

69

 and in-house 

602 

MATLAB scripts. 

603 

Quantifying 13C mass shifts from high-resolution MS1 and MS2 data 

604 

The extraction of mass spectrometry data was performed in MATLAB. First, retention times 

605 

of CL, acyl carnitine, PC and PG species were confirmed using accurate precursor mass and 

606 

MS2 fragmentation patterns. Accurate masses for the monoisotopic and expected mass shifts 

607 

were calculated using ion formulas. The resulting ion cluster was used to visualise the analyte’s 

608 

MS1 chromatogram and to integrate the area under the peak. These extracted MS1 ion 

609 

intensities were then reported as 

13

C fractional enrichments

27

610 

A labelled acyl chain could be incorporated into CLs with or without modification (e.g., 

611 

elongation or desaturation). Thus, MS2 data was used to verify localisation of 

13

C-FA among 

612 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

32 

the acyl chains. From the MS2 data we can derive the ratio of labelled to unlabelled product 

613 

ions fragmented from a labelled precursor. Using relative ion intensities of labelled and 

614 

unlabelled acyl anion [FA - H]

-

, the ion intensity of a labelled precursor (from MS1) is 

615 

apportioned to its constituent acyl chains in terms of their labelled and unlabelled forms. The 

616 

percentage labelling of a given acyl chain among CL or PG pools was then calculated by 

617 

summing up these labelled and unlabelled constituents across all quantified CL or PG species. 

618 

Quantification and statistical analysis 

619 

Student’s t-test, one- or two-way ANOVA with Dunnett’s or Tukey’s multiple comparisons, 

620 

Nonparametric Wilcoxon signed rank test, and linear regression statistical tests were performed 

621 

with GraphPad Prism 9.0 (GraphPad Software). P < 0.05 was considered significant. Data are 

622 

reported as mean ± standard error of the mean (SEM), standard deviation (SD), or min to max 

623 

values of at least three independent determinations as indicated in figure legends. Schematic 

624 

diagrams were created with Biorender.com. 

 

625 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

33 

Extended Data 

626 

 

Extended Data Fig. 1. Palmitate is a minor carbon source to the TCA cycle.

 

627 

(A)

 

Percentage of glucose, glutamine, and FA 16:0 relative to starting amount, and amount of lactate in culture 

628 

media over 24 hours in 1mL culture medium. N=9 per cell line. 

629 

(B)

 

Percentage of glucose and glutamine relative to starting amount in 1 mL (blue) or 2 mL (red) culture medium. 

630 

N=9 per cell line. 

631 

(C)

 

Fractional contribution of [U-

13

C]-FA16:0 to citrate, oxoglutarate, succinate, fumarate, and malate. N=9 per 

632 

cell line. Statistically significant differences between PCa cells to PNT1 cells are shown. 

633 

(D)

 

Correlations of [U-

13

C]-FA 16:0 to citrate (plotted), oxoglutarate, succinate, fumarate, and malate to FAO. 

634 

Error bars represent 

 SD. P and r

2

 values by linear regression. 

635 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

34 

(E)

 

Correlations of [U-

13

C]-FA 16:0 to citrate (plotted), oxoglutarate, succinate, fumarate, and malate to active 

636 

mitochondria content. Error bars represent 

 SD. P and r

2

 values by linear regression. 

637 

(F)

 

Fractional contribution of [U-

13

C]-glucose to pyruvate (red), [U-

13

C]-glutamine to intracellular glutamine 

638 

(blue), [U-

13

C]-FA 16:0 to palmitoyl-carnitine (orange) after 6 hours of labelling in LNCaP, MR42D and 

639 

MR49F cells. N=9 per cell line. 

640 

(G)

 

Fractional contributions of [U-

13

C]-glucose, glutamine, or FA 16:0 to citrate, oxoglutarate and malate. N=9 

641 

per cell line. 

642 

(H)

 

Percentage of glucose, glutamine, and FA 16:0 m0 and m16 relative to starting amount, comparing unlabelled 

643 

and [U-

13

C]-FA 16:0 media at 0 and 6 hours in representative C4-2B spheroids. N=3 per group. 

644 

(I)

 

Fractional contribution of [U-

13

C]-FA 16:0 to palmitoyl-carnitine in 3D spheroid LNCaP, C4-2B, and 22RV1 

645 

models. N=4-5 per cell line. 

646 

(J)

 

Fractional contribution of [U-

13

C]-FA 16:0 to succinate, fumarate, and malate in 3D spheroid models. N=3 

647 

per cell line.  

648 

Graphs show mean 

 SEM unless stated otherwise. *p<0.05 by One-way ANOVA with Dunnett’s multiple 

649 

comparisons test, unless stated otherwise.

 

 

650 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

35 

 

Extended Data Fig. 2. Palmitate cannot maintain TCA cycle activity under glucose and 

651 

glutamine deprivation. 

652 

(A)

 

[1-

14

C]-FA 16:0 oxidation rates in the presence of 5 mM, 3 mM, 1 mM, or 0 mM glucose. N=3 per cell line.  

653 

* compared to 5 mM glucose, p<0.05 by One-way ANOVA with Dunnett’s multiple comparisons test. 

654 

(B)

 

Impact of glucose- or glucose and glutamine- deprivation on metabolic activity assessed by MTT assay to 48 

655 

hours, normalised to day 0 results. Dotted line at 24 hours indicates replenishment of media. 

656 

(C)

 

Impact of glucose- or glucose and glutamine- deprivation on cell viability assessed by live cell counts at 0, 6 

657 

and 24 hours normalised to day 0 results. 

658 

(D)

 

Relative abundance (normalised to +glucose/+glutamine) of [U-

13

C]-FA 16:0 to oxoglutarate or malate. N=3-

659 

9 per cell line.  

660 

Graphs show mean 

 SEM. *compared to +glucose/+glutamine, ^compared to -glucose/+glutamine, p<0.05 by 

661 

One-way ANOVA with Dunnett’s multiple comparisons test.

 

662 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

36 

 

663 

Extended Data Fig. 3. LCFAs undergo incomplete FAO in prostate cells.

 

664 

Fractional contributions of [U-

13

C]-FA 16:0, 18:0, 18:2 to shorter chain acyl-carnitine species. N=9 per cell line. 

665 

Graphs show mean 

 SEM.

 

 

666 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

37 

 

667 

Extended Data Fig, 4. FA 18:2 predominantly enriches CL pools compared to other FAs. 

668 

(A)

 

Enrichment fraction of PC 32:1 and 34:1 in PC3 and C4-2B cells over 48 hours exposure to[U-

13

C]-FA 

669 

16:0. N=1 per cell line. 

670 

(B)

 

Comparison of fractional contribution of [U-

13

C]-FA 16:0 to PC and CL species at 6 hours in PC3 and C4-

671 

2B cells. N=3-9 per cell line. P<0.05 by two-way ANOVA with Tukey’s multiple comparisons test. 

672 

(C)

 

Fractional contribution of [U-

13

C]-FA 16:0, 18:0, 18:1, 18:2 (150 

M) to CL species containing the same 

673 

acyl chains at 6 hours. N=3 per cell line. 

674 

(D)

 

Overall percentage of fatty acyls (of the same length and saturation as labelled substrate) contained in CL 

675 

species labelled by [U-

13

C]-FA 16:0, 18:0, 18:1 or 18:2. Unlabelled 

12

C controls show level of background 

676 

labelling. N=3 per cell line. 

677 

Graphs show mean 

 SEM unless stated otherwise.

 

678 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

38 

 

679 

Extended Data Fig. 5. CPT1a

KD

 reduces FA 18:2 incorporation to CLs. 

680 

(A)

 

Percentage of 18:2 or 16:0 acyl chain in CLs respectively labelled by [U-

13

C]-FA 18:2 or 16:0 in scrambled 

681 

(scr) and siCPT1a cells. N=9 per group. 

682 

(B)

 

Fractional contribution of [U-

13

C]-FA 18:2 or 16:0 to 18:2 or 16:0 acyl chains of CL species. N=9 per group. 

683 

(C)

 

Delta differences between fractional contribution of [U-

13

C]-FA 18:2 or 16:0 to acyl chains of CL species. 

684 

N=7-8 per group. Statistically significant differences within [U-

13

C]-FA 18:2 or 16:0 groups to 0 difference, 

685 

determined by nonparametric Wilcoxon signed rank test. Error bars represent 

 min to max values. 

686 

Graphs show mean 

 SEM unless stated otherwise. ns, not significant, *p<0.05 by unpaired t-tests between 

687 

siCPT1a and scr control groups unless stated otherwise.

 

 

688 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

39 

Acknowledgements 

689 

N.T.S. is supported by the Australian Rotary Health/Rotary Club of Blacktown City ‘Mel Grey’ 

690 

PhD scholarship. L.M.B. and A.J.H. acknowledge grant support from The Movember 

691 

Foundation/Prostate Cancer Foundation of Australia (MRTA3). A.J.H. is supported by a 

692 

Robinson Fellowship from the University of Sydney and funding from the University of 

693 

Sydney. L.M.B. is supported by a Principal Cancer Research Fellowship produced with the 

694 

financial and other support of Cancer Council SA’s Beat Cancer Project on behalf of its donors 

695 

and the State Government of South Australia through the Department of Health and was 

696 

supported by a Future Fellowship from the Australian Research Council (FT130101004). We 

697 

thank the Sydney Mass Spectrometry facility for access to LC-MS instruments; Julia Scott and 

698 

Dr Zeyad Nassar in the Butler lab (University of Adelaide) for assistance with 3D spheroid and 

699 

mitotracker experiment development; Dr Helen McGuire and the Sydney Flow Cytometry 

700 

facility for assistance with mitotracker experiments; and Prof. Lisa Horvath and members of 

701 

the Butler lab for scientific discussions. 

702 

Author contributions 

703 

Conceptualisation by A.J.H. and L.E.Q. Project administration, visualisation, writing-original 

704 

draft by N.T.S. Investigation by N.T.S., M.F.H.S., and A.W. Methodology by L.E.Q. Software 

705 

by N.T.S. and L.E.Q. Formal analysis by N.T.S., L.E.Q and A.J.H. Writing-editing & reviewing 

706 

by N.T.S., L.E.Q., A.J.H., and L.M.B. Funding acquisition, resources by A.J.H. Supervision by 

707 

A.J.H., L.E.Q., and L.M.B. 

708 

Declaration of interests 

709 

The authors declare that they have no competing interests.

 

 

710 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

40 

References:

 

711 

Swinnen, J. V., Esquenet, M., Goossens, K., Heyns, W. & Verhoeven, G. Androgens 

712 

stimulate fatty acid synthase in the human prostate cancer cell line LNCaP. 

Cancer Res.

 

713 

57

, 1086-1090 (1997). 

714 

2 Yue, 

S.

 et al.

 Cholesteryl ester accumulation induced by PTEN loss and PI3K/AKT 

715 

activation underlies human prostate cancer aggressiveness. 

Cell Metab.

 

19

, 393-406, 

716 

doi:10.1016/j.cmet.2014.01.019 (2014). 

717 

3 Migita, 

T.

 et al.

 Fatty acid synthase: a metabolic enzyme and candidate oncogene in 

718 

prostate cancer. 

J. Natl. Cancer Inst.

 

101

, 519-532, doi:10.1093/jnci/djp030 (2009). 

719 

Nassar, Z. D.

 et al.

 Human DECR1 is an androgen-repressed survival factor that 

720 

regulates PUFA oxidation to protect prostate tumor cells from ferroptosis. 

Elife

 

9

721 

doi:10.7554/eLife.54166 (2020). 

722 

5 Balaban, 

S.

 et al.

 Extracellular Fatty Acids Are the Major Contributor to Lipid 

723 

Synthesis in Prostate Cancer. 

Mol. Cancer Res.

 

17

, 949-962, doi:10.1158/1541-

724 

7786.Mcr-18-0347 (2019). 

725 

6 Fidelito, 

G.

 et al.

 Multi-substrate Metabolic Tracing Reveals Marked Heterogeneity and 

726 

Dependency on Fatty Acid Metabolism in Human Prostate Cancer. 

Mol. Cancer Res.

 

727 

21

, 359-373, doi:10.1158/1541-7786.Mcr-22-0796 (2023). 

728 

7 Iglesias-Gato, 

D.

 et al.

 The Proteome of Primary Prostate Cancer. 

Eur. Urol.

 

69

, 942-

729 

952, doi:10.1016/j.eururo.2015.10.053 (2016). 

730 

Butler, L. M.

 et al.

 Lipidomic Profiling of Clinical Prostate Cancer Reveals Targetable 

731 

Alterations in Membrane Lipid Composition. 

Cancer Res.

 

81

, 4981-4993, 

732 

doi:10.1158/0008-5472.CAN-20-3863 (2021). 

733 

Ahmad, F., Cherukuri, M. K. & Choyke, P. L. Metabolic reprogramming in prostate 

734 

cancer. 

Br. J. Cancer

 

125

, 1185-1196, doi:10.1038/s41416-021-01435-5 (2021). 

735 

10 

Costello, L. C. & Franklin, R. B. The intermediary metabolism of the prostate: a key to 

736 

understanding the pathogenesis and progression of prostate malignancy. 

Oncology

 

59

737 

269-282, doi:10.1159/000012183 (2000). 

738 

11 

Liu, Y. Fatty acid oxidation is a dominant bioenergetic pathway in prostate cancer. 

739 

Prostate Cancer Prostatic Dis.

 

9

, 230-234, doi:10.1038/sj.pcan.4500879 (2006). 

740 

12 

Liu, Y., Zuckier, L. S. & Ghesani, N. V. Dominant uptake of fatty acid over glucose by 

741 

prostate cells: a potential new diagnostic and therapeutic approach. 

Anticancer Res.

 

30

742 

369-374 (2010). 

743 

13 

Van Heijster, F. H. A., Breukels, V., Jansen, K. F. J., Schalken, J. A. & Heerschap, A. 

744 

Carbon sources and pathways for citrate secreted by human prostate cancer cells 

745 

determined by NMR tracing and metabolic modeling. 

Proc. Natl. Acad. Sci. U. S. A.

 

746 

119

, doi:10.1073/pnas.2024357119 (2022). 

747 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

41 

14 Dueregger, 

A.

 et al.

 Differential Utilization of Dietary Fatty Acids in Benign and 

748 

Malignant Cells of the Prostate. 

PLoS One

 

10

, e0135704, 

749 

doi:10.1371/journal.pone.0135704 (2015). 

750 

15 

Mah, C. Y., Nassar, Z. D., Swinnen, J. V. & Butler, L. M. Lipogenic effects of androgen 

751 

signaling in normal and malignant prostate. 

Asian J. Urol

 

7

, 258-270, 

752 

doi:

https://doi.org/10.1016/j.ajur.2019.12.003

 (2020). 

753 

16 

Fritz, I. B. & Yue, K. T. Long-chain carnitine acyltransferase and the role of 

754 

acylcarnitine derivatives in the catalytic increase of fatty acid oxidation induced by 

755 

carnitine. 

J. Lipid Res.

 

4

, 279-288 (1963). 

756 

17 

Carracedo, A., Cantley, L. C. & Pandolfi, P. P. Cancer metabolism: fatty acid oxidation 

757 

in the limelight. 

Nat. Rev. Cancer

 

13

, 227-232, doi:10.1038/nrc3483 (2013). 

758 

18 

Ricciardi, M. R.

 et al.

 Targeting the leukemia cell metabolism by the CPT1a inhibition: 

759 

functional preclinical effects in leukemias. 

Blood

 

126

, 1925-1929, doi:10.1182/blood-

760 

2014-12-617498 (2015). 

761 

19 

Schlaepfer, I. R.

 et al.

 Lipid Catabolism via CPT1 as a Therapeutic Target for Prostate 

762 

Cancer. 

Mol. Cancer Ther.

 

13

, 2361-2371, doi:10.1158/1535-7163.mct-14-0183 

763 

(2014). 

764 

20 

Pike, L. S., Smift, A. L., Croteau, N. J., Ferrick, D. A. & Wu, M. Inhibition of fatty acid 

765 

oxidation by etomoxir impairs NADPH production and increases reactive oxygen 

766 

species resulting in ATP depletion and cell death in human glioblastoma cells. 

Biochim 

767 

Biophys Acta

 

1807

, 726-734, doi:

https://doi.org/10.1016/j.bbabio.2010.10.022

 (2011). 

768 

21 Yao, 

C.-H.

 et al.

 Identifying off-target effects of etomoxir reveals that carnitine 

769 

palmitoyltransferase I is essential for cancer cell proliferation independent of 

β

-

770 

oxidation. 

PLoS Biol.

 

16

, e2003782, doi:10.1371/journal.pbio.2003782 (2018). 

771 

22 

O'Connor, R. S.

 et al.

 The CPT1a inhibitor, etomoxir induces severe oxidative stress at 

772 

commonly used concentrations. 

Sci. Rep.

 

8

, 6289, doi:10.1038/s41598-018-24676-6 

773 

(2018). 

774 

23 He, 

L.

 et al.

 Carnitine palmitoyltransferase-1b deficiency aggravates pressure overload-

775 

induced cardiac hypertrophy caused by lipotoxicity. 

Circulation

 

126

, 1705-1716, 

776 

doi:10.1161/circulationaha.111.075978 (2012). 

777 

24 

Carta, G., Murru, E., Banni, S. & Manca, C. Palmitic Acid: Physiological Role, 

778 

Metabolism and Nutritional Implications. 

Front. Physiol.

 

8

, 902, 

779 

doi:10.3389/fphys.2017.00902 (2017). 

780 

25 

Ahn, W. S. & Antoniewicz, M. R. Parallel labeling experiments with [1,2-13C]glucose 

781 

and [U-13C]glutamine provide new insights into CHO cell metabolism. 

Metab Eng.

 

15

782 

34-47, doi:

https://doi.org/10.1016/j.ymben.2012.10.001

 (2013). 

783 

26 Duan, 

L.

 et al.

 13C tracer analysis suggests extensive recycling of endogenous CO2 in 

784 

vivo. 

Cancer Metab.

 

10

, 11, doi:10.1186/s40170-022-00287-8 (2022). 

785 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

42 

27 

Buescher, J. M.

 et al.

 A roadmap for interpreting (13)C metabolite labeling patterns 

786 

from cells. 

Curr. Opin. Biotechnol.

 

34

, 189-201, doi:10.1016/j.copbio.2015.02.003 

787 

(2015). 

788 

28 

Bishop, J. L.

 et al.

 The Master Neural Transcription Factor BRN2 Is an Androgen 

789 

Receptor–Suppressed Driver of Neuroendocrine Differentiation in Prostate Cancer. 

790 

Cancer Discov.

 

7

, 54-71, doi:10.1158/2159-8290.Cd-15-1263 (2017). 

791 

29 Joshi, 

M.

 et al.

 CPT1A Supports Castration-Resistant Prostate Cancer in Androgen-

792 

Deprived Conditions. 

Cells

 

8

, 1115 (2019). 

793 

30 

Nassar, Z. D.

 et al.

 Fatty Acid Oxidation Is an Adaptive Survival Pathway Induced in 

794 

Prostate Tumors by HSP90 Inhibition. 

Mol. Cancer Res.

 

18

, 1500-1511, 

795 

doi:10.1158/1541-7786.mcr-20-0570 (2020). 

796 

31 

Sutherland, R. M., McCredie, J. A. & Inch, W. R. Growth of Multicell Spheroids in 

797 

Tissue Culture as a Model of Nodular Carcinomas2. 

J Natl Cancer Inst.

 

46

, 113-120, 

798 

doi:10.1093/jnci/46.1.113 (1971). 

799 

32 Jones, 

D. 

T.

 et al.

 3D Growth of Cancer Cells Elicits Sensitivity to Kinase Inhibitors 

800 

but Not Lipid Metabolism Modifiers. 

Mol. Cancer Ther.

 

18

, 376-388, 

801 

doi:10.1158/1535-7163.MCT-17-0857 (2019). 

802 

33 

Tidwell, T. R., Røsland, G. V., Tronstad, K. J., Søreide, K. & Hagland, H. R. Metabolic 

803 

flux analysis of 3D spheroids reveals significant differences in glucose metabolism 

804 

from matched 2D cultures of colorectal cancer and pancreatic ductal adenocarcinoma 

805 

cell lines. 

Cancer Metab.

 

10

, 9, doi:10.1186/s40170-022-00285-w (2022). 

806 

34 Ma, 

Y.

 et al.

 Functional analysis of molecular and pharmacological modulators of 

807 

mitochondrial fatty acid oxidation. 

Sci. Rep.

 

10

, 1450, doi:10.1038/s41598-020-58334-

808 

7 (2020). 

809 

35 

Sullivan, Lucas B.

 et al.

 Supporting Aspartate Biosynthesis Is an Essential Function of 

810 

Respiration in Proliferating Cells. 

Cell

 

162

, 552-563, 

811 

doi:

https://doi.org/10.1016/j.cell.2015.07.017

 (2015). 

812 

36 Le, 

A.

 et al.

 Glucose-Independent Glutamine Metabolism via TCA Cycling for 

813 

Proliferation and Survival in B Cells. 

Cell Metab.

 

15

, 110-121, 

814 

doi:10.1016/j.cmet.2011.12.009 (2012). 

815 

37 Koves, 

T. 

R.

 et al.

 Peroxisome Proliferator-activated Receptor-

γ

 Co-activator 1

α

-

816 

mediated Metabolic Remodeling of Skeletal Myocytes Mimics Exercise Training and 

817 

Reverses Lipid-induced Mitochondrial Inefficiency. 

J. Biol. Chem.

 

280

, 33588-33598, 

818 

doi:10.1074/jbc.m507621200 (2005). 

819 

38 Koves, 

T. 

R.

 et al.

 Mitochondrial Overload and Incomplete Fatty Acid Oxidation 

820 

Contribute to Skeletal Muscle Insulin Resistance. 

Cell Metab.

 

7

, 45-56, 

821 

doi:10.1016/j.cmet.2007.10.013 (2008). 

822 

39 

Colbeau, A., Nachbaur, J. & Vignais, P. M. Enzymic characterization and lipid 

823 

composition of rat liver subcellular membranes. 

Biochim Biophys Acta

 

249

, 462-492, 

824 

doi:10.1016/0005-2736(71)90123-4 (1971). 

825 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

43 

40 

Falabella, M., Vernon, H. J., Hanna, M. G., Claypool, S. M. & Pitceathly, R. D. S. 

826 

Cardiolipin, Mitochondria, and Neurological Disease. 

Trends Endocrinol Metab.

 

32

827 

224-237, doi:10.1016/j.tem.2021.01.006 (2021). 

828 

41 

Ikon, N. & Ryan, R. O. Cardiolipin and mitochondrial cristae organization. 

Biochem 

829 

Biophys Acta Biomembr.

 

1859

, 1156-1163, 

830 

doi:

https://doi.org/10.1016/j.bbamem.2017.03.013

 (2017). 

831 

42 

Schwall, C. T., Greenwood, V. L. & Alder, N. N. The stability and activity of respiratory 

832 

Complex II is cardiolipin-dependent. 

Biochim Biophys Acta

 

1817

, 1588-1596, 

833 

doi:

https://doi.org/10.1016/j.bbabio.2012.04.015

 (2012). 

834 

43 Dudek, 

J.

 et al.

 Cardiac-specific succinate dehydrogenase deficiency in Barth 

835 

syndrome. 

EMBO Mol. Med.

 

8

, 139-154, doi:10.15252/emmm.201505644 (2016). 

836 

44 

Eble, K. S., Coleman, W. B., Hantgan, R. R. & Cunningham, C. C. Tightly associated 

837 

cardiolipin in the bovine heart mitochondrial ATP synthase as analyzed by 31P nuclear 

838 

magnetic resonance spectroscopy. 

J. Biol. Chem.

 

265

, 19434-19440 (1990). 

839 

45 

Taylor, W. A. & Hatch, G. M. Identification of the human mitochondrial linoleoyl-

840 

coenzyme A monolysocardiolipin acyltransferase (MLCL AT-1). 

J. Biol. Chem.

 

284

841 

30360-30371, doi:10.1074/jbc.M109.048322 (2009). 

842 

46 

Zachman, D. K.

 et al.

 The role of calcium-independent phospholipase A2 in cardiolipin 

843 

remodeling in the spontaneously hypertensive heart failure rat heart. 

J. Lipid Res.

 

51

844 

525-534, doi:

https://doi.org/10.1194/jlr.M000646

 (2010). 

845 

47 

Xu, Y. & Schlame, M. The turnover of glycerol and acyl moieties of cardiolipin. 

Chem. 

846 

Phys. Lipids

 

179

, 17-24, doi:

https://doi.org/10.1016/j.chemphyslip.2013.10.005

 

847 

(2014). 

848 

48 

Landriscina, C., Megli, F. M. & Quagliariello, E. Turnover of fatty acids in rat liver 

849 

cardiolipin: Comparison with other mitochondrial phospholipids. 

Lipids

 

11

, 61-66, 

850 

doi:10.1007/BF02532585 (1976). 

851 

49 Wahjudi, 

P. 

N.

 et al.

 Turnover of nonessential fatty acids in cardiolipin from the rat 

852 

heart. 

J. Lipid Res.

 

52

, 2226-2233, doi:10.1194/jlr.M015966 (2011). 

853 

50 

Nagarajan, S. R., Butler, L. M. & Hoy, A. J. The diversity and breadth of cancer cell 

854 

fatty acid metabolism. 

Cancer Metab

 

9

, 2, doi:10.1186/s40170-020-00237-2 (2021). 

855 

51 Oemer, 

G.

 et al.

 Fatty acyl availability modulates cardiolipin composition and alters 

856 

mitochondrial function in HeLa cells. 

J. Lipid Res.

 

62

, 100111, 

857 

doi:10.1016/j.jlr.2021.100111 (2021). 

858 

52 

Hoch, F. L. Cardiolipins and biomembrane function. 

Biochimica et Biophysica Acta 

859 

(BBA)

 

1113

, 71-133, doi:10.1016/0304-4157(92)90035-9 (1992). 

860 

53 

Cao, J., Liu, Y., Lockwood, J., Burn, P. & Shi, Y. A Novel Cardiolipin-remodeling 

861 

Pathway Revealed by a Gene Encoding an Endoplasmic Reticulum-associated Acyl-

862 

CoA:Lysocardiolipin Acyltransferase (ALCAT1) in Mouse. 

J. Biol. Chem.

 

279

, 31727-

863 

31734, doi:10.1074/jbc.m402930200 (2004). 

864 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

44 

54 

Lu, Y.-W. & Claypool, S. M. Disorders of phospholipid metabolism: an emerging class 

865 

of mitochondrial disease due to defects in nuclear genes. 

Frontiers in Genetics

 

6

866 

doi:10.3389/fgene.2015.00003 (2015). 

867 

55 

Xu, Y., Malhotra, A., Ren, M. & Schlame, M. The Enzymatic Function of Tafazzin*. 

J. 

868 

Biol. Chem.

 

281

, 39217-39224, doi:

https://doi.org/10.1074/jbc.M606100200

 (2006). 

869 

56 

Hostetler, K. Y., Galesloot, J. M., Boer, P. & Van Den Bosch, H. Further studies on the 

870 

formation of cardiolipin and phosphatidylglycerol in rat liver mitochondria. Effect of 

871 

divalent cations and the fatty acid composition of CDP-diglyceride. 

Biochim Biophys 

872 

Acta

 

380

, 382-389, doi:10.1016/0005-2760(75)90106-x (1975). 

873 

57 Zadra, 

G.

 et al.

 Inhibition of de novo lipogenesis targets androgen receptor signaling in 

874 

castration-resistant prostate cancer. 

Proc. Natl. Acad. Sci. U. S. A.

 

116

, 631-640, 

875 

doi:10.1073/pnas.1808834116 (2019). 

876 

58 Parida, 

P. 

K.

 et al.

 Limiting mitochondrial plasticity by targeting DRP1 induces 

877 

metabolic reprogramming and reduces breast cancer brain metastases. 

Nature Cancer

 

878 

4

, 893-907, doi:10.1038/s43018-023-00563-6 (2023). 

879 

59 

Yang, J. H.

 et al.

 Snail augments fatty acid oxidation by suppression of mitochondrial 

880 

ACC2 during cancer progression. 

Life Science Alliance

 

3

, e202000683, 

881 

doi:10.26508/lsa.202000683 (2020). 

882 

60 

Flaig, T. W.

 et al.

 Lipid catabolism inhibition sensitizes prostate cancer cells to 

883 

antiandrogen blockade. 

Oncotarget

 

8

, 56051-56065, doi:10.18632/oncotarget.17359 

884 

(2017). 

885 

61 Zhang, 

H.

 et al.

 Lipidomics reveals carnitine palmitoyltransferase 1C protects cancer 

886 

cells from lipotoxicity and senescence. 

J Pharm Anal

 

11

, 340-350, 

887 

doi:10.1016/j.jpha.2020.04.004 (2021). 

888 

62 

Paradies, G., Paradies, V., De Benedictis, V., Ruggiero, F. M. & Petrosillo, G. 

889 

Functional role of cardiolipin in mitochondrial bioenergetics. 

Biochim Biophys Acta

 

890 

1837

, 408-417, doi:

https://doi.org/10.1016/j.bbabio.2013.10.006

 (2014). 

891 

63 Zhong, 

H.

 et al.

 Mitochondrial control of apoptosis through modulation of cardiolipin 

892 

oxidation in hepatocellular carcinoma: A novel link between oxidative stress and 

893 

cancer. 

Free Radic. Biol. Med.

 

102

, 67-76, 

894 

doi:

https://doi.org/10.1016/j.freeradbiomed.2016.10.494

 (2017). 

895 

64 Faubert, 

B.

 et al.

 Lactate Metabolism in Human Lung Tumors. 

Cell

 

171

, 358-371.e359, 

896 

doi:10.1016/j.cell.2017.09.019 (2017). 

897 

65 Neinast, 

M. 

D.

 et al.

 Quantitative Analysis of the Whole-Body Metabolic Fate of 

898 

Branched-Chain Amino Acids. 

Cell Metab.

 

29

, 417-429.e414, 

899 

doi:

https://doi.org/10.1016/j.cmet.2018.10.013

 (2019). 

900 

66 

Kiebish, M. A., Han, X., Cheng, H., Chuang, J. H. & Seyfried, T. N. Cardiolipin and 

901 

electron transport chain abnormalities in mouse brain tumor mitochondria: lipidomic 

902 

evidence supporting the Warburg theory of cancer. 

J. Lipid Res.

 

49

, 2545-2556, 

903 

doi:10.1194/jlr.M800319-JLR200 (2008). 

904 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint 

2024.01.16.575611v1.full-html.html
background image

45 

67 

Nyquist, M. D.

 et al.

 TALEN-engineered AR gene rearrangements reveal endocrine 

905 

uncoupling of androgen receptor in prostate cancer. 

Proc. Natl. Acad. Sci. U. S. A.

 

110

906 

17492-17497, doi:doi:10.1073/pnas.1308587110 (2013). 

907 

68 Li, 

Y.

 et al.

 Androgen receptor splice variants mediate enzalutamide resistance in 

908 

castration-resistant prostate cancer cell lines. 

Cancer Res.

 

73

, 483-489, 

909 

doi:10.1158/0008-5472.Can-12-3630 (2013). 

910 

69 Chambers, 

M. 

C.

 et al.

 A cross-platform toolkit for mass spectrometry and proteomics. 

911 

Nat. Biotechnol.

 

30

, 918-920, doi:10.1038/nbt.2377 (2012). 

912 

70 Lu, 

W.

 et al.

 Metabolite Measurement: Pitfalls to Avoid and Practices to Follow. 

Annu. 

913 

Rev. Biochem.

 

86

, 277-304, doi:10.1146/annurev-biochem-061516-044952 (2017). 

914 

 

915 

.

CC-BY-NC-ND 4.0 International license

perpetuity. It is made available under a

preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in 

The copyright holder for this

this version posted January 20, 2024. 

https://doi.org/10.1101/2024.01.16.575611

doi: 

bioRxiv preprint