A surface proton antenna in carbonic anhydrase II supports ... - eLife

0 downloads 0 Views 4MB Size Report
colocalized with the monocarboxylate transporter MCT1 in MCF-7 breast cancer ... Lactate efflux from cancer cells is mediated primarily by the monocarboxylate.
1

A surface proton antenna in carbonic anhydrase II supports lactate transport in cancer cells

2

Sina Ibne Noor1, Somayeh Jamali1, Samantha Ames1, Silke Langer1, Joachim W. Deitmer1

3

& Holger M. Becker1,2

4 5

1

6

Kaiserslautern, Germany; 2Institute of Physiological Chemistry, University of Veterinary Medicine

7

Hannover, D-30559 Hannover, Germany

8

For correspondence: [email protected]

Division of General Zoology, Department of Biology, University of Kaiserslautern, D-67653

9 10

Abstract

11

Many tumor cells produce vast amounts of lactate and acid, which have to be removed from the cell

12

to avoid intracellular lactacidosis and suffocation of metabolism. In the present study, we show that

13

proton-driven lactate flux is enhanced by the intracellular carbonic anhydrase CAII, which is

14

colocalized with the monocarboxylate transporter MCT1 in MCF-7 breast cancer cells. Co-expression

15

of MCTs with various CAII mutants in Xenopus oocytes demonstrated that CAII facilitates MCT

16

transport activity via CAII-Glu69 and CAII-Asp72, which could function as surface proton antennae for

17

the enzyme. While CAII-Glu69 and CAII-Asp72 seem to mediate proton transfer between enzyme and

18

transporter, CAII-His

19

involved in proton shuttling between the two proteins, but mediates binding between MCT and CAII.

20

Taken together, the results suggest that CAII features a moiety that exclusively mediates proton

21

exchange with the MCT to facilitate transport activity.

, the e tral residue of the e z

e s i tra ole ular proto shuttle, is not

1

22

Introduction

23

Many tumor cells, especially those which reside in a hypoxic environment, rely on glycolysis to meet

24

their increased demand for energy and biosynthetic precursors (Hanahan & Weinberg, 2011; Schulze

25

& Harris, 2012; Parks et al., 2013). Thereby they produce considerable amounts of acid and lactate,

26

which have to be removed from the cell to avoid intracellular lactacidosis and suffocation of their

27

metabolism. Lactate efflux from cancer cells is mediated primarily by the monocarboxylate

28

transporters MCT1 (SLC16A1) and MCT4 (SLC16A3), both of which transport lactate together with a

29

proton across the cell membrane (Poole & Halestrap, 1993; Parks et al., 2013). This MCT-mediated H+

30

efflux can exacerbate extracellular acidification and support the formation of a hostile environment,

31

which favors tumor growth and allows tumor cells to escape conventional cancer therapies (Kennedy

32

& Dewhirst, 2010; Parks et al., 2011, 2013; Gillies et al., 2012). While expression of MCT1 is not

33

altered with varying oxygen tension, expression of MCT4 has been shown to be upregulated in

34

different cancer cells under hypoxia (Ullah et al., 2006). However, expression of the two isoforms

35

strongly varies among different tumor species and cell lines. In the MCF-7 breast cancer cell line,

36

used in this study, lactate flux is exclusively mediated by MCT1 under both normoxic and hypoxic

37

conditions (Jamali et al., 2015).

38

Another family of key proteins in tumor acid/base regulation is carbonic anhydrases, which catalyze

39

the reversible hydration of CO2 to HCO3- + H+. While the role of the cancer-specific isoform CAIX in

40

tumor development and progression has been extensively studied, physiological data on CAII in

41

cancer cells are still relatively scarce. CAII was found in different types of brain tumors, with the most

42

malignant species exhibiting the strongest expression, and was linked to poor prognosis in patients

43

suffering from those tumor types (Parkkila et al., 1995; Haapasalo et al., 2007). Furthermore, CAII

44

was suggested to mediate malignant behavior of pulmonary neuroendocrine tumors (Zhou et al.,

45

2015). Experiments on breast cancer cells demonstrated that CAII is upregulated in the highly

46

tumorigenic MDA-MB-231 cell line, when the cells are exposed to the chemotherapeutic drug

47

doxorubicin (Mallory et al., 2005). On the other hand, a reduction in CAII expression was proposed to 2

48

promote tumor cell motility and contribute to tumor growth and metastasis in non-small cell lung

49

cancer and gastric carcinoma (Chiang et al., 2002; Li et al., 2012) and was found to be associated with

50

tumor differentiation and poor prognosis in patients with pancreatic cancer (Sheng et al., 2013),

51

suggesting a differential function of CAII in different tumor types.

52

Catalysis in CA occurs in two distinct and separate stages: the interconversion of CO2 and HCO3-

53

followed by the transfer of an H+ to the bulk solution to regenerate the zinc-bound hydroxide, the

54

latter of which being the rate-limiting step in the overall reaction (Tu et al., 1989; Lindskok, 1997). In

55

CAII, the fastest of the α-CAs (Steiner et al., 1975; Boone et al., 2014), H+ transfer between the zinc-

56

bound water and the solvent surrounding the enzyme is facilitated by the side chain of His64, which

57

shuttles H+ between the bulk solvent and a network of well-ordered hydrogen-bonded water

58

ole ules i the e z

e s a ti e-site cavity (Fisher et al., 2007a). Furthermore, H+ transfer efficiency

59

of His64 is fine-tuned by several hydrophilic residues (Tyr7, Asn62 and Asn67), which contribute to

60

the stabilization of the intervening water molecules within the active-site cavity and influence

61

orientation and the pKa value of His64 (Fisher et al., 2007b). This makes the residue His64 essential

62

for CAII to reach full enzymatic activity. By using an algorithm for mapping proton wires in proteins,

63

Shinobu & Agmon (2009) proposed that the active site H+ wire exits to the protein surface, and leads

64

to Glu69 and Asp72, located on an electronegative patch on the rim of the active site cavity. Based

65

on that observation, the authors proposed that positively charged, protonated buffer molecules dock

66

in that area, from which a proton is delivered to the active site when the enzyme works in the

67

dehydration direction.

68

Cytosolic CAII has been found to facilitate transport function of various acid/base transporters

69

including the Cl-/HCO3- exchangers AE1 and AE2 (Vince & Reithmeier, 1998, 2000; Sterling et al.,

70

2001), the Na+/HCO3- cotransporters NBCe1 and NBCn1 (Gross et al., 2002; Loiselle et al., 2004;

71

Becker et al., 2007), and the Na+/H+ exchanger NHE1 (Li et al., 2002, 2006). Augmentation of

72

acid/base transport by CAII requires both direct binding between transporter and enzyme as well as

73

CAII catalytic activity and has been coined tra sport

eta olo . First evidence for a transport 3

74

metabolon, formed between CAII and an acid/base transporter, has been presented in 1993 by Kifor

75

et al. (1993) for the Cl-/HCO3- exchanger AE1. CAII could be immunoprecipitated with AE1 when

76

antiserum against the N-terminal of AE1 was used, while serum directed against the C-terminal of

77

the transporter failed to immunoprecipitate CAII, suggesting that CAII physically binds to the C-

78

terminal tail of AE1 (Vince and Reithmeier, 1998). These data were confirmed by affinity blotting and

79

a solid phase binding assay with CAII and a GST fusion protein of the AE1 C-terminal (Vince and

80

Reithmeier, 1998). Single site mutations identified the acidic cluster D887ADD in the C-terminal tail of

81

AE1 as binding site for CAII (Vince and Reithmeier, 2000; Vince et al., 2000). Binding of CAII to this

82

cluster would tether the enzyme close to the transporter pore of AE1 near the inner cell surface. This

83

location has been suggested to ideally position CAII to hydrate incoming CO2 and directly supply the

84

AE1 transporter with a localized substrate pool (Vince and Reithmeier, 2000). Indeed, inhibition of

85

CAII catalytic activity decreased transport activity of AE1, heterologously expressed in HEK293 cells

86

by up to 60% (Sterling et al., 2001).

87

First evidence for a direct interaction between the Na+/HCO3- cotransporter NBCe1 and CAII was

88

presented by Gross et al. (2002) using isothermal titration calorimetry. In analogy to the CAII-binding

89

cluster D887ADD found in AE1 (Vince and Reithmeier, 2000), the authors suggested the cluster

90

D986NDD within the C-terminal of NBCe1 as the putative CAII binding site. Functional interaction

91

between NBCe1 and CAII was further shown by heterologous protein expression in Xenopus oocytes

92

(Becker and Deitmer, 2007). Both injection and co-expression of CAII increased NBCe1-mediated

93

membrane current, membrane conductance and Na+ influx during application of CO2/HCO3- in an

94

ethoxzolamide-sensitive manner.

95

Evidence for an interaction between NHE1 and intracellular CAII was obtained by measuring the

96

recovery from a CO2-induced acid load in AP1 cells, transfected with NHE1 (Li et al., 2002).

97

Cotransfection of NHE1 with CAII almost doubled the rate of pH recovery as compared to cells

98

expressing NHE1 alone, while cotransfection with the catalytically inactive mutant CAII-V143Y even

99

decreased the rate of pH recovery, indicating a physical interaction between NHE1 and catalytically 4

100

active CAII. Physical interaction between the two proteins was demonstrated by co-

101

immunoprecipitation of heterologously expressed NHE1 and CAII (Li et al., 2002). A micro titer plate

102

binding assay with a GST fusion protein of the NHE1 C-terminal tail revealed that CAII binds to the

103

penultimate group of 13 amino acids of the C-terminal tail (R790IQRCLSDPGPHP), with the amino acids

104

S796 and D797 playing an essential role in binding (Li et al., 2001, 2006).

105

While a considerable amount of data indicates a physical and functional interaction between various

106

acid/base transporters and carbonic anhydrases, several studies also questioned such transport

107

metabolons. Lu et al. (2006) could not observe a CAII-mediated increase in membrane conductance

108

in NBCe1-expressing Xenopus oocytes, even when fusing CAII to the C-terminal of NBCe1. In line with

109

these experiments, Yamada et al. (2011) could find no increase in the membrane current during

110

application of CO2/HCO3- when co-expressing wild-type NBCe1A or the mutant NBCe1-Δ

111

the putative CAII binding site D986NDD) with CAII. The concept of a physical interaction between

112

HCO3- transporters and CAII has also been challenged by a binding study of Piermarini and colleagues

113

(2007). The authors were able to reproduce the findings of other groups by showing that GST fusion

114

proteins of the C-terminal tails of NBCe1, AE1 and NDCBE (SLC4A8) can bind to immobilized CAII in a

115

micro titer plate binding assay. However, when reversing the assay or using pure peptides, no

116

increased binding of CAII to the immobilized GST fusion proteins could be detected (Piermarini et al.,

117

2007). It was concluded that a bicarbonate transport metabolon may exist, but that CAII might not

118

directly bind to the transporters. That CAII activity could improve substrate supply to bicarbonate

119

transporters even without the requirement for a metabolon, involving direct physical interaction,

120

was also pointed out in a study on AE1 transport activity by Al-Samir et al. (2013). By using Förster

121

resonance energy transfer measurements and immunoprecipitation experiments with tagged

122

proteins, the authors showed no binding or close co-localization of AE1 and CAII. Functional

123

measurements in red blood cells and theoretical modeling suggested that transport activity of AE1

124

can be best supported by CAII, when the enzyme is equally distributed within the cell s cytosol (Al-

p la ki g

5

125

Samir et al., 2013). For detailed reviews on transport metabolons see McMurtrie et al., 2004; Moraes

126

& Reithmeier, 2012; Deitmer & Becker, 2013; Becker et al., 2014.

127

We have previously shown that transport activity of MCT1 and MCT4 is enhanced by CAII, when the

128

two proteins are heterologously co-expressed in Xenopus oocytes (Becker et al., 2005, 2010, 2011;

129

Becker & Deitmer 2008). In contrast to the transport metabolons described before, enhancement of

130

MCT / tra sport fu tio is i depe de t of the e z

131

catalytic activity with 6-Ethoxy-2-benzothiazolesulfonamide (EZA) and co-expression of MCT1/4 with

132

the catalytically inactive mutant CAII-V143Y failed to suppress the CAII-induced facilitation of MCT1/4

133

transport activity (Becker et al., 2005; Becker & Deitmer, 2008). First evidence that this non-catalytic

134

interaction between MCT1/4 and CAII requires direct binding between the two proteins was shown

135

by injection of CAII that was bound to an antibody prior to the injection. In this experiment CAII was

136

not able to enhance transport activity of MCT1 in Xenopus oocytes, suggesting a sterical suppression

137

of the interaction by the antibody (Becker & Deitmer, 2008). In the same study, truncation of the

138

MCT1 C-terminal tail led to loss of interaction between MCT1 and CAII in Xenopus oocytes (Becker &

139

Deitmer, 2008). By introduction of single site mutations in MCT1 and MCT4, the glutamic acidic

140

clusters E489EE and E431EE within the C-terminal tail of MCT1 and MCT4, respectively, could be

141

identified as binding sites for CAII (Stridh et al., 2012; Noor et al., 2015). In both studies binding was

142

confirmed both on the functional level by measuring MCT transport activity in Xenopus oocytes and

143

by pull-down assays using GST-fusion proteins. Direct interaction between MCT1 and CAII was not

144

only shown in vitro and by heterologous protein expression in Xenopus oocytes, but also in mouse

145

astrocytes via an in situ proximity ligation assay (Stridh et al., 2012). Since CA catalytic activity is not

146

required to facilitate MCT transport function, it was hypothesized that CAII might utilize parts of its

147

intramolecular proton pathway to function as a proton antenna for the transporter (Almquist et al.,

148

2010; Becker et al., 2011). Protonatable residues with overlapping Coulomb cages could form proton-

149

attractive domains and could share a proton at a very fast rate, exceeding the upper limit of

150

diffusion-controlled reactions (Ädelroth et al., 2004; Friedman et al., 2005). When these residues are

e s atal ti a ti it . Both, inhibition of CA

6

151

located in proteins or lipid head groups at the plasma membrane, they can collect protons from the

152

solution and direct them to the entrance of a proton-transfer pathway of a membrane-anchored

153

protei , a phe o e o ter ed proto - olle ti g a te

154

2006). The need for such a proton antenna is based on the observation that H+ cotransporters, such

155

as MCTs, extract H+ from the surrounding area at rates well above the capacity for simple diffusion to

156

replenish their immediate vicinity. Therefore, the transporter must exchange H+ with protonatable

157

sites at the plasma membrane, which could function as proton collectors for the transporter

158

(Martínez et al., 2010). The finding that the mutant CAII-H64A, missing the central residue of the

159

e z

160

that CAII may function as proton antenna for the MCT, to provide or subtract protons to or from the

161

transporter, respectively, via its intramolecular H+ shuttle (Becker et al., 2011). In line with this, we

162

could recently show that integration of a proton antenna, in form of 6 histidine residues, into the C-

163

terminal tail of MCT4 could facilitate MCT4 transport activity even in the absence of carbonic

164

anhydrase (Noor et al., 2017).

165

Transport activity of MCTs was not only shown to be facilitated by intracellular CAII, but also by the

166

extracellular CA isoforms CAIV and CAIX (Klier et al., 2011, 2014; Jamali et al., 2015). Hypoxia-

167

regulated CAIX is considered to be a key protein in tumor acid/base regulation. CAIX, the expression

168

of which is usually linked to poor prognosis, is tethered to the membrane via a transmembrane

169

domain, with the catalytic center facing the extracellular site. Like other fast carbonic anhydrases,

170

CAIX is equipped with an intramolecular proton shuttle for rapid exchange of H+ between the

171

catalytic center and the surrounding bulk solution. We could demonstrate that knockdown of CAIX

172

with siRNA reduced MCT1-mediated lactate transport in hypoxic MCF-7 cells, while inhibition of CA

173

catalytic activity with EZA had no effect on MCT1 transport activity (Jamali et al., 2015). Co-

174

expression of MCT1/4 with CAIX in Xenopus oocytes revealed that CAIX-mediated facilitation of MCT

175

transport activity requires the enzymes intramolecular proton shuttle His200 (Jamali et al., 2015).

176

Furthermore, knockdown of CAIX, but not inhibition of CA catalytic activity, reduced proliferation of

a Ädelroth et al.,

; Brä dé et al.,

e s i tra ole ular proto shuttle, fails to fa ilitate MCT tra sport a ti it , led to the otio

7

177

hypoxic MCF-7 cells, indicating that the CAIX-driven increase in lactate efflux is crucial for proper

178

function of cancer cells (Jamali et al., 2015).

179

In the present study we investigated whether CAII can facilitate proton-driven lactate transport in

180

MCF-7 breast cancer cells. Knockdown of CAII, which is closely co-localized to MCT1 in MCF-7 cells,

181

resulted in a significant reduction of MCT transport activity in both normoxic and hypoxic cells. We

182

then further analyzed the molecular mechanism underlying the CAII-mediated facilitation of MCT

183

transport activity. Our results showed that CAII features a moiety (Glu69 + Asp72) that exclusively

184

mediates proton exchange with the transporter, while the central residue of its intramolecular

185

proton shuttle (His64) mediates the binding of CAII to MCT1 and MCT4.

186 187 188

Results

189

Intracellular CAII facilitates lactate transport in MCF-7 breast cancer cells

190

We have recently shown that extracellular CAIX facilitates lactate transport in hypoxic cancer cells by

191

acting as a proton antenna for MCT1/4 at the extracellular face of the plasma membrane (Jamali et

192

al., 2015). Since proton-driven lactate transport would also benefit from a proton antenna at the

193

cytosolic face of the cell membrane, we now investigated whether proton/lactate cotransport in

194

cancer cells is facilitated by intracellular CAII. Therefore, we knocked down CAII in normoxic (20% O2)

195

and hypoxic (1% O2) MCF-7 breast cancer cells using siRNA and determined MCT-mediated

196

proton/lactate cotransport by measuring changes in intracellular pH (pHi) during application and

197

removal of 3 and 10 mM lactate in the nominal absence of CO2/HCO3- (Figure 1A, B). Indeed,

198

knockdown of CAII resulted in a significant decrease in the rate of change in pHi ΔpHi/Δt

199

application and removal of lactate, indicating a decrease in the lactate transport rate with reduced

200

CAII (Figure 1C, D). Interestingly, reduction in transport activity was more pronounced in hypoxic cells

201

than under normoxic conditions (25%-33% reduced in normoxic cells, 47%-57% reduced in hypoxic

oth duri g

8

202

cells), even though expression of CAII, in contrast to CAIX, is not upregulated under hypoxia (Figure 1

203

- figure supplement 1; Jamali et al., 2015). A possible reason for this effect might be that the increase

204

in proton/lactate transport, mediated by extracellular CAIX under hypoxia, challenges the need for a

205

faster proton supply at the intracellular site. Knockdown efficiency of CAII was checked by qRT-PCR

206

and western blot analysis, showing a 60% reduction in normoxic and hypoxic cells (Figure 1 - figure

207

supplement 1). To investigate whether the CAII-induced augmentation in MCT1 transport activity

208

requires catalytic activity of the enzyme, lactate transport in MCF-7 cells can be determined in the

209

presence of the CA inhibitor EZA. We had shown this experiment in a recent study on MCF-7 cells

210

with the same settings as used in the present study (Jamali et al., 2015). In that study application of

211

30 µM EZA had no effect on lactate transport, as measured with the lactate-sensitive FRET

212

nanosensor Laconic (San Martín et al., 2013) and by determining the rate of change in intracellular

213

pH, in normoxic and hypoxic MCF-7 cells, respectively, neither in the presence nor in the absence of

214

5% CO2/15 mM HCO3- (see Figures 3 h, i, and S2 in Jamali et al., 2015). Since EZA did not alter lactate

215

transport in MCF-7 cells under any condition, it can be concluded that the observed reduction in

216

lactate transport by knockdown of CAII is independent of the e z

217

Knockdown of CAII decreased the basal pHi, as measured at the beginning of the experiment, by

218

approximately 0.1 pH units, both in normoxic and in hypoxic MCF-7 cells (Figure 1 - figure

219

supplement 2 A). This acidification indicates that CAII plays an important role in the acid/base

220

regulation of MCF-7 cells, independent the oxygen tension, either directly by providing HCO3- to the

221

buffer system, or indirectly by interacting with the cell s acid/base transporters. Since an intracellular

222

acidification leads to a less favorable gradient for H+-coupled lactate influx, the observed reduction in

223

ΔpHi/Δt duri g la tate appli atio

224

investigate the influence of pHi on la tate tra sport, e plotted ΔpHi/Δt duri g appli atio of

225

lactate (which was always carried out as the first pulse) against the initial pHi (before lactate

226

application) for every individual cell (Figure 1 - figure supplement 2 B-E). Under none of the four

227

o ditio s a positi e orrelatio

e s catalytic activity.

ould also e the result of this ha ge i the H + gradient. To M

et ee ΔpHi/Δt a d pHi could be observed. From these results it 9

228

can be concluded that the decrease in pHi, as induced by knockdown of CAII, seems to play only a

229

minor role in the observed reduction in lactate transport.

230 231

CAII is co-localized with MCT1 in cancer cells

232

We have recently investigated which MCT isoforms mediate lactate transport in MCF-7 cells under

233

normoxic and hypoxic conditions (Jamali et al., 2015). In that study application of 300 nM AR-

234

C155858 fully inhibited the lactate-induced acidification in normoxic and of hypoxic MCF-7 cells. AR-

235

C155858 has been shown to inhibit transport activity of MCT1 (and MCT2 when expressed without

236

its ancillary protein embigin in Xenopus oocytes), but has no effect on MCT4 transport activity (Ovens

237

et al., 2010a,b). By measuring the rate of change in pHi during application of different lactate

238

concentrations we determined Km values of ~5 mM lactate form both normoxic and hypoxic MCF-7

239

cells (Jamali et al., 2015). For MCT1, a Km value between 3–8 mM had been determined in various cell

240

types (Carpenter & Halestrap, 1994; Bröer et al., 1997, 1998). For MCT2, the Km value was found to

241

be 0.74 (Bröer et al., 1999; Heidtmann et al., 2015), while for MCT4 Km values between 17–35 mM

242

have been reported (Dimmer et al., 2000). Western blot analysis for MCT1, MCT2 and MCT4 revealed

243

sharp bands only for MCT1 in MCF-7 cells both under normoxic and hypoxic conditions (Jamali et al.,

244

2015). Taken together, these results indicate that lactate transport in MCF-7 cells is exclusively

245

mediated by MCT1, both under normoxic and hypoxic conditions.

246

Antibody staining of MCT1 and CAII revealed a homogenous distribution of both proteins within the

247

cells (Figure 2A). To investigate whether CAII is colocalized with MCT1 in MCF-7 cells, we performed

248

an in situ proximity ligation assay (PLA). The PLA indicated close proximity (< 40 nm) between MCT1

249

and CAII, both in normoxic and hypoxic cells (Figure 2B1, B2). The amount of PLA signals per nucleus

250

did not significantly differ between normoxic and hypoxic cells (Figure 2C). Knockdown of CAII

251

resulted in a significant reduction in the signal (Figure 2B3, C). When the PLA was carried out without

252

primary antibodies as negative control, no PLA signals could be detected (Figure 2B4, C).

10

253

Taken together, these data suggest that CAII directly interacts with MCT1 to facilitate transport

254

activity.

255 256

CAII supports proliferation of cancer cells

257

To investigate whether the CAII-mediated augmentation in lactate flux facilitates cancer cell

258

proliferation, we determined the number of MCF-7 cells kept under different conditions for up to

259

three days (Figure 3A-D). Indeed, knockdown of CAII significantly decreased cell proliferation (as

260

compared to cells transfected with non-targeting negative control siRNA) both under normoxia and

261

hypoxia, while transfection with non-targeting negative control siRNA had no significant effects on

262

cell proliferation. Interestingly, total inhibition of lactate transport with 300 nM AR-C155858

263

decreased cell proliferation by the same degree as did knockdown of CAII (as compared to untreated

264

cells). Inhibition of CAII catalytic activity with 30 µM EZA, however, had no effect on cell proliferation

265

(as compared to untreated cells). The effects of AR-C155858 and EZA are in line with previous

266

findings that inhibition of MCT transport activity significantly reduces proliferation of MCF-7 and

267

MDA-MB-231 cells, while inhibition of CA catalytic activity has no effect on cell proliferation (Jamali

268

et al., 2015). The striking effect of CAII knockdown on normoxic and hypoxic cells suggests the

269

possibility that the CAII protein might have yet another role in cell proliferation, in addition to its

270

function as a facilitator of lactate efflux and its catalytic function in cellular pH regulation.

271 272

CAII facilitates a proton-collecting apparatus at its surface to drive MCT transport activity

273

We have previously shown that CAII facilitates transport activity of MCT1 and MCT4 when

274

heterologously expressed in Xenopus oocytes. CAII-mediated facilitation of MCT1/4 transport activity

275

was found to be independent of the e z

276

activity with EZA and co-expression of MCT1/4 with the catalytically inactive mutant CAII-V143Y

277

failed to suppress the CAII-induced facilitation of MCT1/4 transport activity (Becker et al., 2005;

e s atal ti a ti it , si e both inhibition of CA catalytic

11

278

Becker & Deitmer, 2008). Since CA catalytic activity is not required to facilitate MCT transport

279

function, it was hypothesized that CAII might utilize parts of its intramolecular proton pathway to

280

function as a proton antenna for the transporter (Almquist et al., 2010; Becker et al., 2011). To

281

investigate which moieties within the CAII protein could mediate this antennal function, we tested

282

the influence of various CAII mutants on MCT1/4 transport activity using the Xenopus oocyte

283

expression system.

284

By using an algorithm for mapping proton wires in proteins, Shinobu and Agmon (2009) suggested

285

that CAII-Glu69 and CAII-Asp72, located on an electronegative patch on the rim of the active site

286

cavity of CAII, could fu tio as a do ki g statio

287

to the catalytic center. The position of the two amino acids is shown in the cartoon in Figure 4 A. To

288

test whether CAII-Glu69 or CAII-Asp72 are involved in the facilitation of MCT transport activity, we

289

mutated these two amino acids, as well as the nearby Asp71, and co-expressed the resulting mutants

290

with either MCT1 or MCT4 in Xenopus oocytes. Transport activity of MCT1/4 was determined by

291

measuring changes in intracellular proton concentration ([H+]i) during application and removal of 3

292

and 10 mM lactate, while CAII catalytic activity was checked by application of 5% CO2 / 10 mM HCO3-

293

(Figure 4B; Figure 4 – figure supplement 1A). CAII-WT increased transport activity of MCT1 by 80-

294

100%, as measured by the rate of change in [H+]i Δ[H+]i/Δt duri g application and removal of lactate

295

(Figure 4C, D). However, no significant increase in transport activity was observed, when MCT1 was

296

co-expressed with CAII-E69Q or CAII-D72N, while mutation of the nearby Asp71 (CAII-D71N) had no

297

effect on MCT1 transport activity (Figure 4B-D). Co-expression of the CAII mutants with MCT4 gave

298

similar results as those obtained with MCT1 (Figure 4 – figure supplement 1A-C). Western blot

299

analysis showed that protein expression levels were not altered by the mutations in CAII (Figure 4 –

300

figure supplement 2).

301

It has been shown previously that addition of exogenous H+ donors/acceptors, like imidazole or

302

carnosine, can rescue proton shuttling in carbonic anhydrase in vitro (An et al., 2002; Tu et al., 1989).

303

To check whether chemical rescue of the H+ shuttle in CAII-E69Q and CAII-D72N also can restore the

for protonated buffer molecules to deliver protons

12

304

functional interaction between these mutants and MCT1, we injected 4-methylimidazole (4-MI) into

305

oocytes co-expressing MCT1 and CAII-E69Q/CAII-D72N. Oocytes expressing MCT1 alone or co-

306

expressing MCT1 + CAII-WT or MCT1 + CAII-D71N, respectively, were used as control. The effective

307

free volume of a oo te is arou d .

308

400 mM solution of 4-MI should result in an intracellular 4-MI concentration of 30 mM.

309

Crystallization studies have shown that 4-MI i ds to CAII ear the His

310

a moiety near Glu69 and Asp72 (Figure 5A; Duda et al., 2001, 2003). After injection of 4-MI, MCT1

311

transport activity was enhanced after co-expressing CAII-E69Q or CAII-D72N to a similar extent as

312

with co-expression with CAII-WT and CAII-D71N (Figure 5B-D). Since 4-MI acts as mobile H+ buffer,

313

the compound is expected to increase the cytosolic buffer strength (βi), which in turn alters changes

314

in [H+]i. Therefore, we calculated net H+ fluxes (JH) from the rate of change in intracellular pH during

315

application of 3 and 10 mM lactate and βi (Figure 5E; Figure 5 – figure supplement 1A). Injection of 4-

316

MI induced a significant increase in lactate-induced JH in oocytes co-expressing MCT1 with CAII-E69Q

317

and D72N, respectively, while leaving JH in oocytes co-expressing MCT1 with CAII-WT or CAII-D71N,

318

as well as oocytes expressing MCT1 alone, unaltered (Figure 5E). Again the same results could be

319

observed with MCT4 (Figure 5 – figure supplements 1 and 2). These data indicate that the exogenous

320

H+ donor/acceptor 4-MI can rescue the facilitation of MCT1/4 transport activity by CAII-E69Q and

321

CAII-D72N.

μl )euthen et al., 2002). Therefore, injection of 27.6 nl of a

π-stacking to Trp5 and to

322 323

CAII-Glu69 and Asp72 are not involved in binding between CAII and MCTs

324

We have previously shown that CAII-mediated enhancement of MCT1/4 transport activity requires

325

direct binding of the enzyme to an acidic cluster within the transporters C-terminal tails (Stridh et

326

at., 2012; Noor et al., 2015). To check whether the failure of CAII-E69Q and CAII-D72N to enhance

327

MCT transport activity is due to loss of binding between transporter and enzyme, we performed a

328

pull-down assay with GST fusion proteins of the C-terminal tail of MCT1 and MCT4, respectively (GST-

329

MCT1-CT, GST-MCT4-CT), and lysates from oocytes expressing CAII-WT, CAII-E69Q, CAII-D71N, and 13

330

CAII-D72N, respectively (Figure 6). All three CA mutants could be pulled down with GST-MCT1-CT and

331

GST-MCT4-CT, with no evident changes in signal intensity between the mutants and CAII-WT. As

332

negative control, CAII-WT was pulled down with GST alone, resulting in no or only very weak signals.

333

These results indicate that the failure of CAII-E69Q and CAII-D72N to enhance MCT transport activity

334

is not due to a loss of direct binding between transporter and enzyme.

335 336

CAII-Glu69 and Asp72 do not support CAII catalytic activity

337

By using an algorithm for mapping proton wires in proteins, Shinobu and Agmon (2009) had shown

338

that Glu69 and Asp72 form an extension of the CAII proton wire. From this model, the authors

339

proposed that positively charged, protonated buffer molecules could dock in that area, from which a

340

proton is delivered to the active site to facilitate catalytic activity. To investigate whether Glu69 and

341

Asp72 play a role in CAII catalytic function, we determined CAII catalytic activity in oocyte lysates by

342

gas analysis mass spectrometry (Figure 7A, B). CA atal ti a ti it is deter i ed

343

e ri h e t LE , which is influenced both by the rate of hydration and dehydration of CO2 (see

344

Materials & Methods). Therefore a change in LE indicates changes of CA catalytic activity both in the

345

hydration and dehydration reaction. Mutation of CAII-Glu69, Asp71, or Asp72 did not alter CAII

346

catalytic activity, as compared to CAII-WT, while mutation of CAII-His64 (the central residue of the

347

CAII intramolecular H+ shuttle) resulted in a 78% reduction in CA activity (Figure 7A, B). Even when

348

both Glu69 and Asp72 were mutated together (CAII-E69Q/D72N), no alterations in CAII catalytic

349

activity could be observed (Figure 7A, B). The results from the mass spectrometry were confirmed my

350

easuri g Δ[H+]i/Δt i

the log

CAII-expressing oocytes during application and removal of CO2/HCO3-.

351

E pressio of CAII resulted i a sig ifi a t i rease i Δ[H+]i/Δt, as compared to native oocytes (Figure

352

7C, D). Mutation of CAII-Glu69, Asp71, Asp72, or double mutation of Glu69/Asp72 resulted in no

353

significant change in CA activity, neither during application nor removal of CO2/HCO3-. Only mutation

354

of CAII-His64 induced a significant decrease in catalytic activity. These data indicate that neither

355

Glu69 nor Asp72 are required for full catalytic activity in CAII. Interestingly, mutation of His64 to Ala 14

356

reduced CAII catalytic activity to 25%, when measured by gas analysis mass spectrometry (Figure 7

357

B), but only to 50%, when measured by Δ[H+]i/Δt i i ta t oo tes Figure C, D . These fi di gs i fer

358

that the cytosol of oocytes already contains certain, still unidentified, buffer compounds that can

359

support catalytic activity of the mutant.

360 361

CAII-His64 mediates binding of CAII to MCT1/4, but is not involved in H+ shuttling between the

362

proteins

363

Our previous studies had shown that CAII-mediated facilitation of MCT transport activity requires

364

CAII-His64, the central residue of the enzymes intramolecular H+ shuttle (Becker et al., 2011).

365

Furthermore, our structural models suggest that His64 mediates binding between CAII and an acidic

366

cluster within the C-terminal tails of MCT1 and MCT4 (Figure 8A; Figure 8 – figure supplement 1A;

367

Stridh et al., 2012; Noor et al., 2015). To investigate whether CAII-His64 mediates binding and/or

368

proton shuttling between MCT1/4 and CAII, we co-expressed MCT1/4 with a CAII in which the

369

histidine at position 64 was either mutated to alanine (CAII-H64A) or to lysine (CAII-H64K). Mutation

370

of His to Ala should disrupt both intramolecular H+ shuttling in CAII and binding of the enzyme to the

371

glutamate residues in MCT1/4. Mutation of His to Lys does also disable H + shuttling, due to the

372

increase in pKa, but should still allow binding of CAII to the glutamate residues in the C-terminal tail

373

of MCT1/4. Indeed, tra sport a ti it

374

application and removal of lactate, was not enhanced by co-expression with CAII-H64A, while CAII-

375

H64K enhanced MCT transport activity to the same extent as did CAII-WT (Figure 8B-D; Figure 8 –

376

figure supplement 1B-D).

377

Binding of the CAII mutants to MCT1/4 was evaluated by a pull-down assay with GST fusion proteins

378

of the C-terminal tail of MCT1 and MCT4, respectively (GST-MCT1-CT, GST-MCT4-CT) and lysates

379

from oocytes expressing CAII-WT, CAII-H64A, and CAII-H64K, respectively (Figure 8E, F; Figure 8 –

380

figure supplement 1E, F). CAII-H64A was not able to bind to the C-terminal tail of MCT1 and MCT4,

of MCT

a d MCT , as

easured

Δ[H+]i/Δt duri g

15

381

respectively, showing only 12 and 14% signal intensity of the band for CAII-WT, and therefore not

382

being significantly different from the negative control, in which CAII-WT was pulled down with GST

383

alone. In contrast, CAII-H64K did still bind to the C-terminal tail of MCT1 and MCT4, with no

384

significant alterations in signal intensity as compared to the band for CAII-WT.

385

Catalytic activity of CAII-H64A and CAII-H64K was determined in oocyte lysates by gas analysis mass

386

spectrometry (Figure 8 – figure supplement 2A, B). Both mutations of His64 to Ala and Lys resulted in

387

a significant decrease in CA catalytic activity with a reduction to 53% (CAII-H64K) and 23% (CAII-

388

H64A), respectively, as compared to CAII-WT. However, mutation of His64 did not change the

389

expression level of CAII in oocytes (Figure 8 – figure supplement 2C, D). The finding that catalytic

390

activity of CAII-H64K is strongly reduced, compared to CAII-WT, but still significantly higher than the

391

activity of CAII-H64A indicates that the lysine at position 64 might still be able to function as

392

intramolecular H+ shuttle, even though with significantly less efficacy than histidine.

393

Taken together, these results indicate that His64, the central residue of the CAII intramolecular H+

394

shuttle, mediates i di g of CAII to the tra sporter s C-terminal tail rather than H+-transfer between

395

the two proteins. However, since the H64K mutation still seems to retain some shuttling activity, a

396

role of His64 in H+-transfer cannot be fully ruled out.

397 398

The histidines in the N-terminus of CAII are not involved in binding of CAII to MCT1

399

It was previously proposed by Vince et al. (2000) that CAII binds to the C-terminal tail of the Cl-/HCO3-

400

anion exchanger AE1 via a cluster of five histidine and one lysine residues in the enzyme s N-terminal

401

domain.

402

H3P/H4Q/K9A/H10K/H15K/H17S;

403

(which does not bind to AE1) failed to bind to the transporter. Indeed the same CAII-mutant was also

404

not able to facilitate MCT1 transport activity (Becker et al., 2008). To investigate the role of His3,

405

His4, Lys9, His10, and His15 in the facilitation of MCT transport activity, we mutated every single one

This

assumption

was

based

oi ed

on

the

finding

that

a

CAII-mutant

(CAII-

CAII-HEX ), mimicking the N-terminal domain of CAI

16

406

of these amino acids, either alone or in combination, and co-expressed the resulting mutants with

407

MCT1 in Xenopus oocytes. However, none of the mutations resulted in a loss of functional interaction

408

between MCT1 and CAII (Figure 9), indicating that none of the histidines or lysine in the N-terminal of

409

CAII is directly involved in the facilitation of MCT transport activity. A possible reason for the inability

410

of CAII-HEX to facilitate MCT transport activity might be that the introduction of charged and bulky

411

amino acids into the N-terminal domain may prohibit access of the MCT C-terminal tail to the binding

412

site CAII-His64.

413 414 415

Discussion

416

Our previous studies have shown that CAII facilitates transport activity of MCT1 and MCT4

417

independent of the enzyme s catalytic activity, since both inhibition of CA activity with EZA and

418

molecular disruption of the catalytic center (CAII-V143Y) did not reduce the CAII-induced increase in

419

MCT transport activity (Becker et al., 2005; 2008). Since CA catalytic activity is not required to

420

facilitate MCT transport function, we hypothesized that CAII might utilize parts of its intramolecular

421

proton pathway to function as a proton antenna for the transporter (Almquist et al., 2010; Becker et

422

al., 2011). The proton pathway of CAII consists of a water wire, coordinated by Asn62 and Asn67,

423

which extends from the active site to the histidine residue at position 64 (Fisher et al., 2007a,b).

424

From there the proton is passed on between the imidazole ring in His64 and exogenous buffer

425

molecules surrounding the enzyme (Fisher et al., 2007a,b). In addition, Shinobu and Agmon (2009)

426

presented a model in which the active site H+ wire exits to the protein surface, and leads to Glu69

427

and Asp72, located on an electronegative patch on the rim of the active-site cavity. Based on this

428

model, the authors proposed that positively charged, protonated buffer molecules could dock in that

429

area, from which a proton is delivered to the active site, when the enzyme works in the dehydration

430

direction. However, this assumption is purely based on mathematical modelling and has, to our 17

431

knowledge, not yet been evaluated experimentally. In the present study, we could show that both

432

Glu69 and Asp72 are essential for CAII-mediated facilitation of MCT1/4 transport activity both in the

433

influx and efflux direction. Mutation of Glu69 and Asp72 did, however, not alter CAII catalytic

434

activity, as demonstrated in vitro by gas analysis mass spectrometry and in vivo by determining the

435

rate of change in intracellular H+ concentration in oocytes during application and removal of CO2.

436

These results suggest that this proton-collecting moiety is not involved in the facilitation of CAII

437

catalytic activity, but rather mediates a second, independent function, which is the rapid supply and

438

removal of protons from the pore of the adjacent transporter, rendering CAII a bona fide proton

439

antenna. Whether this kind of interaction does only exist between CAII and MCTs, or whether CAII

440

could also function as proton antenna for other proton-coupled membrane transporters, like proton-

441

coupled amino acid transporters or Na+/H+ exchangers, remains to be shown.

442

Our previous studies had shown that CAII-mediated facilitation of MCT transport activity requires

443

CAII-His64, which mediates the exchange of protons between catalytic center and surrounding bulk

444

solution during the CAII catalytic cycle. However, our structural models suggested that His64 rather

445

mediates binding of CAII to Glu489/Glu491 in MCT1 (Stridh et al., 2012) and Glu431/Glu433 in MCT4

446

(Noor et al., 2015). To investigate whether His64 mediates binding or proton shuttling between MCT

447

and CAII, or both, we exchanged the histidine at position 64 either to alanine or to lysine. Mutation

448

of His to Ala should disable both intramolecular H+ shuttling in CAII and binding of the enzyme to the

449

glutamate residues in MCT1/4. Mutation of His to Lys does also disable H+ shuttling, since lysine has a

450

considerably higher pKa than the imidazole ring in histidine, which does not allow fast

451

protonation/deprotonation reactions at physiological pH. However, lysine should still be able to form

452

hydrogen bonds with the glutamate residues in the C-terminal tail of MCT1 and MCT4, and should

453

therefore still enable CAII to bind to the transporter. While MCT transport activity was not enhanced

454

by CAII-H64A, CAII-H64K facilitated MCT transport activity to the same extent as CAII-WT. These

455

results let us conclude that His64 may not be involved in proton transfer between MCT and CAII, but

456

instead mediates binding of the enzyme to the transporters C-terminal tail. However, since CAII18

457

H64K still displayed a higher catalytic activity than CAII-H64A, it cannot be ruled out that CAII-H64K

458

still retains some proton shuttling activity. Indeed, it has been shown that lysine residues that are

459

buried in the protein interior could display considerably lower pKa values than in water (Isom et al.,

460

2011). Therefore, it might be possible that CAII-H64K can still retain enough shuttling activity to

461

facilitate MCT-mediated lactate transport.

462

Taken together, these results suggest that CAII facilitates MCT transport activity by a completely

463

different mechanism than the catalytic activity: Catalytic activity of CAII is facilitated by proton

464

shuttling via His64, while Glu69 and Asp72 are ineffective in supporting CA activity. When interacting

465

with MCT1/4, His64 does function as binding site, while proton transfer between transporter and

466

enzyme seems to be mediated by Glu69 and Asp72 (Figure 10).

467

The physiological need for CAII serving as such a proton antenna for MCTs might derive from the low

468

apparent diffusion rate of protons in a strongly buffered solution like the cytosol. By applying a

469

mathematical model of Brownian diffusion, Martinez et al. (2010) determined the maximum supply

470

capacity of substrates to various transporters and enzymes. For the MCT1 the authors calculated that

471

the diffusion rate of lactate, which is present in the cell within the millimolar range, exceeds the

472

turnover rate of the transporter by several magnitudes, supporting the notion that for this molecule

473

the cytosol is a well-mixed compartment. For protons, however, the authors calculated that the

474

maximum supply rate is approximately 15 times lower than the apparent turnover number of MCT1,

475

as calculated by Ovens et al. (2010a). In other words, the model demonstrates that MCT1 extracts H+

476

from the cytosol at rates well above the capacity for simple diffusion to replenish its immediate

477

vicinity. This paradoxical result implies that the transporter does not extract H+ substrates directly

478

from the bulk cytosol, but fro

479

aqueous phase and the transport site (Martinez et al., 2010). The findings of the present study

480

suggest that CAII could serve as such a harvester which captures protons from protonatable residues

481

at the plasma membrane and funnels them to the transporter pore (Figure 10). By this, CAII would

a i ter ediate har esti g

o part e t located between the

19

482

counteract the local depletion of protons in the direct vicinity of the transporter to support transport

483

activity. Since the direction of proton movement is depending on the ion gradient, this principle can

484

be applied both in the influx and in the efflux direction. In case of proton/lactate efflux, CAII would

485

harvest H+ from protonatable residues near the plasma membrane and shuttle them to the

486

transporter, in case of proton/lactate influx, CAII would capture the H+ from the transporter pore

487

and distribute them to protonatable residues near the plasma membrane, following the

488

electrochemical H+ gradient. Indeed, we could previously show that protons, which are focally

489

applied to MCT1-expressing oocytes, moved faster along inner face of the plasma membrane when

490

CAII was injected, suggesting that CAII could allocate H+ along the plasma membrane and thereby

491

increase the area in which H+ can be released from the membrane into the bulk solution (Becker &

492

Deitmer, 2008). By this, CAII would counteract the accumulation of protons at the transporter pore,

493

which would result in reduction of transport activity in the influx direction.

494

Even though the mathematical model from Martinez et al. (2010) suggests a 15 fold lower supply

495

rate of H+, our measurements in Xenopus oocytes have shown that CAII enhances transport activity

496

of MCT a d MCT

497

activity is either facilitated by other proton-collecting molecules, or - more likely - that MCTs are

498

already able to extract protons from adjacent protonatable sites, even though not as efficient as

499

when cooperating with CAII.

500

The relevance of this proton antenna for proper cell function becomes evident from our physiological

501

experiments on MCF-7 breast cancer cells. First evidence for a functional interaction between MCT

502

and CAII comes from the finding that knockdown of CAII reduces lactate transport in MCF-7 cells.

503

That this decrease in transport activity is due to a loss of CAII catalytic activity could be ruled out,

504

since we could show in a recent study that inhibition of CA activity by application of EZA did not alter

505

lactate flux in MCF-7 cells under the same conditions (Jamali et al., 2015). Therefore, it can be

506

assumed that CAII facilitates lactate flux in cancer cells by a non-catalytic, direct interaction. Indeed,

o l

arou d t o fold. This i plies that in the absence of CAII MCT transport

20

507

CAII is in close co-localization with MCT1 in MCF-7 cells, as shown by in situ proximity ligation assay.

508

Thus it appears likely that CAII can form a protein complex with MCT1 in the cells to facilitate lactate

509

flux by functioning as a proton antenna for the transporter. In line with these findings, we previously

510

showed that CAII also facilitates lactate transport in mouse cerebellar and white matter astrocytes,

511

i depe de t fro

512

and CAII are in close proximity in astrocytes as shown by in situ PLA. Interestingly, colocalization

513

could only be detected when an antibody directed against the intracellular loop between TM 6 and 7

514

of MCT1 was used. When the PLA was carried out using an antibody against the C-terminal of MCT1,

515

carrying the CAII binding site, no proximity between MCT1 and CAII could be detected (Stridh et al.,

516

2012). Astrocytes are highly glycolytic cells, which are widely believed to export lactate as an energy

517

fuel for adja e t euro s, a phe o e o

518

Magistretti, 1994; Magistretti, 2006; Barros and Deitmer, 2010). These findings infer that CA-

519

mediated facilitation of MCT1/4 transport activity might be a more general, or at least common,

520

phenomenon in glycolytic cells which have to extrude high amounts of lactate.

521

In the present study, knockdown of CAII resulted in a more pronounced reduction in MCT1 transport

522

activity in hypoxic than in normoxic cancer cells, even so expression of CAII was not increased under

523

hypoxia. This effect might be attributed to extracellular CAIX, which functions as a proton antenna

524

for MCT1 in MCF-7 cells under hypoxic conditions (Jamali et al., 2015). CAIX operates only on the

525

extracellular side of the membrane, which might lead to the formation of a proton micro-domain on

526

the cytosolic side of the membrane in the absence of intracellular CAII, since supply or removal of H+

527

is enhanced on the extracellular side, leading to an increase in MCT transport activity. This would

528

result in increased accumulation or depletion of H+ at the cytosolic side, depending on transport

529

direction, which would impair MCT transport rate. With a proton antenna on both sides of the

530

membrane, formation of H+ micro-domains would be suppressed both at the cis- and at the trans-

531

side. In such a scenario, extracellular CAIX and intracellular CAII would cooperate

532

pri iple , providing protons to the transporter on one side and removing them on the other side of

the e z

e s atal ti a ti it

“tridh et al.,

. Like in the cancer cells, MCT1

oi ed Astrocyte-to-Neuron Lactate Shuttle Pelleri a d

a push a d pull

21

533

the cell membrane (Figure 10). By this mechanism, intracellular CAII and extracellular CAIX could

534

cooperate to enhance lactate transport in cancer cells under hypoxic conditions. Since CAIX is already

535

expressed in normoxic MCF-7 cells at a low level (Jamali et al., 2015), this push and pull principle

536

could also occur in normoxic cells, even though to a lower extend. The lower, yet robust, expression

537

of extracellular CAIX under normoxia could also explain the reduced effect of CAII knockdown on

538

lactate transport in normoxic MCF-7 cells. In line with this notion, we previously found in Xenopus

539

oocytes that intracellular CAII can also work in concert with another extracellular carbonic

540

anhydrase, CAIV, to ensure rapid shuttling of protons and lactate across the cell membrane (Klier et

541

al., 2014).

542

In CAII, H+ shuttling between enzyme and transporter is mediated by Glu69 and Asp72. CAIX seems to

543

lack an analogue moiety within its catalytic domain. However, it features a 59 amino acids long

544

proteoglycan-like (PG) domain that is unique to CAIX among the CA family (Pastorek & Pastorekova,

545

2015). The PG domain of human CAIX contains 18 glutamate and 8 aspartate residues, which have

546

been suggested to function as an intramolecular proton buffer, which could support CAIX catalytic

547

activity when operating in an acidic environment (Innocenty et al., 2009). We could recently show

548

that both truncation of the CAIX PG domain and application of a PG-binding antibody (M75)

549

decreased CAIX-induced facilitation in Xenopus oocytes and breast cancer cells (Ames et al., 2018).

550

Those findings let us to the conclusion that the CAIX PG domain could function as extracellular

551

proton antenna for monocarboxylate transporters within the acidic environment of a solid tumor.

552

The importance of CAII for cell viability was demonstrated by a cell proliferation assay. Knockdown of

553

CAII reduced proliferation of MCF-7 cells to the same extent as did full inhibition of lactate transport

554

with AR-C155858, while inhibition of CA catalytic activity with EZA had no effect on proliferation.

555

These results are correspond to previous findings that knockdown of CAIX or interference with the

556

CAIX PG domain by an antibody, but not inhibition of CA catalytic activity, decreases cell proliferation

557

in hypoxic MCF-7 and MDA-MB-231 cancer cells (Jamali et al., 2015; Ames et al., 2018). Apparently,

22

558

removal of the proton antenna on one side of the membrane is sufficient to reduce MCT-mediated

559

proton/lactate efflux from the cell. This in turn could lead to intracellular lactacidosis and impairment

560

of metabolism, which will ultimately result in a decrease in cell proliferation.

561

Knockdown of CAII decreased, but did not abolish lactate transport in MCF-7 cells. However,

562

knockdown of CAII decreased cell proliferation to the same extend as did inhibition of MCT1

563

transport activity with AR-C155858. This striking effect of CAII knockdown on normoxic and hypoxic

564

cells suggests that the CAII protein might have another role in cell proliferation, which is beyond its

565

function as a facilitator of lactate efflux. Since cell proliferation is not decreased by inhibition of CA

566

activity with EZA, the supporting effect of CAII seems to be independent of the enzyme s catalytic

567

activity. Zhou et al. (2015) showed that knockdown of CAII in NCI-H727 and A549 cancer cell lines

568

resulted in significant reduction in clonogenicity in vitro, and marked suppression of tumor growth in

569

vivo. By using an apoptosis gene array, the authors found a novel association of CAII-mediated

570

apoptosis with specific mitochondrial apoptosis–associated proteins. These findings suggest that the

571

CAII protein features multiple functions within the cell, including cellular acid/base regulation,

572

facilitation of lactate transport, and regulation of apoptosis.

573

In summary, our results show that CAII supports lactate flux in cancer cells, presumably by

574

functioning as a proton antenna for MCTs. Therefore, CAII features a molecular moiety that

575

exclusively mediates proton exchange with the transporter, while parts of its intramolecular proton

576

shuttle, pivotal for catalytic activity, mediate binding to the MCT. This enhancement of proton

577

movement does not only drive basic lactate flux, but is also a prerequisite for further facilitation of

578

lactate transport by CAIX under hypoxic conditions, as increased glycolytic activity requires even

579

higher lactate transport capacity in cancer cells.

580 581 582 23

583

Materials and Methods

584 585

Key resources table Reagent type (species) or resource

Designation

cell line (homo sapiens)

MCF-7

recombinant DNA reagent

pGHJ-CAII

recombinant DNA reagent

pGHJ-MCT1

recombinant DNA reagent

pGHJ-MCT4

recombinant DNA reagent

pGEX-MCT1-CT

recombinant DNA reagent

pGEX-MCT4-CT

antibody antibody antibody antibody antibody

antibody antibody antibody marker

rabbit anti-carbonic anhydrase II polyclonal antibody goat anti-MCT1 polyclonal antibody mouse anti-β-Actin monoclonal antibody mouse anti-GST Tag monoclonal antibody goat anti-rabbit IgG horseradish peroxidase-conjugated secondary antibody goat anti-mouse IgG horseradish peroxidase-conjugated secondary antibody Alexa Fluor 546 donkey anti-rabbit IgG Alexa Fluor 488 donkey anti-goat IgG Alexa Fluor 488 Phalloidin

Source or reference German Collection of Microorganisms and Cell Cultures (DSMZ) Becker et al., 2015 PMID: 16174776 Bröer et al., 1998 PMID: 9639576 Dimmer et al., 2000 PMID: 10926847 Stridh et al., 2012 PMID: 22451434 Noor et al., 2015 PMID: 25561737

Identifiers

Additional information

ACC-115

Millipore

AB1828

1:500

Santa Cruz Biotechnology

sc-14917

1:300

Santa Cruz Biotechnology

sc-47778

1:2500

Millipore

05-782

1:400

Santa Cruz Biotechnology

sc2004

1:2000

Santa Cruz Biotechnology

sc-2031

1:2000

Invitrogen

A10040

1:1000

Invitrogen

A-11055

1:1000

Life Technologies

A12379

1:500

586 587

Cultivation of MCF-7 cells

588

The human breast adenocarcinoma cell line MCF-7 was purchased from the German Collection of

589

Microorganisms and Cell Cultures DSMZ, Braunschweig, Germany (DSMZ-No. ACC-115). Cells were

590

cultured i

591

#353001, Fisher Scientific) in RPMI-1640 medium (Sigma-Aldrich, Schnelldorf, Germany),

592

supplemented with 10% fetal calf seru , % Mi i al Eagle s Mediu , .

593

glucose, 16 mM NaHCO3, and 1% penicillin/streptomycin, pH 7.2, at 37°C in 5% CO2, 95% air

594

(normoxia) or 5% CO2, 1% O2, 94% N2 (hypoxia) in humidified cell culture incubators. Cells were

tissue ulture dishes Cor i g™ Fal o ™ Eas -Grip Tissue Culture Dish

M hu a i suli ,

M

24

595

subcultivated for a maximum of 15 passages. The cell line was tested negative for contamination

596

with mycoplasma.

597 598

siRNA-mediated knockdown of CAII

599

CAII was knocked down in MCF-7 cells, using siRNA (Ambion Silencer® Select anti CA2 siRNA, s2249,

600

Life Technologies). Non-targeting negative control siRNA (Ambion Silencer® Select Negative Control

601

No. 1 siRNA) was used as control. Cells were transfected with 50 pmol of siRNA, using Lipofectamine

602

RNAiMAX transfection reagent (Life Technologies) in OptiMEM medium (Thermo Fisher). Transfected

603

cells were incubated for 3 days under normoxic or hypoxic conditions. Knockdown efficiency was

604

calculated by measurement of CAII RNA levels with quantitative real-time PCR and CAII protein levels

605

with western blot analysis.

606 607

Quantitative real-time PCR

608

Total RNA was extracted from MCF-7 cells, using the RNeasy MinElute Cleanup Kit (Qiagen GmbH,

609

Hilden, Germany). cDNA was synthesized with SuperScript III Reverse Transcriptase (Life

610

Technologies) and random hexamer primers (200 ng; #SO142, Fisher Scientific GmbH, Schwerte,

611

Germany). Real time PCR was carried out with SYBR Green (Po erUp™ SYBR Green Master Mix,

612

applied biosystems) on an AriaMx real-time PCR system (Agiland Technologies). All samples were run

613

in duplicate. RPL27 was used as a reference gene. The characteristics of the primers are shown in

614

Table 1. Calculation of the gene expression level was carried out using the comparative threshold

615

method 2-ΔΔCT.

616 617

Western blot analysis of CAII

618

To determine expression levels of CAII, MCF-7 cells were harvested by trypsinization and lysed in lysis

619

buffer (50 mM NaCl, 25 mM Tris, 0.5% (v/v) Triton X-100) with protease inhibitor (‘o he O plete™, 25

620

Mini, EDTA-free Protease Inhibitor Cocktail Tablets) for 1h at 4°C. The lysate was cleared from cell

621

debris by centrifugation. Protein concentration was determined with Bradford Reagent (Protein

622

Assay Dye Reagent Concentrate, Bio-Rad). 20 µg of total protein were separated on a 12%

623

polyacrylamide gel and blotted on a polyvinylidene membrane (Roti®-PVDF, Carl-Roth). Unspecific

624

binding sites were blocked with 5% milk powder, dissolved in PBS for 2 h. CAII was detected using

625

primary anti-CAII antibody (dilution 1:500; rabbit anti-carbonic anhydrase II polyclonal antibody,

626

AB1828, Millipore) and a goat anti-rabbit IgG horseradish peroxidase-conjugated secondary antibody

627

(dilution 1:2000; sc2004, Santa Cruz Biotechnology). After documentation the membrane was

628

washed 2x 10 min with stripping buffer (1.5% (w/v) glycine, 0.1% (w/v) SDS, 1% (v/v) Tween20, pH

629

. . As loadi g o trol β-Actin was labelled with Mouse-Anti-β-Actin (1:2500; sc-47778; Santa Cruz

630

Biotechnology). Quantification of the band intensity was carried out with the software ImageJ. The

631

sig al for CAII as or alized to the sig al for β-Actin in the same lane.

632 633

pH imaging in MCF-7 cells

634

Changes in intracellular pH in MCF-7 cells were measured with the pH sensitive dye

635

Seminaphthorhodafluor 5- and 6- ar o li a id F I itroge ™ “NA‘F F-AM, Life Technologies)

636

using a confocal laser scanning microscope (LSM 700 with AxioExaminer.D1 upright microscope,

637

equipped with a 40x water immersion objective (Zeiss C-Apochromat 40x/1.20 W, Carl Zeiss

638

Microscopy GmbH, Frankfurt, Germany). SNARF-5F is a ratiometric pH indicator which exhibits a

639

significant pH-dependent emission shift from yellow-orange to deep red fluorescence when

640

switching from acidic and basic pH (Han & Burgess, 2010). With a pKa of 7.2 SNARF-5 is

641

recommended for measuring cytosolic pHi (Han & Burgess, 2010). The fluorescence emission

642

spectrum of SNARF-5 has an isosbestic point at ~590 nm. The isosbestic point is the wavelength at

643

which fluorescence emission of SNARF-5 does not change under varying pH values, which provides

644

the possibility of ratiometric imaging. For ratiometric imaging SNARF-5 was excited at 555 nm with a

645

scanning frequency of 0.4 Hz. The emitted light was separated with a variable dichroic mirror at 590 26

646

nm in a 590 nm fraction and the signals of the 590 nm fraction.

648

MCF-7 cells were loaded with 10 µM of the acetoxymethyl ester of SNARF 5 for 10-15 minutes at RT

649

in a 35 x 10 mm tissue culture dish which was also used as bath chamber. Loading efficiency was

650

checked by taking snapshots from time to time in the setup. After loading, cells were constantly

651

perfused with medium at room temperature at a rate of 2 ml/min using a tubing system. The

652

medium had the following composition: 143 mM NaCl, 5 mM KCl, 2 mM CaCl 2, 1 mM MgSO4, 1 mM

653

Na2HPO4, 10 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES), pH 7.2. In lactate-

654

containing media NaCl was substituted by Na-L-lactate in equimolar amounts. To reproducibly

655

measure the rate of lactate-induced acidification, cells were depleted of lactate for at least 15

656

minutes prior to lactate application. For application of lactate, the application tube was switched

657

between different beakers containing the desired solutions.

658

Data analysis was carried out with ImageJ (National Institutes of Health, USA). For analysis a region of

659

interest (ROI) was drawn around each cell (Figure 1 – Figure Supplement 3 A3). The intensity of each

660

ROI was recorded for each frame and the values were stored digitally in a spread sheet.

661

In order to convert the fluorescent ratio into pH values, the system was calibrated by the use of

662

nigericin (10 µM; Life Technologies) as recently described (Forero-Quintero et al., 2017). Nigericin is a

663

microbial toxin derived from Streptomyces hygroscopicus, which functions as a K+/H+ exchanging

664

ionophore (Kovbasnjuk et al., 1991). At high K+ concentrations (130 mM K+ used in this study),

665

nigericine allows to equilibrate intracellular and extracellular pH. For calibration SNARF-5 loaded cells

666

were subsequently superfused with saline, adjusted to pH 6.0, 6.5, 7.0, 7.5, and 8.0, respectively, in

667

the presence of nigericin (Figure 1 – Figure Supplement 3 B). Steady state values were calculated by

668

exponential regression fitting for every single cell (Figure 1 – Figure Supplement 3 B). The fluorescent

669

ratio was blotted against the extracellular pH and the data were fitted using a Boltzmann function: y=

A − A

1+�

�−� ��

+A 27

670

with A1 = initial value, A2 = final value, x0 = center, dx = time constant. The resulting values are

671

given in the inset in Figure 1 – Figure Supplement 3 C.

672

Based on these data, pH values were calculated from the fluorescent ratio (R) for every data point

673

using the formula

674

pH = x + dx ∗ ln

675

A −A R−A

−1 .

The resulting pH values were plotted against the time, as exemplarily shown in Figure 1 A. From the

676

resulting curve the maximum rate of change in pHi ΔpHi/Δt duri g appli atio of la tate i flu a d

677

withdrawal (efflux) of lactate was determined for every single cell by linear regression fitting using

678

OriginPro 8.6.

679

For all imaging experiments the number of repetitions is given as n = number of cells / number of

680

batches.

681 682

Antibody staining of MCF-7 cells

683

MCF-7 cells, growing on glass cover slips, were rinsed twice in phosphate buffered saline (PBS) and

684

fixated in 4% paraformaldehyde in PBS for 20 min. Cells were permeabilized with 0.5% Triton X-100

685

and unspecific binding sites were blocked with 3% bovine serum albumin (BSA) and 1% normal goat

686

serum (NGS) for 2h at room temperature. Cells were incubated with primary antibodies (rabbit anti-

687

CAII polyclonal antibody, AB1828, Millipore, dilution 1:500 and goat anti-MCT1 polyclonal antibody,

688

sc-14917, Santa Cruz, dilution 1:300) overnight at 7°C. Cells were washed with PBS and incubated

689

with the secondary antibody (1:1000; Alexa Fluor 546 donkey anti-rabbit IgG and Alexa Fluor 488

690

donkey anti-goat IgG; Thermo Fisher). Cells were mounted on a microscope slide using mounting

691

ediu

ith DAPI ProLo g™ Gold A tifade Mou ta t ith DAPI, Ther o Fisher a d analyzed with

692

a confocal laser scanning microscope (Leica TCS SP5, Leica Microsystems, Wetzlar, Germany) with a

693

63x objective (HCX PL APO lambda blue 63x 1.4 OIL, Leica Microsystems).

694 695 28

696

In situ proximity ligation assay in MCF-7 cells

697

Colocalization of MCT1 and CAII was examined using the Duolink in situ Proximity Ligation Assay kit

698

(Sigma-Aldrich), as described by the manufacturer and by Söderberg et al., (2008). In short, MCF-7

699

cells, growing on glass coverslips were fixated in 4% paraformaldehyde solution (Roti-Histofi

700

Roth, Karlsruhe, Germany) for 30 min. After fixation, u spe ifi binding sites were blocked with 3%

701

bovine serum albumin (BSA; Sigma-Aldrich) and 1% normal goat serum (NGS; Sigma-Aldrich) for 2

702

hours at RT in a humid chamber. Cells were incubated with primary antibodies (rabbit anti-CAII

703

polyclonal antibody, AB1828, Millipore, dilution 1:200 and goat anti-MCT1 polyclonal antibody, sc-

704

14917, Santa Cruz, dilution 1:200) for 2 hours at room temperature. The following steps were

705

perfor ed a ordi g to the

706

incubated with the PLA probes for one hour at 37°C in a humid chamber. After washing, cells were

707

incubated with ligation-ligase solution. The ligation reaction was carried out at 37°C for 30 min in a

708

humid chamber. After washing, cells were incubated with the amplification-polymerase solution for

709

100 min at 37°C in a darkened humid chamber. For better visualization of the cells, F-actin was

710

stained with Alexa Fluor 488 Phalloidin (1:500; A12379, Life technologies). Cells were washed again

711

and mounted on a microscope slide using the mounting media supplied with the kit. Cells incubated

712

without primary antibodies were used as procedure controls. The resulting staining was visualized

713

using a confocal laser-scanning microscope (LSM 700; Carl Zeiss GmbH, Oberkochen, Germany),

714

equipped with a Zeiss EC Plan-Neofluar 40x/1.3 objective. Images were analyzed and the PLA signals

715

quantified using ImageJ.

%;

a ufa turer s proto ol. Briefl , the ells were washed two times and

716 717

Measurement of cell proliferation

718

120 µl of MCF-7 cell suspension (5x105 cells/ml) was added to 24-well plates, containing 70 µl

719

OptiMEM medium (Thermo Fisher), supplemented either with siRNA (Ambion Silencer® Select anti

720

CA2 siRNA, s2249 or Ambion Silencer® Select Negative Control No. 1 siRNA) and Lipofectamine

721

RNAiMAX transfection reagent (Life Technologies), or with the CA inhibitor EZA (Sigma Aldrich) and 29

722

the MCT1 inhibitor AR-C155858 (Santa Cruz Biotechnology), respectively. Cells were incubated for

723

3.5 h under normoxic conditions in a cell culture incubator. After incubation, the wells were filled up

724

to 0.75 µl with RPMI-1640 medium (Sigma-Aldrich), supplemented with 10% fetal calf serum, 2%

725

Mi i al Eagle s Mediu ,

726

penicillin/streptomycin, pH 7.2, and placed at 37°C in 5% CO2, 95% air (normoxia) or 5% CO2, 1% O2,

727

94% N2 (hypoxia) in humidified cell culture incubators. After addition of RPMI-1640 medium the

728

inhibitors were present at a concentration of 30 µM (EZA) and 300 nM (AR-C155858), respectively.

729

Cells were counted after 0 days (immediately after the 3.5 h incubation period), 1 day, 2 days, and 3

730

days. Cells were washed with PBS and fixed with 4% paraformaldehyde in PBS for 30 min. After

731

washing, nuclei were stained with 10 µM of the intercalating fluorescent dye Hoechst 33342

732

(ThermoFisher) for 20 minutes. Pictures were taken with a fluorescent microscope (Leica DM IRB),

733

equipped with a 10x objective (Leica 10x PH1). For every condition four wells were used. Three

734

images were taken from each well at random locations, yielding 12 pictures for every condition

735

(n=12/4). The number of nuclei per image was counted using the program ImageJ. Therefore each

736

picture was converted to a monochrome image (with threshold set to 30 – 255). Fused nuclei were

737

separated

738

set to 0-infinity, circularity set to 0.00-1.00). After counting the number of nuclei per picture was

739

converted to number of nuclei per mm2.

ith the o

.

a d

M hu a

i suli ,

atershed a d ou ted

M glu ose,

ith the o

M NaHCO3, and 1%

a d A al ze Parti les size

740 741

Site directed mutation of CAII

742

Site directed mutation of CAII was carried out by PCR using Phusion High-Fidelity DNA Polymerase

743

(Thermo Fisher) and modified primers, which contained the desired mutation. Primers used for

744

creation of the different mutants are shown in Table 2. Human CAII, cloned in the oocyte expression

745

vector pGEM-He-Juel, was used as template. PCR was cleaned up using the GeneJET PCR Purification

746

Kit (Thermo Fisher) and the template was digested with DpnI (Fermentas FastDigest DpnI, Thermo

747

Fisher) before transformation into E.coli DH5α cells. 30

748

Heterologous protein expression in Xenopus oocytes

749

cDNA coding for human CAII-WT or mutants of CAII, rat MCT1, and rat MCT4, respectively, cloned

750

into the oocyte expression vector pGEM-He-Juel was transcribed in vitro with T7 RNA-Polymerase

751

(mMessage mMachine, Ambion Inc., Austin, USA) as described earlier (Becker et al., 2004; Becker,

752

2014). Xenopus laevis females were purchased from the Radboud University, Nijmegen, Netherlands.

753

Segments of ovarian lobules were surgically removed under sterile conditions from frogs

754

anaesthetized with 1 g/l of ethyl 3-aminobenzoate methanesulfonate (MS-222, Sigma-Aldrich), and

755

rendered hypothermic. The procedure was approved by the Landesuntersuchungsamt Rheinland-

756

Pfalz,

757

Verbraucherschutz und Lebensmittelsicherheit, Oldenburg (33.19-42502-05-17A113). As described

758

earlier (Becker et al., 2004; Becker, 2014), oocytes were singularized by collagenase (Collagenase A,

759

Roche, Mannheim, Germany) treatment in Ca2+-free oocyte saline (pH 7.8) at 28°C for 2 h. The

760

singularized oocytes were left overnight in an incubator at 18°C in Ca2+-containing oocyte saline (pH

761

7.8) to recover. Oocytes of the stages V and VI were injected with 5 ng of cRNA coding for MCT1 or

762

MCT4, either together with 12 ng of cRNA coding for CAII or alone. Measurements were carried out 3

763

to 6 days after injection of cRNA. 4-MI was dissolved in DEPC-H2O at a concentration of 400 mM. 27.6

764

nl of the 4-MI solution were injected on the day of the experiments.

765

The oocyte saline had the following composition: 82.5 mM NaCl, 2.5 mM KCl, 1 mM CaCl 2, 1 mM

766

MgCl2, 1 mM Na2HPO4, 5 mM HEPES; titrated with NaOH to the desired pH. In CO2/HCO3-- and

767

lactate-containing saline, NaCl was substituted by NaHCO3 or Na-L-lactate in equimolar amounts.

Koblenz

(23

177-07/A07-2-003

§6)

and

the

Niedersächsisches

Landesamt

für

768 769

Measurement of intracellular H+ concentration in Xenopus oocytes

770

Changes in intracellular H+ concentration in oocytes were determined with ion-sensitive

771

microelectrodes under voltage-clamp conditions, using single-barreled microelectrodes. For

772

production of the pH-sensitive electrode a borosilicate glass capillary with filament, 1.5 mm in

31

773

diameter, was pulled to a micropipette with a tip opening of 0.5 – 1 µm. To achieve a hydrophobic

774

inner surface, the tip of the micropipette is backfilled with a drop of 5% tri-N-butylchlorosilane in

775

99.9% pure carbon-tetrachloride using a thin glass capillary and baked for 4.5 min at 450°C on a hot

776

plate under a fume hood. Silanized micropipettes were backfilled with a drop (approximately 0.5 µl)

777

of H+-selective cocktail (Hydrogen ionophore I - cocktail A, 95291, Sigma-Aldrich), using a thin glass

778

capillary and stored in a wet chamber for at least 30 min, so that the viscous cocktail can reach the

779

front end of the tip. Afterwards the cocktail is covered with 0.1 M Na-citrate solution, pH 6.0, to form

780

a liquid membrane. The electrode was connected to an amplifier (custom build at the University of

781

Kaiserslatern) with a chloride silver wire. For production of the reference and current electrodes

782

borosilicate glass capillaries with filament, 1.5 mm in diameter, were pulled to micropipettes with a

783

tip opening of 1 – 2 µm, backfilled with 3 M KCl and connected to an Axoclamp 2B amplifier (Axon

784

Instruments). For calibration of the pH-sensitive electrode, it was superfused with oocyte saline,

785

adjusted to pH 7.0, followed by superfusion with saline adjusted to pH 6.4. Calibration of the

786

electrode was carried out before every single experiment. After calibration of the pH-sensitive

787

electrode, the bath perfusion was topped and a single oocyte was placed into the bath. The

788

reference and current clamp electrodes were fist impaled into the oocyte. Then the pH-sensitive

789

ele trode

790

direction of flow of the bath perfusion. Since diffusion of protons within the cell is very slow, it is of

791

major importance to position the pH-sensitive microelectrode as close as possible to the inner face of

792

the plasma membrane in order to measure fast changes in intracellular H+-concentration (Bröer et

793

al., 1998). All experiments were carried out at room temperature (22-25°C). During the whole

794

experiment the oocyte was superfused with oocyte saline at a flow rate of 2 ml/min using a tubing

795

system. For application of lactate and CO2/HCO3-, respectively, the application tube was switched

796

between different beakers containing the desired solutions. The measurements were stored digitally

797

using custom made PC software (Neumann, 2018; copy archived at https://github.com/General-

798

Zoology/iClamp) based on the program LabView (National Instruments). A step by step instruction

as i paled a d positio ed to the fro t of the oo te

e

ra e, dire tl i to the

32

799

for the production of ion-sensitive microelectrodes and their use in Xenopus oocytes can be found in

800

Becker (2014).

801

For calculation of pH, the electrode potential (Ve), recorded during the calibration, was plotted

802

against the pH of the two calibration solutions (pH 7.0 and pH 6.4) and a linear fit was created, using

803

the spread sheet program OriginPro 8.6. Afterwards, the pH values for every recorded data point

804

were calculated using the formula

805

pH = intercept + slope * Ve.

806

Since pH is defined as -log ([H+]), the change in [H+] at a given change in pH depends on the baseline

807

pH, which could lead to misinterpretation of the real change in proton concentration. For this reason

808

proton concentration was calculated using the formula

809

[H+] = 10-pH x 109 (nM)

810

for every recorded data point.

811

Transport activity of the MCT was then determined by measuring the rate of change in [H +]i

812

(Δ[H+]i/Δt) during application or removal of substrate by linear fitting.

813

To make sure that the change in [H+]i during application of lactate is the result of MCT transport

814

activity and does not depend on passive diffusion of lactic acid or endogenous lactate transporters,

815

we previously determined Δ[H+]i/Δt duri g appli atio of la tate i Xenopus oocytes, injected with

816

H2O instead of cRNA (Becker et al., 2004). In H2O-injected oocytes no lactate-induced changes in [H+]i

817

could be observed, indicating that the lactate-induced acidification in MCT-expressing oocytes is the

818

result of MCT-mediated proton-coupled lactate transport (Becker et al., 2004).

819 820

Cal ulation of int insi

821

I tri si

822

carried out at the end of each experiment (for example see Figure 4B . βi was calculated from the

823

ha ge i i tra ellular pH ΔpHi) and the change in intracellular HCO3- concentration (Δ[HCO3-]i)

824

uffer apa it

uffe apa ity (βi) and lactate-induced proton flux (JH) βi) of oocytes was determined from the pulse of 5% CO2 / 10 mM HCO3-,

during application of CO2 using the formula 33

825

βi = ΔpHi / Δ[HCO3-]i (mM).

826

The exact calculations are shown on page three in the Figure Source Data 1 from Figure 5.

827

Since all experiments were carried out in the nominal absence of CO2, the CO2/HCO3--dependent

828

uffer apa it βCO2) was omitted from the calculation.

829

Lactate-induced proton flux (JH) was calculated by multiplying the rate of change in intracellular pH

830

duri g la tate appli atio

831

JH = ΔpH/t βi (mM).

832

The al ulatio of βi and JH is shown on page two in the Figure Source Data 1 from Figure 5.

ΔpH/t

ith βi using the formula

833 834

Pull-down of CAII

835

Pull-down of CAII with GST-fusion proteins of the C-termini of MCT1 and MCT4 has been described in

836

detail previously (Noor et al., 2015). In brief C-termini of MCT1 and MCT4 were cloned into the

837

expression vector pGEX-2T (GE Healthcare Europe GmbH) and transformed into E. coli BL21 cells.

838

Protein expression was induced by addition of 0.8 mM isopropyl-β-D-thiogalactopyranosid (IPTG). 3

839

hours after induction, cells were harvested, resuspended in phosphate-buffered saline (PBS) and

840

lysed in lysis buffer (PBS, supplemented with 2 mM MgCl2, 1% Triton X-100, and protease inhibitor

841

cocktail tablets from Roche). Bacterial lysates were centrifuged for 15 min at 4°C, 12000 x g and the

842

supernatant, containing the GST-fusion protein (bait protein), was collected for further use. CAII-WT

843

and mutants of CAII, respectively, were expressed in Xenopus oocytes. For each experiment 25

844

oocytes were lysed in lysis buffer. Oocyte lysates were centrifuged for 15 min at 4°C, 12000 x g and

845

the supernatant (prey protein) was collected for further use.

846

The pull-down experiment was carried out using the Pierce GST protein interaction pull-down kit

847

(Thermo Fisher). Briefly, for immobilization of GST-fusion protein, bacterial lysate was added to

848

glutathione agarose and incubated for 2 hours at 4°C with end-over-end mixing. After incubation, the

849

excess bait protein was removed by centrifugation and the beads were washed five times with wash

850

buffer (1 PBS : 1 lysis buffer). 400 µl of oocyte lysate, containing CAII, was added to the column and 34

851

incubated for 2 hours at 4 °C with end-over-end mixing. After incubation, the excess prey protein was

852

removed by centrifugation and the beads were washed five times with wash buffer. Protein was

853

eluted fro

854

To determine the relative amount of GST and CAII, an equal volume of the samples was analyzed by

855

western blotting. GST was detected using a primary anti-GST antibody (dilution 1:400; anti-GST tag

856

mouse monoclonal IgG, no. 05-782, Millipore) and a goat anti-mouse IgG horseradish peroxidase-

857

conjugated secondary antibody (dilution 1:2000; sc-2031, Santa Cruz). CAII was detected using

858

primary anti-CAII antibody (dilution 1:400; rabbit anti-carbonic anhydrase II polyclonal antibody,

859

AB1828, Millipore) and a goat anti-rabbit IgG horseradish peroxidase-conjugated secondary antibody

860

(dilution 1:2000; Santa Cruz). Quantification of the band intensity was carried out with the software

861

ImageJ. To overcome variations in the signal intensity between different blots, signal intensity of

862

each band for CAII was normalized to the signal intensity of the band from the pull-down of CAII-WT

863

with the GST-fusion protein of the C-terminal tail. To account for variations in the amount of GST-

864

fusion protein, each normalized signal for CAII was divided by the corresponding, normalized signal

865

for GST.

the eads ith

μl of elutio

uffer

M glutathio e i PB“, pH . .

866 867

Determination of CA catalytic activity via mass spectrometry

868

Catalytic activity of CAII in Xenopus oocytes was determined by monitoring the

869

doubly labelled 13C18O2 through several hydration and dehydration steps of CO2 and HCO3- at 24°C

870

(Silverman, 1982; Sültemeyer et al., 1990). The reaction sequence of 18O loss from 13C18O18O (m/z =

871

49) over the intermediate product 13C18O16O (m/z = 47) and the end product 13C16O16O (m/z = 45) was

872

monitored with a quadrupole mass spectrometer (OmniStar GSD 320; Pfeiffer Vacuum, Asslar,

873

Germany). The relative 18O enrichment was calculated from the measured 45, 47, and 49 abundance

874

as a function of time according to: log enrichment = log [49x100/(49+47+45)]. For the calculation of

875

CA activity, the rate of

876

over the time, using OriginPro 8.6. The rate was compared with the corresponding rate of the non-

18

18

O depletion of

O degradation was obtained from the linear slope of the log enrichment

35

877

catalyzed reaction. Enzyme activity in units (U) was calculated from these two values as defined by

878

Badger and Price (1989). From this definition, one unit corresponds to 100% stimulation of the non-

879

catalyzed 18O depletion of doubly labelled 13C18O2. For each experiment a batch of 20 native oocytes

880

or 20 oocytes expressing CAII-WT or a mutant of CAII, were lysed in 80 µl oocyte saline, pipetted into

881

the cuvette and the catalyzed degradation was determined for 10 min.

882 883

Calculation and statistics

884

Statistical values are presented as means ±standard error of the mean. Number of repetitions (n) is

885

indicated in each figure panel. For western blot analysis number of technical replicates (blots) /

886

number of biological replicates (cell lysates or pull-down eluates) are separated by a dash. For

887

calculation of significance in differences, Student´s t-test was used. In the figures shown, a

888

sig ifi a e le el of p ≤ .

is

arked ith *, p ≤ .

ith ** a d p ≤ .

ith ***.

889 890 891

Funding

892

The work was funded by the Deutsche Forschungsgemeinschaft (BE 4310/6-1; to H.M.B.), the

893

Stiftung Rheinland-Pfalz für Innovation (961-386261/957; to H.M.B.) the Research Initiative BioComp

894

(to H.M.B. and J.W.D), and the Landesschwerpunkt Membrantransport (to H.M.B. and J.W.D).

895 896

Conflict of interest

897

The authors declare no conflict of interest.

36

898

References

899

Ädelroth P, Brzezinski P. 2004. Surface-mediated proton-transfer reactions in membrane-bound

900

proteins.

901

10.1016/j.bbabio.2003.10.018.

902

Al-Samir S, Papadopoulos S, Scheibe RJ, Meißner JD, Cartron J-P, Sly WS, Alper SL, Gros G, Endeward

903

V. 2013. Activity and distribution of intracellular carbonic anhydrase II and their effects on the

904

transport activity of anion exchanger AE1/SLC4A1. The Journal of Physiology 591:4963–82. DOI:

905

10.1113/jphysiol.2013.251181.

906

Almquist J, Lang P, Prätzel-Wolters D, Deitmer JW, Jirstrand M, Becker HM. 2010. A Kinetic Model of

907

the Monocarboxylate Transporter MCT1 and its Interaction with Carbonic Anhydrase II. Journal of

908

Computer Science & Systems Biology 3:107–116. DOI: 10.4172/jcsb.1000066.

909

Ames S, Pastoreková S, Becker HM. 2018. The proteoglycan-like domain of carbonic anhydrase IX

910

mediates non-catalytic facilitation of lactate transport in cancer cells. Oncotarget [Accepted].

911

An H, Tu C, Duda DM, Montanez-Clemente I, Math K, Laipis PJ, McKenna R, Silverman DN. 2002.

912

Chemical rescue in catalysis by human carbonic anhydrases II and III. Biochemistry 41:3235–42.

913

Badger MR, Price GD. 1989. Carbonic Anhydrase Activity Associated with the Cyanobacterium

914

Synechococcus PCC7942. Plant Physiology 89:51–60. DOI: 10.1104/pp.89.1.51.

915

Barros LF, Deitmer JW. 2010. Glucose and lactate supply to the synapse. Brain Research Reviews 63.

916

Elsevier B.V.:149–159. DOI: 10.1016/j.brainresrev.2009.10.002.

917

Becker HM. 2014. Transport of Lactate: Characterization of the Transporters Involved in Transport at

918

the Plasma Membrane by Heterologous Protein Expression in Xenopus Oocytes. In: Neuromethods,

919

pp. 25–43. DOI: 10.1007/978-1-4939-1059-5_2.

Biochimica

et

Biophysica

Acta

-

Bioenergetics

1655:102–115.

DOI:

37

920

Becker HM, Bröer S, Deitmer JW. 2004. Facilitated lactate transport by MCT1 when coexpressed with

921

the sodium bicarbonate cotransporter (NBC) in Xenopus oocytes. Biophysical journal 86:235–47. DOI:

922

10.1016/S0006-3495(04)74099-0.

923

Becker HM, Deitmer JW. 2008. Nonenzymatic proton handling by carbonic anhydrase II during H+-

924

lactate cotransport via monocarboxylate transporter 1. The Journal of Biological Chemistry

925

283:21655–67. DOI: 10.1074/jbc.M802134200.

926

Becker HM, Deitmer JW. 2007. Carbonic anhydrase II increases the activity of the human electrogenic

927

Na+/HCO3-

928

10.1074/jbc.M700066200.

929

Becker HM, Hirnet D, Fecher-Trost C, Sültemeyer D, Deitmer JW. 2005. Transport activity of MCT1

930

expressed in Xenopus oocytes is increased by interaction with carbonic anhydrase. The Journal of

931

Biological Chemistry 280:39882–9. DOI: 10.1074/jbc.M503081200.

932

Becker HM, Klier M, Deitmer JW. 2010. Nonenzymatic augmentation of lactate transport via

933

monocarboxylate transporter isoform 4 by carbonic anhydrase II. The Journal of Membrane Biology

934

234:125–35. DOI: 10.1007/s00232-010-9240-y.

935

Becker HM, Klier M, Deitmer JW. 2014. Carbonic anhydrases and their interplay with acid/base-

936

coupled membrane transporters. In: Frost SC and McKenna R (eds) Sub-cellular biochemistry.

937

Subcellular Biochemistry. Dordrecht: Springer Netherlands, pp. 105–34. DOI: 10.1007/978-94-007-

938

7359-2_7.

939

Becker HM, Klier M, Schüler C, McKenna R, Deitmer JW. 2011. Intramolecular proton shuttle

940

supports not only catalytic but also noncatalytic function of carbonic anhydrase II. Proceedings of the

941

National

942

10.1073/pnas.1014293108.

943

Boone CD, Pinard M, McKenna R, Silverman DN. 2014. Catalytic mechanis

944

anhydrases: CO2 hydration and proton transfer. Frost SC and McKenna R (eds) Sub-cellular

cotransporter.

Academy

of

The

Sciences

Journal

of

the

of

Biological

United

States

Chemistry

of

282:13508–21.

America

108:3071–6.

DOI:

DOI:

of α-class carbonic

38

945

biochemistry

946

10.1007/978-94-007-7359-2_3.

947

Brändén M, Sandén T, Brzezinski P, Widengren J. 2006. Localized proton microcircuits at the

948

biological membrane-water interface. Proceedings of the National Academy of Sciences of the United

949

States of America 103:19766–70. DOI: 10.1073/pnas.0605909103.

950

Bröer S, Schneider HP, Bröer A, Rahman B, Hamprecht B, Deitmer JW. 1998. Characterization of the

951

monocarboxylate transporter 1 expressed in Xenopus laevis oocytes by changes in cytosolic pH. The

952

Biochemical journal 333:167–74.

953

Chiang WL, Chu SC, Yang SS, Li MC, Lai JC, Yang SF, Chiou HL, Hsieh YS. 2002. The aberrant expression

954

of cytosolic carbonic anhydrase and its clinical significance in human non-small cell lung cancer.

955

Cancer Letters 188:199–205. DOI: 10.1016/S0304-3835(02)00078-2.

956

Deitmer JW, Becker HM. 2013. Transport metabolons with carbonic anhydrases. Frontiers in

957

Physiology 4:291. DOI: 10.3389/fphys.2013.00291.

958

Duda DM, Tu C, Qian M, Laipis P, Agbandje-McKenna M, Silverman DN, McKenna R. 2001. Structural

959

and kinetic analysis of the chemical rescue of the proton transfer function of carbonic anhydrase II.

960

Biochemistry 40:1741–8.

961

Duda D, Govindasamy L, Agbandje-McKenna M, Tu C, Silverman DN, McKenna R. 2003. The refined

962

atomic structure of carbonic anhydrase II at 1.05 A resolution: implications of chemical rescue of

963

proton transfer. Acta crystallographica. Section D, Biological Crystallography 59:93–104.

964

Fisher SZ, Maupin CM, Budayova-Spano M, Govindasamy L, Tu C, Agbandje-McKenna M, Silverman

965

DN, Voth GA, McKenna R. 2007a. Atomic crystal and molecular dynamics simulation structures of

966

human carbonic anhydrase II: Insights into the proton transfer mechanism. Biochemistry 46:2930–

967

2937. DOI: 10.1021/bi062066y.

968

Fisher SZ, Tu C, Bhatt D, Govindasamy L, Agbandje-McKenna M, McKenna R, Silverman DN. 2007b.

969

Speeding Up Proton Transfer in a Fast Enzyme: Kinetic and Crystallographic Studies on the Effect of 39

75.

Subcellular

Biochemistry.

Dordrecht:

Springer

Netherlands:31–52.

DOI:

970

Hydrophobic Amino Acid Substitutions in the Active Site of Human Carbonic Anhydrase II.

971

Biochemistry 46:3803–3813. DOI: 10.1021/bi602620k.

972

Forero-Quintero LS, Deitmer JW, Becker HM. 2017. Reduction of epileptiform activity in ketogenic

973

mice: The role of monocarboxylate transporters. Scientific Reports 7:4900. DOI: 10.1038/s41598-017-

974

05054-0.

975

Friedman R, Nachliel E, Gutman M. 2005. Molecular dynamics of a protein surface: ion-residues

976

interactions. Biophysical Journal 89:768–81. DOI: 10.1529/biophysj.105.058917.

977

Gillies RJ, Verduzco D, Gatenby RA. 2012. Evolutionary dynamics of carcinogenesis and why targeted

978

therapy does not work. Nature reviews. Cancer 12:487–93. DOI: 10.1038/nrc3298.

979

Gross E, Pushkin A, Abuladze N, Fedotoff O, Kurtz I. 2002. Regulation of the sodium bicarbonate

980

cotransporter kNBC1 function: role of Asp986, Asp988 and kNBC1-carbonic anhydrase II binding. The

981

Journal of Physiology 544:679–685. DOI: 10.1113/jphysiol.2002.029777.

982

Haapasalo J, Nordfors K, Järvelä S, Bragge H, Rantala I, Parkkila A-K, Haapasalo H, Parkkila S. 2007.

983

Carbonic anhydrase II in the endothelium of glial tumors: a potential target for therapy. Neuro-

984

oncology 9:308–313. DOI: 10.1215/15228517-2007-001.

985

Han J, Burgess K. 2009. Fluorescent indicators for intracellular pH. Chemical Reviews 110:2709–2728.

986

DOI: 10.1021/cr900249z.

987

Hanahan D, Weinberg RA. 2011. Hallmarks of cancer: The next generation. Cell 144:646–674. DOI:

988

10.1016/j.cell.2011.02.013.

989

Håkansson K, Carlsson M, Svensson LA, Liljas A. 1992. Structure of native and apo carbonic anhydrase

990

II and structure of some of its anion-ligand complexes. Journal of Molecular Biology 227:1192–204.

991

DOI: 10.1016/0022-2836(92)90531-N.

40

992

Heidtmann H, Ruminot I, Becker HM, Deitmer JW. 2015. Inhibition of monocarboxylate transporter

993

by N-cyanosulphonamide S0859. European Journal of Pharmacology 762:344–349. DOI:

994

10.1016/j.ejphar.2015.05.049.

995

Innocenti A, Pastorekova S, Pastorek J, Scozzafava A, De Simone G, Supuran CT. 2009. The

996

proteoglycan region of the tumor-associated carbonic anhydrase isoform IX acts as anintrinsic buffer

997

optimizing CO2 hydration at acidic pH values characteristic of solid tumors. Bioorganic & Medicinal

998

Chemistry Letters 19:5825–8. DOI: 10.1016/j.bmcl.2009.08.088.

999

Isom DG, Castaneda CA, Cannon BR, Garcia-Moreno E. B. 2011. Large shifts in pKa values of lysine

1000

residues buried inside a protein. Proceedings of the National Academy of Sciences 108:5260–5265.

1001

DOI: 10.1073/pnas.1010750108.

1002

Jamali S, Klier M, Ames S, Barros LF, McKenna R, Deitmer JW, Becker HM. 2015. Hypoxia-induced

1003

carbonic anhydrase IX facilitates lactate flux in human breast cancer cells by non-catalytic function.

1004

Scientific reports 5:13605. DOI: 10.1038/srep13605.

1005

Kennedy KM, Dewhirst MW. 2010. Tumor metabolism of lactate: the influence and therapeutic

1006

potential for MCT and CD147 regulation. Future Oncology (London, England) 6:127–48. DOI:

1007

10.2217/fon.09.145.

1008

Kifor G, Toon MR, Janoshazi A, Solomon AK. 1993. Interaction between red cell membrane band 3

1009

and cytosolic carbonic anhydrase. The Journal of Membrane Biology 134:169–179.

1010

Klier M, Andes FT, Deitmer JW, Becker HM. 2014. Intracellular and extracellular carbonic anhydrases

1011

cooperate non-enzymatically to enhance activity of monocarboxylate transporters. The Journal of

1012

Biological Chemistry 289:2765–75. DOI: 10.1074/jbc.M113.537043.

1013

Klier M, Schüler C, Halestrap AP, Sly WS, Deitmer JW, Becker HM. 2011. Transport activity of the

1014

high-affinity monocarboxylate transporter MCT2 is enhanced by extracellular carbonic anhydrase IV

1015

but not by intracellular carbonic anhydrase II. The Journal of Biological Chemistry 286:27781–91. DOI:

1016

10.1074/jbc.M111.255331. 41

1017

Kovbasnjuk ON, Antonenko YN, Yaguzhinsky LS. 1991. Proton dissociation from nigericin at the

1018

membrane-water interface, the rate-limiting step of K+/H+ exchange on the bilayer lipid membrane.

1019

FEBS letters 289:176–8. Available at: http://www.ncbi.nlm.nih.gov/pubmed/1655522.

1020

Li X-J, Xie H-L, Lei S-J, Cao H-Q, Meng T-Y, Hu Y-L. 2012. Reduction of CAII Expression in Gastric

1021

Cancer: Correlation with Invasion and Metastasis. Chinese Journal of Cancer Research 24:196–200.

1022

DOI: 10.1007/s11670-012-0196-6.

1023

Li X, Alvarez B V, Casey JR, Reithmeier RAF, Fliegel L. 2002. Carbonic anhydrase II binds to and

1024

enhances activity of the Na+/H+ exchanger. The Journal of Biological Chemistry 277:36085–91. DOI:

1025

10.1074/jbc.M111952200.

1026

Li X, Liu Y, Alvarez B V, Casey JR, Fliegel L. 2006. A novel carbonic anhydrase II binding site regulates

1027

NHE1 activity. Biochemistry 45:2414–24. DOI: 10.1021/bi051132d.

1028

Lindskog S. 1997. Structure and mechanism of carbonic anhydrase. Pharmacology & Therapeutics

1029

74:1–20. Available at: http://www.ncbi.nlm.nih.gov/pubmed/9336012.

1030

Loiselle FB. 2004. Regulation of the human NBC3 Na+/HCO3- cotransporter by carbonic anhydrase II

1031

and PKA. AJP: Cell Physiology 286:1423–1433. DOI: 10.1152/ajpcell.00382.2003.

1032

Lu J, Daly CM, Parker MD, Gill HS, Piermarini PM, Pelletier MF, Boron WF. 2006. Effect of human

1033

carbonic anhydrase II on the activity of the human electrogenic Na/HCO3 cotransporter NBCe1-A in

1034

Xenopus oocytes. The Journal of biological chemistry 281:19241–50. DOI: 10.1074/jbc.M602181200.

1035

Magistretti PJ. 2006. Neuron-glia metabolic coupling and plasticity. The Journal of Experimental

1036

Biology 209:2304–11. DOI: 10.1242/jeb.02208.

1037

Mallory JC, Crudden G, Oliva A, Saunders C, Stromberg A, Craven RJ. 2005. A novel group of genes

1038

regulates susceptibility to antineoplastic drugs in highly tumorigenic breast cancer cells. Molecular

1039

Pharmacology 68:1747–1756. DOI: 10.1124/mol.105.016519.

42

1040

Martínez C, Kalise D, Barros LF. 2010. General requirement for harvesting antennae at Ca and H

1041

channels and transporters. Frontiers in Neuroenergetics 2:1–8. DOI: 10.3389/fnene.2010.00027.

1042

McMurtrie HL, Cleary HJ, Alvarez B V, Loiselle FB, Sterling D, Morgan PE, Johnson DE, Casey JR. 2004.

1043

The bicarbonate transport metabolon. Journal of Enzyme Inhibition and Medicinal Chemistry 19:231–

1044

236. DOI: 10.1080/14756360410001704443.

1045

Moraes TF, Reithmeier RAF. 2012. Membrane transport metabolons. Biochimica et Biophysica Acta -

1046

Biomembranes 1818:2687–2706. DOI: 10.1016/j.bbamem.2012.06.007.

1047

Neumann G. 2018. iClamp. Github. https://github.com/General-Zoology/iClamp

1048

Noor SI, Dietz S, Heidtmann H, Boone CD, McKenna R, Deitmer JW, Becker HM. 2015. Analysis of the

1049

binding moiety mediating the interaction between monocarboxylate transporters and carbonic

1050

anhydrase II. Journal of Biological Chemistry 290:4476–4486. DOI: 10.1074/jbc.M114.624577.

1051

Noor “I, Pou ssegur J, Deit er JW, Be ker HM.

1052

transport activity of the monocarboxylate transporter MCT4. The FEBS Journal 284:149–162. DOI:

1053

10.1111/febs.13964.

1054

Ovens MJ, Davies AJ, Wilson MC, Murray CM, Halestrap AP. 2010a. AR-C155858 is a potent inhibitor

1055

of monocarboxylate transporters MCT1 and MCT2 that binds to an intracellular site involving

1056

transmembrane helices 7-10. The Biochemical Journal 425:523–30. DOI: 10.1042/BJ20091515.

1057

Ovens MJ, Manoharan C, Wilson MC, Murray CM, Halestrap AP. 2010b. The inhibition of

1058

monocarboxylate transporter 2 (MCT2) by AR-C155858 is modulated by the associated ancillary

1059

protein. Biochemical Journal 431:217–225. DOI: 10.1042/BJ20100890.

1060

Parkkila AK, Herva R, Parkkila S, Rajaniemi H. 1995. Immunohistochemical demonstration of human

1061

carbonic anhydrase isoenzyme II in brain tumours. The Histochemical Journal 27:974–982. DOI:

1062

10.1007/BF00175572.

. I tegratio of a proto a te

a fa ilitates

43

1063

Parks SK, Chiche J, Pouysségur J. 2013. Disrupting proton dynamics and energy metabolism for cancer

1064

therapy. Nature Reviews Cancer 13:611–23. DOI: 10.1038/nrc3579.

1065

Parks SK, Chiche J, Pouyssegur J. 2011. pH control mechanisms of tumor survival and growth. Journal

1066

of Cellular Physiology 226:299–308. DOI: 10.1002/jcp.22400.

1067

Pastorek J, Pastorekova S. 2015. Hypoxia-induced carbonic anhydrase IX as a target for cancer

1068

therapy:

1069

10.1016/j.semcancer.2014.08.002.

1070

Pellerin L, Magistretti PJ. 1994. Glutamate uptake into astrocytes stimulates aerobic glycolysis: a

1071

mechanism coupling neuronal activity to glucose utilization. Proceedings of the National Academy of

1072

Sciences of the United States of America 91:10625–9.

1073

Piermarini PM, Kim EY, Boron WF. 2007. Evidence against a direct interaction between intracellular

1074

carbonic anhydrase II and pure C-terminal domains of SLC4 bicarbonate transporters. The Journal of

1075

Biological Chemistry 282:1409–21. DOI: 10.1074/jbc.M608261200.

1076

Poole RC, Halestrap AP. 1993. Transport of lactate and other monocarboxylates across mammalian

1077

plasma membranes. The American journal of physiology 264:761-82.

1078

Schulze A, Harris AL. 2012. How cancer metabolism is tuned for proliferation and vulnerable to

1079

disruption. Nature 491:364–73. DOI: 10.1038/nature11706.

1080

Sheng W, Dong M, Zhou J, Li X, Dong Q. 2013. Down regulation of CAII is associated with tumor

1081

differentiation and poor prognosis in patients with pancreatic cancer. Journal of Surgical Oncology

1082

107:536–43. DOI: 10.1002/jso.23282.

1083

Shinobu A, Agmon N. 2009. Mapping proton wires in proteins: carbonic anhydrase and GFP

1084

chromophore biosynthesis. Journal of Physical Chemistry A 113:7253–7266. DOI: 10.1021/jp8102047.

1085

Silverman DN. 1982. Carbonic anhydrase: oxygen-18 exchange catalyzed by an enzyme with rate-

1086

contributing proton-transfer steps. Methods in enzymology 87:732–52.

From

biology

to

clinical

use.

Seminars

in

Cancer

Biology

31:52–64.

DOI:

44

1087

Söderberg O, Leuchowius K-J, Gullberg M, Jarvius M, Weibrecht I, Larsson L-G, Landegren U. 2008.

1088

Characterizing proteins and their interactions in cells and tissues using the in situ proximity ligation

1089

assay. Methods (San Diego, Calif.) 45:227–32. DOI: 10.1016/j.ymeth.2008.06.014.

1090

Steiner H, Jonsson B ‐H, Lindskog S. 1975. The Catalytic Mechanism of Carbonic Anhydrase:

1091

Hydrogen‐Isotope Effects on the Kinetic Parameters of the Human C Isoenzyme. European Journal of

1092

Biochemistry 59:253–259. DOI: 10.1111/j.1432-1033.1975.tb02449.x.

1093

Sterling D, Reithmeier RAF, Casey JR. 2001. A transport metabolon: Functional interaction of carbonic

1094

anhydrase II and chloride/bicarbonate exchangers. Journal of Biological Chemistry 276:47886–47894.

1095

DOI: 10.1074/jbc.M105959200.

1096

Stridh MH, Alt MD, Wittmann S, Heidtmann H, Aggarwal M, Riederer B, Seidler U, Wennemuth G,

1097

McKenna R, Deitmer JW, Becker HM. 2012. Lactate flux in astrocytes is enhanced by a non-catalytic

1098

action

1099

10.1113/jphysiol.2011.220152.

1100

Sültemeyer DF, Fock HP, Canvin DT. 1990. Mass Spectrometric Measurement of Intracellular

1101

Carbonic Anhydrase Activity in High and Low C(i) Cells of Chlamydomonas: Studies Using O Exchange

1102

with C/O Labeled Bicarbonate. Plant Physiology 94:1250–1257.

1103

Tu CK, Silverman DN, Forsman C, Jonsson BH, Lindskog S. 1989. Role of histidine 64 in the catalytic

1104

mechanism of human carbonic anhydrase II studied with a site-specific mutant. Biochemistry

1105

28:7913–8.

1106

Ullah MS, Davies AJ, Halestrap AP. 2006. The plasma membrane lactate transporter MCT4, but not

1107

MCT1, is up-regulated by hypoxia through a HIF- α-dependent mechanism. Journal of Biological

1108

Chemistry 281:9030–9037. DOI: 10.1074/jbc.M511397200.

1109

Vince JW, Carlsson U, Reithmeier RAF. 2000. Localization of the Cl-/HCO3- anion exchanger binding

1110

site to the amino-terminal region of carbonic anhydrase II. Biochemistry 39:13344–9.

of

carbonic

anhydrase

II.

The

Journal

of

Physiology

590:2333–51.

DOI:

45

1111

Vince JW, Reithmeier RAF. 1998. Carbonic anhydrase II binds to the carboxyl terminus of human

1112

band 3, the erythrocyte C1-/HCO3- exchanger. The Journal of biological chemistry 273:28430–7.

1113

Vince JW, Reithmeier RAF. 2000. Identification of the carbonic anhydrase II binding site in the Cl(-

1114

)/HCO(3)(-) anion exchanger AE1. Biochemistry 39:5527–33.

1115

Yamada H, Horita S, Suzuki M, Fujita T, Seki G. 2011. Functional role of a putative carbonic anhydrase

1116

II-binding domain in the electrogenic Na+-HCO3- cotransporter NBCe1 expressed in Xenopus oocytes.

1117

Channels (Austin, Tex.) 5:106–9. DOI: 10.1073/pnas.1008705107.he.

1118

Zeuthen T, Zeuthen E, Klaerke DA. 2002. Mobility of ions, sugar, and water in the cytoplasm of

1119

Xenopus oocytes expressing Na+-coupled sugar transporters (SGLT1). Journal of Physiology 542:71–

1120

87. DOI: 10.1113/jphysiol.2001.014530.

1121

Zhou Y, Mokhtari RB, Pan J, Cutz E, Yeger H. 2015. Carbonic anhydrase II mediates malignant

1122

behavior of pulmonary neuroendocrine tumors. American Journal of Respiratory Cell and Molecular

1123

Biology 52:183–192. DOI: 10.1165/rcmb.2014-0054OC.

46

1124

Tables

1125 1126

Table 1: Characteristics of primers used for qRT-PCR Gene CA II RPL27

Primer

Se 5’→ 3’

Tm (°C)

forward reverse forward reverse

AAACAAAGGGCAAGAGTGCTGACT TTTCAACACCTGCTCGCTGCTG GGTGGTTGCTGCCGAAATGGG TGTTCTTCACGATGACAGCTTTGCG

57.2 58.3 58.9 58.6

Amplicon size (bp) 173 101

Location 680-703 831-852 30-50 106-130

1127 1128 1129 1130

Table 2: Characteristics of primers used for single-site mutation of CAII

1131

Shown are the sense primers for single-site mutation CAII. Nucleotides that differ from the wild-type

1132

sequence are labelled in blue. The antisense primers had the inverse complement sequence of the

1133

sense primers. Multiple mutations of CAII were created from single-site mutants, using the primers

1134

given in the table. CAII-H64A, also used in this study, was kindly provided by Dr. Robert McKenna,

1135

University of Florida, Gainesville, U.S.A.

Mutation

Se 5’→ 3’

Tm (°C)

CAII-H3A

CCGAGGATGTCCGCTCACTGGGGGTACGGC

76.3

CAII-H4A

CCGAGGATGTCCCATGCCTGGGGGTACGGC

76.3

CAII-H3A/H4A

CCGAGGATGTCCGCTGCCTGGGGGTACGGC

77.7

CAII-K9A

GGGTACGGCGCACACAACGGACCTGAG

72.6

CAII-H10A

GGGTACGGCAAAGCCAACGGACCTGAG

71.0

CAII-H15A

GGACCTGAGGCCTGGCATAAGGACTTCC

71.0

CAII-H64K

CAACAATGGTAAAGCTTTCAACG

57.1

CAII-E69Q

CATGCTTTCAACGTGCAGTTTGAT

59.3

CAII-D71N

GGAGTTTAATGACTCTCAGG

55.2

CAII-D72N

TTTGATAACTCTCAGGACAAAGCA

57.6

1136

47

1137

Figure Legends

1138

Figure 1. CAII facilitates lactate-induced proton flux in MCF-7 breast cancer cells. (A, B) Original

1139

recordings of lactate-induced changes in intracellular pH (pHi) in normoxic (A) and hypoxic (B) MCF-7

1140

breast cancer cells either treated with negative control siRNA (control, black traces) or CAII-siRNA

1141

(CAII knockdown, blue traces). (C, D ‘ate of ha ge i i tra ellular pH ΔpHi/Δt) during application

1142

(C) and withdrawal (D) of lactate in normoxic and hypoxic MCF-7 breast cancer cells either treated

1143

with negative control siRNA or CAII-siRNA (mean + SEM). Knockdown of CAII results in a significant

1144

reduction of lactate-induced pH change both under normoxic and hypoxic conditions. The black

1145

asterisks above the bars for CAII knockdown cells refer to the corresponding bars of the control cells.

1146

The blue and gray significance indicators above the bars for hypoxic cells refer to the corresponding

1147

bars of normoxic cells. *p≤ .

1148

The following figure supplement is available for figure 1:

1149

Figure 1 - figure supplement 1. Determination of CAII knockdown efficiency in MCF-7 cells.

1150

Figure 1 - figure supplement 2. Influence of pHi on lactate transport.

1151

Figure 1 - figure supplement 3. Calibration of SNARF-5 in MCF-7 cells.

1152

Figure 1 - source data 1. Original dataset for Figure 1.

, **p≤ .

, ***p≤0.001, n.s. no significance; “tude t s t-test.

1153 1154

Figure 2. MCT1 and CAII are co-localized in MCF-7 breast cancer cells. (A) Antibody staining of CAII

1155

(A1) and MCT1 (A2) in MCF-7 cells. (A3) Overlay of the fluorescence signals for MCT1 (red), CAII

1156

(green) and the nuclei marker DAPI (blue). Specificity of primary antibodies was tested by incubating

1157

MCF-7 cells only with secondary antibodies (A4). (B) in situ proximity ligation assay (PLA) of MCT1 and

1158

CAII in MCF-7 breast cancer cells, incubated under normoxia (B1) and hypoxia (B2), respectively, and

1159

normoxic MCF-7 cells in which CAII was knocked down using siRNA (B3). The red dots indicate co-

1160

localization of MCT1 and CAII with a maximum distance of < 40 nm. (B4) Negative control of an in situ

1161

PLA without primary antibodies. For better visualization of the cells, F-actin was stained with

1162

fluorescence labelled phalloidin (green). (C) Quantification of the PLA signals, shown as signals per 48

1163

nucleus (mean + SEM). The significance indicators above the bars refer to the values of the PLA for

1164

normoxic cells. ***p≤ .

1165

The following figure supplement is available for figure 2:

1166

Figure 2 - source data 1. Original dataset for Figure 2.

, n.s. no significance; “tude t s t-test.

1167 1168

Figure 3. CAII supports proliferation of MCF-7 breast cancer cells. (A, B) Staining of nuclei with

1169

Hoechst 33342 (blue) in MCF-7 cells after 3 days in culture under normoxic (A) or hypoxic (B)

1170

conditions. Cells remained either untreated (control; A1, B1), mock-transfected with non-targeting

1171

negative control siRNA (siNeg; A2, B2), transfected with siRNA against CAII (siCAII; A3, B3), incubated

1172

with 30 µM of the CA inhibitor EZA (A4, B4), or incubated with 300 nM of the MCT1 inhibitor AR-

1173

C155858 (A5, B5). (C, D) Total number of nuclei/mm2 in normoxic (C) and hypoxic (D) MCF-7 cell

1174

cultures, kept for 0–3 days under the conditions as described in (A) and (B). For every data point four

1175

dishes of cells were used and three pictures were taken from each dish at random locations, yielding

1176

12 pictures/data point (n = 12/4). The blue asterisks indicate significance in differences between

1177

siCAII and siNeg, the orange asterisks between control and AR-C155858, and the yellow significance

1178

indicators between control and EZA. *p≤ .

1179

test.

1180

The following figure supplement is available for figure 3:

1181

Figure 3 - source data 1. Original dataset for Figure 3.

, **p≤ .

, ***p≤ .

, .s. o sig ifi a e; “tude t s t-

1182 1183

Figure 4. CAII-Glu69 and Asp72 are crucial for the facilitation of MCT1 transport activity. (A)

1184

Structural model of human CAII (PDB-ID: 2CBA; Håkansson et al., 1992). Glu69 and Asp72, which have

1185

been suggested to form a proton-collecting antenna are labelled in red and green, respectively. The

1186

adjacent Asp71 is labelled in gray. His64, the central residue of the intramolecular proton shuttle, is

1187

labelled in yellow. (B) Original recordings of the change in intracellular H+ concentration in oocytes

1188

expressing MCT1 (black trace), MCT1 + CAII-WT (blue trace), MCT1 + CAII-E69Q (red trace), MCT1 + 49

1189

CAII-D71N (gray trace), and MCT1 + CAII-D72N (green trace), respectively, during application of 3 and

1190

10 mM of lactate and application of 5% CO2 / 10 mM HCO3-. (C, D) Rate of change in intracellular H+

1191

o e tratio

Δ[H+]/Δt as i du ed

appli atio

C a d re o al D of

a d

M la tate,

1192

respectively, in oocytes expressing MCT1 (dark gray), MCT1 + CAII-WT (blue), MCT1 + CAII-E69Q

1193

(red), MCT1 + CAII-D71N (light gray), and MCT1 + CAII-D72N (green), respectively (mean + SEM). The

1194

black significance indicators refer to MCT1, the blue significant indicators refer to MCT1 + CAII-WT.

1195

*p≤ .

1196

The following figure supplement is available for figure 4:

1197

Figure 4 - figure supplement 1. CAII-Glu69 and Asp72 are crucial for the facilitation of MCT4

1198

transport activity.

1199

Figure 4 - figure supplement 2. Expression levels of CAII are not altered by single-site mutation.

1200

Figure 4 - source data 1. Original dataset for Figure 4.

, **p≤0.01, ***p≤ .

, n.s. no significance; “tude t s t-test.

1201 1202

Figure 5. Chemical rescue of the interaction between MCT1 and CAII-E69Q/D72N by 4-

1203

methylimidazole. (A) Structural model of human CAII, complexed with 4-methylimidazole (4-MI)

1204

(PDB-ID: 1MOO; Duda et al., 2003). 4-MI binds near His64 or in a moiety between Glu69 and Asp 72.

1205

Inset: Close-up of the binding moiety. 4-MI can bind in two alternative confirmations (orange and

1206

green) between Leu57, Asn67, Glu69, Asp72, and Ile91. (B) Original recordings of the change in

1207

intracellular H+ concentration in oocytes expressing MCT1 (black trace), MCT1 + CAII-WT (blue trace),

1208

MCT1 + CAII-E69Q (red trace), MCT1 + CAII-D71N (gray trace), and MCT1 + CAII-D72N (green trace),

1209

respectively, during application of 3 and 10 mM lactate. All oocytes were injected with 4-MI (30 mM)

1210

the day the measurements were carried out. (C, D) Rate of change in intracellular H+ concentration

1211

Δ[H+]/Δt as i du ed

appli atio

C a d re o al D of

a d

M la tate, respe ti el , i

1212

oocytes expressing MCT1 (dark gray), MCT1 + CAII-WT (blue), MCT1 + CAII-E69Q (red), MCT1 + CAII-

1213

D71N (light gray), and MCT1 + CAII-D72N (green), respectively (mean + SEM). The black significance

1214

indicators refer to MCT1, the blue significance indicators refer to MCT1+CAII. All oocytes were 50

1215

injected with 4-MI (30 mM) the day the measurements were carried out. (E) Lactate-induced proton

1216

flux (JH), as calculated from the rate of change in pHi a d the ells i tri si

1217

supplement 1), in oocytes expressing MCT1 (dark gray), MCT1 + CAII-WT (blue), MCT1 + CAII-E69Q

1218

(red), MCT1 + CAII-D71N (light gray), and MCT1 + CAII-D72N (green), respectively, either injected

1219

with 4-MI or not (mean + SEM). The significance indicators above the bars from 4-MI-injected

1220

oocytes refer to the cells expressing the same proteins without 4-MI. *p≤ .

1221

***p≤ .

1222

The following figure supplement is available for figure 5:

1223

Figure 5 - figure supplement 1. Intrinsic buffer capacity of oocytes.

1224

Figure 5 - figure supplement 2. Chemical rescue of the interaction between MCT4 and CAII-

1225

E69Q/D72N by 4-methylimidazole.

1226

Figure 5 - source data 1. Original dataset for Figure 5.

uffer apa it

βi; figure

, **p≤ .

,

, n.s. no significance; “tude t s t-test.

1227 1228

Figure 6. CAII-Glu69 and Asp72 do not mediate binding between MCT1/4 and CAII. (A, C)

1229

Representative western blots for CAII (upper panel) and GST (lower panel). CAII-WT, CAII-E69Q, CAII-

1230

D71N and CAII-D72N were pulled down with a GST fusion protein of the C-terminal of (A) MCT1 (GST-

1231

MCT1-CT) and (C) MCT4 (GST-MCT4-CT), respectively. As negative control CAII-WT was pulled down

1232

with GST alone. (B, D) Relative intensity of the fluorescent signal for CAII (mean + SEM). The signal

1233

intensity of CAII, pulled down with GST-MCT1/4-CT was set to 100%. The significance indicators refer

1234

to the original values of CAII-WT. ***p≤0.001, n.s. no significance; “tude t s t-test.

1235

The following figure supplement is available for figure 6:

1236

Figure 6 - source data 1. Original dataset for Figure 6.

1237 1238

Figure 7. CAII-Glu69 and Asp72 do not support CAII catalytic activity. (A) Original recording of the

1239

log enrichment (LE) of either a lysate of 20 native oocytes (black trace) or a lysate of 20 oocytes

1240

expressing CAII-WT (blue), CAII-H64A (yellow), CAII-E69Q (red), CAII-D71N (gray), CAII-D72N (green), 51

1241

or the double mutant CAII-E69Q/D72N (turquois). The beginning of the traces shows the rate of

1242

degradation of the 18O-labelled substrate in the non-catalyzed reaction; the black arrow indicates the

1243

addition of oocyte lysate. (B) Enzymatic activity of CA in units/ml (mean + SEM). The black asterisks

1244

refer to the values from native oocytes, the blue significance indicators refer to the values of oocytes

1245

expressing CAII-WT. (C, D) Rate of change in intracellular H+ concentratio

1246

application (C) and removal (D) of 5% CO2 / 10 mM HCO3- in native oocytes and oocytes expressing

1247

CAII-WT, CAII-H64A, CAII-E69Q, CAII-D71N, CAII-D72N or the double mutant CAII-E69Q/D72N (mean

1248

+ SEM). **p≤ .

1249

The following figure supplement is available for figure 7:

1250

Figure 7 - source data 1. Original dataset for Figure 7.

, ***p≤ .

Δ[H+]/Δt as i du ed

, .s. o sig ifi a e; “tude t s t-test.

1251 1252

Figure 8. CAII-His64 mediates binding, but no proton transfer between MCT1 and CAII. (A)

1253

Structural model of the binding between the C-terminal tail of MCT1 (raspberry) and CAII (blue).

1254

Binding is mediated by Glu489 (red) and Glu491 (orange) in the MCT1 C-terminal tail and His64

1255

(yellow) in CAII. (B) Original recordings of the change in intracellular H+ concentration in oocytes

1256

expressing MCT1 (black trace), MCT1 + CAII-WT (blue trace), MCT1 + CAII-H64A (yellow trace), and

1257

MCT1 + CAII-H64K (green trace), respectively, during application of 3 and 10 mM of lactate and

1258

application of 5% CO2 / 10 mM HCO3-. (C, D) Rate of change in intracellular H+ concentration

1259

Δ[H+]/Δt as i du ed

appli atio

C a d re o al D of

a d

M la tate, respe ti el , i

1260

oocytes expressing MCT1 (gray), MCT1 + CAII-WT (blue), MCT1 + CAII-H64A (yellow), and MCT1 +

1261

CAII-H64K (green), respectively (mean + SEM). The black significance indicators refer to MCT1, the

1262

blue significance indicators refer to MCT1 + CAII-WT. (E) Representative western blots for CAII (upper

1263

panel) and GST (lower panel). CAII-WT, CAII-H64A, and CAII-H64K were pulled down with a GST

1264

fusion protein of the C-terminal of MCT1 (GST-MCT1-CT). As negative control CAII-WT was pulled

1265

down with GST alone. (F) Relative intensity of the fluorescent signal for CAII (mean + SEM). The signal

52

1266

intensity of CAII, pulled down with GST-MCT1-CT was set to 100%. The significance indicators refer to

1267

the original values of CAII-WT. ***p ≤ .

1268

The following figure supplement is available for figure 8:

1269

Figure 8 - figure supplement 1. CAII-His64 mediates binding, but no proton transfer between MCT4

1270

and CAII.

1271

Figure 8 - figure supplement 2. Catalytic activity and expression levels of CAII His64 mutants.

1272

Figure 8 - source data 1. Original dataset for Figure 8.

, .s. o sig ifi a e, “tude t s t-test.

1273 1274

Figure 9. The histidine residues in the N-terminus of CAII are not involved in the interaction

1275

between CAII and MCT1. (A) Structural model of human CAII (PDB-ID: 1XEV). His3, His4, His10, Lys9,

1276

and His15, which have been suggested to mediate binding of CAII to various acid/base transporters,

1277

are labelled in green. His64, the biding site for MCT1 and MCT4, is labelled in yellow. (B) Rate of

1278

change in intracellular H+ o e tratio

1279

in oocytes expressing MCT1 (dark gray), MCT1 + CAII-WT (blue), or MCT1 + one of the CAII mutants

1280

(green) (mean + SEM). The black significance indicators refer to MCT1, the blue significant indicators

1281

refer to MCT1 + CAII-WT. ***p≤ .

1282

The following figure supplement is available for figure 9:

1283

Figure 9 - source data 1. Original dataset for Figure 9.

Δ[H+]/Δt as i du ed

appli atio of

a d

M la tate,

. .s. o sig ifi a e; “tude t s t-test.

1284 1285

Figure 10. Carbonic anhydrases function as proton antennae for MCTs in glycolytic cancer cells.

1286

Intracellular CAII (blue circle) is anchored to the C-terminal tail of MCT1/4 (raspberry structure) via

1287

CAII-His64 (yellow spot). This binding brings CAII close enough to the transporter pore to shuttle

1288

protons between transporter and surrounding protonatable residues (orange circles). Proton

1289

shuttling is mediated by CAII-Glu69 and CAII-Asp72 (red and green dots). Under hypoxic conditions

1290

CAIX (green circle) binds to MCT1/4 via their chaperon CD147 (ochre structure) to facilitate the

53

1291

exchange of protons between transporter and extracellular protonatable residues (orange circles) via

1292

its proteoglycan-like (PG) domain in a similar fashion as Glu69 and Asp72 in CAII.

1293

The necessity for this proton antenna derives from the slow diffusion of H+ within the highly buffered

1294

cytosol. Lactate, which is produced from glucose (entering the cell by facilitated diffusion via glucose

1295

transporters (light blue structure)) by glycolysis and subsequent conversion of pyruvate, quickly

1296

reaches the MCT by simple diffusion. Protons, however, which are produced during hydrolysis of ATP

1297

(cell work), diffuse very slow within the cell. To allow fast extrusion of protons and lactate from the

1298

cell, the MCT does not extract H+ directly from the bulk cytosol, but from an intermediate harvesting

1299

compartment made up by protonatable residues at the cell membrane and CAII. Like in the

1300

cytoplasm, diffusion of ions in the extracellular space is restricted. Therefore protons have to be

1301

removed from the extracellular site of the transporter by CAIX and shuttled to protonatable residues

1302

at the extracellular face of the cell membrane, from where they can be released to the extracellular

1303

space. By this non-catalytic mechanism intracellular and extracellular carbonic anhydrases could

1304

cooperate non-enzymatically to facilitate proton-driven lactate flux across the cell membrane of

1305

cancer cells.

1306 1307 1308

Figure Supplement Legends

1309

Figure 1 – Figure Supplement 1. Determination of CAII knockdown efficiency in MCF-7 cells. (A ΔCt

1310

values for CAII minus RPL27 of normoxic and hypoxic MCF-7 cells, treated with siRNA against CAII

1311

(blue bars) or non-targeting negative control siRNA (gray bars) (mean + SEM). *p≤ .

1312

test. (B) Relative change in the RNA level of CAII as given by the 2-ΔΔCt values for CAII in normoxic and

1313

hypoxic MCF-7 cells treated with siRNA against CAII compared to cells treated with non-targeting

1314

negative control siRNA (mean + SEM). (C ‘eprese tati e

1315

actin (lower panel) from of normoxic and hypoxic MCF-7 cells, treated with siRNA against CAII or

1316

non-targeting negative control siRNA. (B) Relative intensity of the fluorescent signal for CAII,

ester

; “tude t s t-

lot for CAII upper pa el a d β-

54

1317

normalized to the signal intensity of β-actin in the same probe (mean + SEM). The significance

1318

indicators above the bars for siCAII refer to the values of siNeg. ***p≤ .

; “tude t s t-test.

1319 1320

Figure 1 – Figure Supplement 2. Influence of pHi on lactate transport. (A) Initial intracellular pH

1321

(pHi), as measured at the beginning of the experiment, of normoxic (left row) and hypoxic (right row)

1322

MCF-7 cells, treated with siRNA against CAII (blue) or non-targeting negative control siRNA (gray)

1323

ea ± “EM . ***p≤ .

; “tude t s t-test. (B-E ΔpHi/Δt duri g appli atio of

M la tate i

1324

normoxic (B, C) and hypoxic (D, E) MCF-7 cells, either treated with negative control siRNA (B, D) or

1325

CAII-siRNA (C, E), as shown in Figure 1 B, plotted against the cells initial pHi. Every dot represents one

1326

individual cell. The red lines represent a linear regression fit.

1327 1328

Figure 1 – Figure Supplement 3. Calibration of SNARF-5 in MCF-7 cells. (A) Confocal image of MCF-7

1329

cells, loaded with SNARF-5. (A1) Signal of the emission fraction below 590 nm. (A2) Signal of the

1330

emission fraction above 590 nm. (A3) Signal of the emission fraction below 590 nm divided by the

1331

signal of the emission fraction above 590 nm. (B) Calibration of the fluorescence ratio in the

1332

presence of nigericin and 130 mM K+, at pH 8.0, 7.5, 7.0, 6.5 and 6.0. An exponential equation was

1333

used to calculate the maximum steady state for each pH application, as indicated by the red traces.

1334

Gray traces indicate the confidence bands (for 95% confidence level). (C) Fluorescent ratio plotted

1335

against the extracellular pH (mean ± SEM). A Boltzmann fit (red) was used to calculate the

1336

parameters of conversion (inset). Gray traces indicate the confidence bands (for 95% confidence

1337

level). pH values are calculated as pH = x + dx ∗ ln

1338

A −A R−A

−1 .

1339

Figure 4 – Figure Supplement 1. CAII-Glu69 and Asp72 are crucial for the facilitation of MCT4

1340

transport activity. (A) Original recordings of the change in intracellular H+ concentration in oocytes

1341

expressing MCT4 (black trace), MCT4 + CAII-WT (blue trace), MCT4 + CAII-E69Q (red trace), MCT4 +

1342

CAII-D71N (gray trace), and MCT4 + CAII-D72N (green trace), respectively, during application of 3 and 55

1343 1344

10 mM lactate and application of 5% CO2 / 10 mM HCO3-. (B, C) Rate of change in intracellular H+ o e tratio

Δ[H+]/Δt as i du ed

appli atio

B a d re o al C of

a d

M lactate,

1345

respectively, in oocytes expressing MCT4 (dark gray), MCT4 + CAII-WT (blue), MCT4 + CAII-E69Q

1346

(red), MCT4 + CAII-D71N (light gray), and MCT4 + CAII-D72N (green), respectively (mean + SEM). The

1347

black significance indicators refer to MCT4, the blue significant indicators refer to MCT4 + CAII-WT.

1348

*p≤ .

, p≤ .

, ***p≤0.001. n.s. o sig ifi a e; “tude t s t-test.

1349 1350

Figure 4 – Figure Supplement 2. Expression levels of CAII are not altered by single-site mutation.

1351

(A) Representative western blot for CAII upper pa el a d β-tubulin (lower panel) from native

1352

oocytes and oocytes expressing CAII-WT, CAII-E69Q, CAII-D71N, and CAII-D72N, respectively. (B)

1353

Relative intensity of the fluorescent signal for CAII (mean + SEM). The signal intensity of CAII-WT was

1354

set to 100%. The significance indicators refer to the original values of CAII-WT. ***p≤ .

1355

sig ifi a e; “tude t s t-test.

, .s. o

1356 1357

Figure 5 – Figure Supplement 1. Intrinsic buffer capacity of oocytes. (A, B) Intrinsic buffer capacity

1358

βi), as measured by application of 5% CO2 / 10 mM HCO3-, in oocytes expressing MCT1 (A) or MCT4

1359

(B) either alone or together with CAII-WT, CAII-E69Q, CAII-D71N, and CAII-D72N, respectively, in the

1360

absence and presence of 4-MI in the cytosol (mean + SEM). *p≤ .

1361

test.

, .s. o sig ifi a e; “tude t s t-

1362 1363

Figure 5 – Figure Supplement 2. Chemical rescue of the interaction between MCT4 and CAII-

1364

E69Q/D72N by 4-methylimidazole. (A, B) Rate of change in intracellular H+ o e tratio

1365

as induced by application (A) and removal (B) of 3 and 10 mM lactate, respectively, in oocytes

1366

expressing MCT4 (dark gray), MCT1 + CAII-WT (blue), MCT4 + CAII-E69Q (red), MCT4 + CAII-D71N

1367

(light gray), and MCT4 + CAII-D72N (green), respectively, in the presence of 4-MI in the cytosol (mean

1368

+ SEM). The black significance indicators refer to MCT4, the blue significant indicators refer to MCT4

Δ[H+]/Δt

56

1369 1370

+ CAII-WT. (C) Lactate-induced proton flux (JH), as calculated from the rate of change in pHi and the ells i tri si

uffer apa it

βi; Figure 4 – Figure Supplement 1). The significance indicators above

1371

the bars from 4-MI-injected oocytes refer to the cells expressing the same proteins without 4-MI.

1372

*p≤ .

, **p≤ .

, ***p≤ .

. .s. o sig ifi a e; “tude t s t-test.

1373 1374

Figure 8 – Figure Supplement 1. CAII-His64 mediates binding, but no proton transfer between

1375

MCT4 and CAII. (A) Structural model of the binding between the C-terminal tail of MCT4 (raspberry)

1376

and CAII (blue). Binding is mediated by Glu431 (orange) and Glu433 (red) in the MCT4 C-terminal tail

1377

and His64 (yellow) in CAII. (B) Original recordings of the change in intracellular H+ concentration in

1378

oocytes expressing MCT4 (black trace), MCT4 + CAII-WT (blue trace), MCT4 + CAII-H64A (yellow

1379

trace), and MCT4 + CAII-H64K (green trace), respectively, during application of 3 and 10 mM of

1380

lactate and application of 5% CO2 / 10 mM HCO3-. (C, D) Rate of change in intracellular H+

1381

o e tratio

Δ[H+]/Δt as i du ed

appli atio

C a d re o al D of

a d

M la tate,

1382

respectively, in oocytes expressing MCT4 (gray), MCT4 + CAII-WT (blue), MCT4 + CAII-H64A (yellow),

1383

and MCT4 + CAII-H64K (green), respectively. The black significance indicators refer to MCT4, the blue

1384

significant indicators refer to MCT4 + CAII-WT (mean + SEM). (E) Representative western blots for

1385

CAII (upper panel) and GST (lower panel). CAII-WT, CAII-H64A, and CAII-H64K were pulled down with

1386

a GST fusion protein of the C-terminal of MCT4 (GST-MCT4-CT). As negative control CAII-WT was

1387

pulled down with GST alone. (F) Relative intensity of the fluorescent signal for CAII (mean + SEM).

1388

The signal intensity of CAII, pulled down with GST-MCT4-CT was set to 100%. The significance

1389

indicators refer to the original values of CAII-WT. ***p≤ .

, .s. o sig ifi a e; “tude t s t-test.

1390 1391

Figure 8 – Figure Supplement 2. Catalytic activity and expression levels of CAII His64 mutants. (A)

1392

Original recording of the log enrichment (LE) of either a lysate of 20 native oocytes (black trace) or a

1393

lysate of 20 oocytes expressing CAII-WT (blue), CAII-H64A (yellow), or CAII-H64K (green). The

1394

beginning of the traces shows the rate of degradation of the

18

O-labelled substrate in the non57

1395

catalyzed reaction; the black arrow indicates the addition of oocyte lysate. (B) Enzymatic activity of

1396

CA in units/ml (mean + SEM). The black asterisks refer to the values from native oocytes, the blue

1397

significance indicators refer to the values of oocytes expressing CAII-WT, the yellow significance

1398

indicators refer to the values of oocytes expressing CAII-H64A. (C) Representative western blot for

1399

CAII upper pa el a d β-tubulin (lower panel) from native oocytes and oocytes expressing CAII-WT,

1400

CAII-H64A, and CAII-H64K, respectively. (D) Relative intensity of the fluorescent signal for CAII (mean

1401

+ SEM). The signal intensity of CAII-WT was set to 100%. The significance indicators refer to the

1402

original values of CAII-WT. ***p≤ .

, .s. o sig ifi a e; “tude t s t-test.

58

Figure 1

B

A Normoxia pHi 0.2

Hypoxia pHi 0.2

control

control

3 min

3 min

CAII knock-down

Hypoxia n=41/3

**

n=54/5 n=46/3

***

control

**

n=39/3 n.s.

*** n.s.

*** CAII knock-down

D 0.22

*** control

CAII knock-down

0.20 0.18 0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02 0

Efflux

10 mM Lac

*

Normoxia

n=54/5

***

control

Hypoxia

n=41/3 n=39/3

n=46/3

10 mM Lac

Normoxia

3 mM Lac

3 mM Lac

Influx

10 mM Lac

0.20 0.18 0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02 0

3 mM Lac

DpHi/Dt (min-1)

C 0.22

10 mM Lac

DpHi/Dt (min-1)

3 mM Lac

CAII knock-down

***

CAII knock-down

** ***

*** * ***

control CAII knock-down

Figure 1 - Figure Supplement 1

A 6

*

*

B 1.0 n=6/3 0.8

(43 kDa)

D 1.2 1.0

n=6/3

0.8 0.6

0.2 0

***

***

0.4

Normoxia

siCAII

b-Actin

siNeg vs siCAII

siNeg

(30 kDa)

Hypoxia

siCAII

siNeg

Normoxia Hypoxia siNeg siCAII siNeg siCAII

CAII

0.0

Hypoxia

siCAII

C

Normoxia

0.4 0.2

Rel. amount of CAII

0

siCAII

1

siNeg

2

0.6 Normoxia

-DDCt

3

2

DCt

4

siNeg

5

n=6/3

Hypoxia

Figure 1 - Figure Supplement 2

B 0.24

7.8

*** n=54/5

*** 7.6

0.08

7.4

7.5

control CAII knock-down

0.12 0.08

7.6

7.7

7.8

7.9

8.0

0

8.1

7.4

7.5

7.9

8.0

8.1

7.9

8.0

8.1

E 0.24 Normoxia

Hypoxia

0.20

CAII knock-down 0.16

n=46/3

0.12 0.08 0.04 7.4

7.8

initial pHi

C 0.24

0

7.7

7.6

initial pHi

0.20

7.4

n=41/3

0.16

0.04

n=46/3

Normoxia

7.2

0.12

0

Hypoxia

7.5

7.3

0.16

0.04

n=39/3

DpHi/Dt (min-1)

initial pHi

7.7

Hypoxia Control

0.20

n=54/5

DpHi/Dt (min-1)

n=41/3

DpHi/Dt (min-1)

0.20 7.9

D 0.24

Normoxia Control

DpHi/Dt (min-1)

A 8.0

CAII knock-down 0.16

n=39/3

0.12 0.08 0.04

7.5

7.6

7.7

7.8

initial pHi

7.9

8.0

8.1

0

7.4

7.5

7.6

7.7

7.8

initial pHi

Figure 1 - Figure Supplement 3

A1

A2

590nm

30 µM

100 80 60 40

8.0

7.5

30 µM

C 100

10 min

20

7.0

590nm

30 µM

Fluorescent ratio (a.u.)

Fluorescent ratio (a.u.)

B

A3

6.5

6.0 pHe 7.5

Model Equation Reduced Chi-Sqr Adj. R-Square

80

Amplitude

Boltzmann y = A2+(A1-A2)/(1+exp((x-x0)/dx)) 3.78483 0.99943 Value Std.Err. 108.50451 5.02114 A1 14.36576 2.7177 A2 6.84642 0.04354 x0 dx 0.06249 0.53247

60 40

n = 53/4

20 6.0

6.5

7.0 pHe

7.5

8.0

Figure 2

A1

A2

A3

A4

Anti-CAII

Anti-MCT1

Anti-CAII Anti-MCT1 DAPI

no prim Ab DAPI

30 µM

30 µM

30 µM

30 µM

B1

B2

B3

B4

normoxia

hypoxia

CAII-siRNA

no primary antibody

15 µM

15 µM

15 µM

n.s.

Normoxia 15

Hypoxia CAII-siRNA

10

*** n=9/3

0

n=12/4

5

(normoxia)

n=12/4

PLA signals / nucleus

C 20

Ab MCT1 + Ab CAII

n=9/3

*** no prim. Ab

15 µM

Figure 3

Normoxia A2 siNeg

control

Normoxia A3 siCAII

300 µM

300 µM

Hypoxia B2 siNeg

B1 control

number of cells / mm2

C 500

300 µM

400

control siNeg

300

siCAII

200

n = 12/4 n.s.

100

** ***

0 0

EZA AR-C

*

*** *** 1 2 time of incubation (days)

300 µM

Hypoxia B4 EZA

*** *** 3

300 µM

Hypoxia B5 AR-C155858

300 µM

n.s.

Normoxia

Normoxia A5 Normoxia AR-C155858

300 µM

Hypoxia B3 siCAII

300 µM

Normoxia A4 EZA

300 µM

D 500 number of cells / mm2

A1

Hypoxia

300 µM

Hypoxia

n.s.

400

control siNeg

300

siCAII

200

n = 12/4

EZA AR-C

n.s. n.s.

100

*** **

0 0

*** *** 1 2 time of incubation (days)

*** *** 3

Figure 4

B MCT1

A

MCT1 + CAII-WT MCT1 + CAII-E69Q MCT1 + CAII-D71N MCT1 + CAII-D72N

hCAII His64

Zn

+

[H ]i 20 nM 3 min

Glu69 Asp72

Asp71

3 mM Lac

30 0

150

n=10

*** n.s.

n=12

*** 3 mM Lac

60

***

***

120 90

D 180

*** n.s.

n=13

*** n.s.

n.s.

**

*** n.s.

MCT1 +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N

Efflux

60 30 0

***

***

120 90

n=8 n.s.

n=15

n=13

n=10 n=12

*** * ***

10 mM Lac

150

n=8 n.s.

5% CO2/10 mM HCO3

3 mM

n=15

D[H+]i/Dt (nM/min)

Influx

10 mM Lac

D[H+]i/Dt (nM/min)

C 180

10 mM Lac

*** *

*** *

n.s.

**

*** *

MCT1 +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N

-

Figure 4 - Figure Supplement 1

A

MCT4 MCT4 + CAII-WT MCT4 + CAII-E69Q MCT4 + CAII-D71N MCT4 + CAII-D72N

+

[H ]i 20 nM 4 min

n.s.

***

**

120

*** n.s.

90

30 0

3 mM Lac

60

***

*** n.s.

*** n.s.

n.s.

** *** n.s.

MCT4 +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N

150

-

Efflux n=8

n.s.

***

**

120

*** n.s.

90 60 30 0

**

10 mM Lac

n=8

D[H+]i/Dt (nM/min)

150

5% CO2/10 mM HCO3

C 180

Influx

10 mM Lac

D[H+]i/Dt (nM/min)

B 180

10 mM Lactate

3 mM

3 mM Lactate

** n.s.

*** n.s.

n.s.

*

** n.s.

MCT4 +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N

Figure 4 - Figure Supplement 2

A Native

CAIIWT

CAIIE69Q

CAIID71N

CAIID72N

CAII (30 kDa)

b-Tub (50 kDa) n.s.

rel. amounts of CAII (%)

B 125

n.s.

n.s.

n=2/2 100 50 75 25 0

*** Native

CAIIWT

CAIIE69Q

CAIID71N

CAIID72N

Figure 5

B MCT1 + 4-MI

A

MCT1 + CAII-WT + 4-MI MCT1 + CAII-E69Q + 4-MI MCT1 + CAII-D71N + 4-MI MCT1 + CAII-D72N + 4-MI

hCAII 4-MI Ala64

Asn67

Å

3.2Å

Asp71

16 14 12 10 8 6 4 2 0

7.1Å

Leu57

Asp72

3 mM Lac

D 160 n.s.

n.s.

+ 4-MI

n.s.

3 mM Lac

**

***

**

**

**

n.s.

n.s.

**

n.s.

**

**

D[H+]i/Dt (nM/min)

Influx n=8

10 mM Lac

D[H+]i/Dt (nM/min) JH (mM/min)

E

140 120 100 80 60 40 20 0

4-MI

MCT1 +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N Influx n=15

140 120 100 80 60 40 20 0

n=8

n=8

**

n.s.

n=12

n=8

n.s.

n.s.

n=10

**

10 mM Lac

Efflux n=8

n.s.

n.s.

**

+ 4-MI

n.s.

*

**

*

*

n.s.

**

n.s.

*

10 mM Lac

8.6

Å

3.3

+

[H ]i 20 nM 2 min

6.7Å

Å

4.9

3 mM

Å

4-MI

Asp72

C 160

Glu69

8Å Å 3.9

2.

4-MI

3.7

Glu69

6. 6Å

Ile91

Zn

n.s.

*

MCT1 +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N n=8

n=8

n=8

n.s.

* n=12

n.s.

**

n.s.

MCT1 +4-MI

+CAII- +4-MI WT

+CAII- +4-MI E69Q

+CAII- +4-MI D71N

+CAII- +4-MI D72N

Figure 5 - Figure Supplement 1

4-MI

control

control

4-MI

control

n.s.

n.s.

n.s.

4-MI

n.s.

*

4-MI

4-MI

control

4-MI

control

4-MI

control

4-MI

control

MCT1 +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N

n=8

control

n.s.

bi (mM)

*

n.s.

4-MI

*

n.s.

20 18 16 14 12 10 8 6 4 2 0

control

B

n=8-15

4-MI

18 16 14 12 10 8 6 4 2 0

control

bi (mM)

A

MCT4 +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N

Figure 5 - Figure Supplement 2

JH (mM/min)

n.s.

**

**

n.s.

**

Influx

12

n=8

90 60 30 0

n.s.

** n.s.

**

**

n.s.

** n.s.

**

** n.s.

**

MCT4 +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N

n.s.

10

n.s.

**

+ 4-MI

120

MCT4 +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N

14

**

n.s.

*

***

n.s.

8 6

n.s.

*

n=8

150

10 mM Lac

3 mM

30

n.s.

**

**

**

60

n.s.

n.s.

+ 4-MI

90

0

C

**

Efflux

3 mM

120

n=8

D[H+]i/Dt (nM/min)

150

B 180

Influx

10 mM Lac

D[H+]i/Dt (nM/min)

A 180

n.s.

*

n.s.

n.s.

*

4 2 0

MCT4 +4-MI

+CAII- +4-MI WT

+CAII- +4-MI E69Q

+CAII- +4-MI D71N

+CAII- +4-MI D72N

Figure 6

A

GST +CAIIWT

GST-MCT1-CT +CAII- +CAIIWT E69Q

C +CAIID71N

GST +CAIIWT

+CAIID72N

CAII

CAII

(30 kDa)

(30 kDa)

GST

GST

(30 kDa)

(30 kDa)

125

D 150 n=4/3

n.s. n.s. n.s.

100 75 50 25 0

*** GST +CAIIWT

GST-MCT1-CT +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N

rel. amounts of CAII (%)

rel. amounts of CAII (%)

B 1 50

GST-MCT4-CT +CAII- +CAIIWT E69Q

125

n=4/3

n.s.

+CAIID71N

n.s.

+CAIID72N

n.s.

100 75 50 25 0

*** GST +CAIIWT

GST-MCT´4-CT +CAII- +CAII- +CAII- +CAIIWT E69Q D71N D72N

Figure 7

D[H+]i/Dt (nM/min)

C 250 200 150

Influx n=8

***

n.s.

n.s.

n.s.

n.s.

***

***

***

** **

50 Native CAII- CAII- CAII- CAII- CAII- CAIIWT H64A E69Q D71N D72N E69Q D72N

n.s.

n.s.

***

***

n.s.

***

20

D 250 200

n=7

n=6

n=6

n=9

*** ***

10 n=6

100

0

***

30

0

***

***

n=7

CA activity (U/mL)

Native CAII-WT CAII-H64A LE CAII-E69Q 0.1 CAII-D71N 1 min CAII-D72N CAII-E69Q/D72N

n.s.

40

n=11

B

oocyte lysate

D[H+]i/Dt (nM/min)

A

Native CAII- CAII- CAII- CAII- CAII- CAIIWT H64A E69Q D71N D72N E69Q D72N Efflux n=8

***

n.s.

n.s.

n.s.

n.s.

***

***

***

***

150 100

** **

50 0

Native CAII- CAII- CAII- CAII- CAII- CAIIWT H64A E69Q D71N D72N E69Q D72N

Figure 8

A

B MCT1

MCT1-CT

MCT1 + CAII-WT MCT1 + CAII-H64A MCT1 + CAII-H64K

CT

Glu489

Glu491 hCAII His64 +

[H ]i 25 nM 2 min

Zn

3 mM Lac

**

120

** n.s. **

60 30 0

E

MCT1

GST +CAIIWT CAII

(30 kDa)

GST (30 kDa)

10 mM Lac

90

+CAIIWT

** n.s.

+CAIIH64A

GST-MCT1-CT +CAII+CAIIWT H64A

n.s.

***

F

n=8

**

n.s.

**

120

** n.s.

90

**

60 30 0

+CAIIH64K

+CAIIH64K

150

Efflux

MCT1

+CAIIWT

+CAIIH64A

1 50 125

n=4/2

n.s.

100 75 50 25 0

*** GST +CAIIWT

n.s.

**

** n.s.

10 mM Lac

**

5% CO2/10 mM HCO3

3 mM Lac

n=8

D[H+]i/Dt (nM/min)

n.s.

rel. amounts of CAII (%)

150

D 180

Influx

3 mM Lac

D[H+]i/Dt (nM/min)

C 180

10 mM Lac

*** GST-MCT1-CT +CAII- +CAII- +CAIIWT H64A H64K

+CAIIH64K

-

Figure 8 - Figure Supplement 1

A

B MCT4

MCT4-CT

MCT4 + CAII-WT MCT4 + CAII-H64A MCT4 + CAII-H64K

CT

hCAII

Glu431 His64 Glu433

+

[H ]i 25 nM 2 min

Zn

3 mM Lac

120

** n.s. ***

60 30 0

E

MCT4

GST +CAIIWT CAII

(30 kDa)

GST (30 kDa)

10 mM Lac

90

+CAIIWT

** n.s.

+CAIIH64A

GST-MCT4-CT +CAII+CAIIWT H64A

n.s.

***

F

n=8

n.s.

***

**

120

*n.s. **

90

***

60 30 0

+CAIIH64K

+CAIIH64K

150

Efflux

MCT4

10 mM Lac

***

5% CO2/10 mM HCO3

3 mM Lac

***

D[H+]i/Dt (nM/min)

n=8

n.s.

rel. amounts of CAII (%)

150

D 180

Influx

3 mM Lac

D[H+]i/Dt (nM/min)

C 180

10 mM Lac

+CAIIWT

1 50 125

** n.s.

**

+CAIIH64A

+CAIIH64K

n.s.

n=4/2

100 75 50 25 0

*** GST +CAIIWT

n.s.

*** GST-MCT4-CT +CAII- +CAII- +CAIIWT H64A H64K

-

Figure 8 - Figure Supplement 2

A

C

oocyte lysate

Native

CAIIWT

CAIIH64A

CAIIH64K

CAII (30 kDa)

Native CAII-WT CAII-H64A CAII-H64K

b-Tub (50 kDa)

1 min

D

50

CA activity (U/ml)

n=8

40

***

n=7

30 20

n=7

10

*** ***

0

*** *** ***

n=6

Native

CAIIWT

CAIIH64A

CAIIH64K

150

rel. amounts of CAII (%)

B

LE 0.2

125

n=4/3

n.s.

n.s.

CAIIH64A

CAIIH64K

100 75 50 25

***

0 Native

CAIIWT

Figure 9

His3

His15

His64

B 160 140 120 100 80 60 40 20 0

n.s.

n.s.

n.s.

***

***

***

***

n.s.

n.s.

n.s.

***

***

n.s.

***

*** n.s.

*** *** 3 mM Lac

His4

His10 Lys9

10 mM Lactate

hCAII

D[H+]/Dt (nM/min)

A

MCT1+ CAII WT n=18 n=28

CAII H3A n=16

n.s.

***

CAII H4A n=15

n.s.

***

n.s.

***

n.s.

n.s.

***

***

n.s.

***

CAII CAII CAII CAII CAII H10A H15A H3/4A H3/4/10A K9A n=12 n=8 n=9 n=6 n=12

Figure 10

Lac

-

H

Gluc

+

H

+

MCT1/4

GLUT intracellular

H64 E69 D72

Gluc

H glycolysis

H

CAIX

CD 147

extracellular

+ PG

Pyr

-

Lac LDH

-

H

+

H

+

+

CAII

cell work