1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Abstract 16 LRRK2 mutations

0 downloads 0 Views 20MB Size Report
12. 6Dept. Medical Genetics, University of British Columbia, Vancouver, Canada. 13. 14. 15. Abstract. 16. LRRK2 mutations produce end-stage Parkinson's ...
1 2 3

Initial elevations in glutamate and dopamine neurotransmission decline with age, as does exploratory behavior, in LRRK2 G2019S knock-in mice

4

Mattia Volta2*, Dayne A. Beccano-Kelly3*, Sarah A. Paschall4*, Stefano Cataldi4+, Sarah MacIsaac4+,

5

Naila Kuhlmann4, Chelsie A. Kadgien4, Igor Tatarnikov4, Jesse Fox4, Jaskaran Khinda4, Emma Mitchell4,

6

Sabrina Bergeron4, Heather Melrose5, Matthew J. Farrer6$ & Austen J. Milnerwood1$

7

+ $

8

1

Dept. Neurology & Neurosurgery, Montreal Neurological Institute, McGill University, Montreal, Canada

9

2

Institute for Biomedicine, EURAC, Bolzano, Italy

10

3

Department of Physiology, Anatomy & Genetics, University of Oxford, Oxford, U.K.

11

4

Graduate Program in Neurosciences, University of British Columbia, Vancouver, Canada

12

5

Mayo Clinic, Jacksonville, Florida, U.S.A.

13

6

Dept. Medical Genetics, University of British Columbia, Vancouver, Canada

* equal contributions

14 15 16

Abstract

17

LRRK2 mutations produce end-stage Parkinson’s disease (PD) with reduced nigrostriatal

18

dopamine. Conversely, asymptomatic carriers have increased dopamine turnover and altered

19

brain connectivity. LRRK2 pathophysiology remains unclear, but reduced dopamine and

20

mitochondrial abnormalities occur in aged mutant knock-in (GKI) mice. Conversely, cultured

21

GKI neurons exhibit increased synaptic transmission. We assessed behavior and synaptic

22

glutamate and dopamine function across ages. Young GKI exhibit more vertical exploration,

23

elevated glutamate and dopamine transmission, and aberrant D2-receptor responses. These

24

phenomena decline with age, but are stable in littermates. In young GKI, dopamine transients are

25

slower, independent of DAT, increasing dopamine extracellular lifetime. Slowing of dopamine

26

transients is observed with age in littermates, suggesting premature ageing of dopamine synapses

27

in GKI. Thus, GKI mice exhibit early, but declining, synaptic and behavioral phenotypes,

28

making them amenable to investigation of early pathophysiological, and later parkinsonian-like,

29

alterations. This model will prove valuable in efforts to develop neuroprotection for PD.

30 31 32

1

33

Introduction

34

Parkinson’s disease (PD) is clinically diagnosed when patients present with characteristic

35

progressive motor symptoms, although post-mortem detection of Lewy pathology and nigral cell

36

loss are currently required for confirmation. A recent study suggests nigral cell death may be as

37

low as 0-10% 1-3 years from diagnosis, whereas dopamine functional markers such as tyrosine

38

hydroxylase (TH) and dopamine transporter (DAT) are profoundly reduced at the earliest points

39

assessed 1. The rapid and near complete loss of dopamine functional markers at (or within a few

40

years of) diagnosis argues that ongoing clinical deterioration over several years is due to loss of

41

compensatory mechanisms and / or dysfunction of non-dopaminergic neurons.

42

Although motor symptoms respond well to current therapy (e.g., dopamine replacement

43

by L-DOPA or deep brain stimulation; DBS), PD is a multisystem disorder with a host of L-

44

DOPA unresponsive features. All patients suffer a range of non-motor symptoms, many of which

45

appear to precede motor onset by years or decades 2–8. Cognitive dysfunction is seen in ~40% of

46

newly diagnosed PD cases 9–11 in the form of deficits in attention, executive function, verbal

47

fluency and visuospatial processing, rather than memory per se (although memory is often also

48

impaired) 6,12. This dysexecutive / subcortical syndrome is thought to be due to impaired cortico-

49

striatal basal ganglia processing for action selection 2,13,14. There are no effective symptomatic

50

treatments for many of these non-motor issues, nor are there currently any disease-modifying

51

(neuroprotective) interventions.

52

While the etiology for most Parkinson patients remains unknown, asides age, gene

53

mutations contribute most risk 15. Pathogenic mutations in leucine-rich repeat kinase 2 (LRRK2)

54

account for ~1% of all PD, of which LRRK2 G2019S is most frequent; identified in ~30% of

55

cases in some ethnicities 16,17. Affected LRRK2 individuals develop late-onset, L-DOPA-

56

responsive motor parkinsonism that is clinically and often pathologically indistinguishable from

57

idiopathic PD 18,19. Dopamine PET imaging of affected LRRK2 mutation carriers reveals

58

progressive neurochemical alterations similar to those of sporadic PD, namely impaired

59

presynaptic dopamine function 20,21. Further study in LRRK2 families reveals surprising

60

alterations in clinically asymptomatic, otherwise healthy, mutation carriers, including: i) early

61

increases in dopamine turnover by PET 22, ii) changes in cortical connectivity by resting state

62

MRI and neurochemical changes 20,23,24, and iii) alterations in cognitive tests of executive

63

function 25. Advances in our understanding of PD argue cell death and overt motor dysfunction

2

64

are late occurrences, preceded by dysfunction of dopaminergic and non-dopaminergic systems.

65

In this light, in model systems modelling PD etiology, the underlying pathophysiology and

66

phenotypes should also be expected to be initially subtle, progressive, and include dysfunction of

67

multiple neuronal systems, prior to cell loss.

68

We engineered the LRRK2 G2019S substitution into the endogenous mouse gene

69

(G2019S knock-in mice; GKI) which in vivo produced reductions in basal and pharmacologically

70

evoked nigrostriatal dopamine release in mice aged >12 months by microdialysis 26. This

71

Parkinson’s-like deficit was not observed at 6 months, and occurred despite a normal

72

compliment of nigral neurons and nigrostriatal dopamine markers (TH). Contrastingly, in

73

cortical neurons cultured from the same GKI mice, we observed increases in glutamatergic and

74

GABAergic synaptic transmission at 21 days in vitro 27.

75

To investigate this disparity, we probed dopamine and glutamate release, neuronal

76

morphology, synaptic proteins and behavior in young and aged GKI mice. Young animals

77

exhibit increased exploratory rearing behavior and increases in striatal glutamate and dopamine

78

transmission. As GKI mice age, they exhibit less exploratory rearing and reductions in both

79

glutamate and dopamine transmission. However, extracellular lifetime of single dopamine

80

release events is enhanced in young GKI animals, and maintained in aged animals, at which

81

point wild-type littermate values have increased to parity. Several measures demonstrate the

82

LRRK2 G2019S mutation produces alterations in young adult mice, most of which decline with

83

age, prior to ages where hypodopaminergia, mitochondrial and tau pathology are observed. We

84

provide further evidence that GKI mice provide a valuable model in which to probe early

85

pathophysiological effects of mutant LRRK2, and later classically PD-like deficits. We conclude

86

that understanding the early pathophysiological changes in etiological models may offer the best

87

hope for development of neuroprotection in PD and related diseases.

88 89

Materials & Methods

90

G2019S knock-in mice and behavioral testing

91

C57Bl/6J wild-type (WT) and Lrrk2 G2019S knock-in heterozygous (GKI) mice 26,27 were

92

maintained according to Canadian Council on Animal Care regulations. To avoid confounds of

93

oestrus cycle upon behavior and neural connectivity, only male animals were used in this work.

94

Mice undergoing surgery were weighed at age of use, and all other mice in the colony were

3

95

weighed at a single time point to produce an age vs. weight plot. Separate cohorts of adult

96

animals were tested (once only) at 1-6 and 12-18 months of age. After familiarization to handling

97

over three days with the operator, mice underwent the following behavioral paradigms and

98

videos were analyzed post-hoc using ANY-maze (Stoelting) behavioral tracking software, as

99

previously 28,29. Open field (OF) test: mice explored an arena (48cm x 48cm) for 15 min.

100

Cylinder test: mice were placed in a 1l beaker and video recorded for 5min. The number of

101

rearings and forelimb wall contacts were scored manually off-line. All testing and analysis was

102

performed experimenter blind.

103 104

Electrophysiology

105

Whole-cell patch clamp recording was conducted in striatal projection neurons (SPNs) in 300μm

106

thick coronal slices from 1-18 month-old male mice, as in 30,31. To help preserve cell viability,

107

slices from >12 month old animals were pre-incubated in recovery solution containing (in mM:

108

93 NMDG, 93 HCl, 2.5 KCl, 1.2 NaH2PO4, 30 NaHCO3, 20 HEPES, 25 Glucose, 5 Sodium

109

Ascorbate, 2 Thiourea, 3 Sodium Pyruvate, 10 MgSO4.7H2O, 0.5 CaCl2.2H2O, pH 7.3-7.4,

110

300-310 mOsm) for 15min at 34˚C prior to transfer to a holding chamber. Slices were held and

111

perfused at RT with artificial cerebrospinal fluid (ACSF) containing (in mM): 125 NaCl, 2.5

112

KCl, 25 NaHCO3, 1.25 NaH2PO4, 2 MgCl2, 2 CaCl2, 10 glucose, pH 7.3–7.4, 300–310 mOsm).

113

Cells were visualized by IR-DIC on an Olympus BX51 microscope (20x + 4x magnifier) and

114

SPNs visually identified by somatic size (8–20 μm), morphology and location within the

115

dorsolateral striatum, 50–150 μm beneath the slice surface. Data were acquired by Multiclamp

116

700B amplifier digitized at 10 kHz, filtered at 2 kHz and analyzed in Clampfit10 (Molecular

117

Devices). Pipette resistance (Rp) was 5–8 MΩ when filled with (in mM): 130 Cs

118

methanesulfonate, 5 CsCl, 4 NaCl, 1 MgCl2, 5 EGTA, 10 HEPES, 5 QX-314, 0.5 GTP, 10 Na2-

119

phosphocreatine, and 5 MgATP, 0.1 spermine, pH 7.2, 290 mOsm. Tolerance for series

120

resistance (Rs) was 22 months 26).

240

Longo et al., (2014) found homozygous GKI mice exhibited hyperactivity at all ages

241

tested 35; although such a result might be impacted by lower body weight in mutants. In

242

agreement, we previously observed an increase (~10%) in open field activity in a small cohort of

243

homozygous GKI mice at 6 (but not 12) months in the founding colony 26; whereas hyperactivity

244

was not observed at any age in heterozygous mice 26. Here we tested large cohorts in the open

245

field exploration task and, in agreement with our previous report, found no significant effects

246

(Figure 1.B.i, WT n=66 & 34, GKI n=72 & 44, for 12 months, respectively; 2-way

247

ANOVA: F(1,212)=0.09, 0.002, 0.007 and p=0.8, 0.9 & 0.9 for age, genotype and interaction,

248

respectively). We similarly found no evidence for thigmotaxis at either age point, suggesting a

8

249

lack of anxiety- or anxiolytic-like phenotypes, which may have altered open field exploration

250

(center path ratios were not significantly different).

251

Contrastingly, in the cylinder exploration test conducted in a smaller environment that

252

may stimulate mice or relieve some stress of open field testing, we found significant age and

253

age-genotype interaction effects (Figure 1.B.ii, WT n=50 & 33, GKI n=59 & 49, for 12

254

months, respectively; 2-way ANOVA: F(1,187)=23.6, 7.3, and p=0.0001, 0.008 for age and

255

genotype-age interaction, respectively). Post-test analysis demonstrated a significant ~25%

256

increase in rearing events in GKI mice aged 35% increase in miniature excitatory postsynaptic current (mEPSC) frequency at 3 weeks

271

in vitro 27. To determine whether similar alterations to cortical/thalamic glutamate release occur

272

in GKI mouse brain, we conducted whole-cell patch-clamp recording of dorsolateral striatal

273

medium-sized spiny projection neurons (MSNs or SPNs; referred to herein as SPNs) in slices

274

acutely prepared from young and aged GKI and WT littermate mice (Figure 2). Intrinsic

275

membrane properties were as predicted for SPNs 30,33,38, and although there were statistically

276

significant age effects upon membrane capacitance (Cm), membrane resistance (Rm) and decay

277

time constants (Tau m), there were no genotype or genotype-age interaction effects (see Figure

278

2.Supplement 1; 2-way ANOVA values included). Thus, intrinsic membrane properties of SPNs

279

are not altered by the presence of G2019S LRRK2.

9

280

Analysis of spontaneous EPSCs (sEPSCs, Figure 2.A) revealed a significant main age

281

effect upon event amplitude (Figure 2.A.i. 2-way ANOVA; F(1,165)=3.92, p=0.049, WT n=40(17)

282

& 31(12), GKI n=53(20) & 44(15), for 1-3 and >12 months, respectively), but no genotype or

283

genotype-age interaction effects (F(1,165)=2.1, 1.8, p=0.15 & 0.18, respectively). Contrastingly,

284

analysis of sEPSC frequencies revealed significant main genotype and age effects (Figure 2.A.ii.

285

2-way ANOVA; F(1,165)=6.3, 17.3, p=0.013, 0.0001, respectively), and a significant interaction

286

between genotype and age (F(1,165)=7.8, p=0.006). Post-tests demonstrated a significant ~47%

287

increase in sEPSC frequency in GKI SPNs aged 1-3m, relative to WT littermates (p=0.0004).

288

This phenomenon was significantly reduced with age in GKI mice (p=0.0001), but not WT

289

animals (p=0.6). To more discretely interrogate the age-dependency of this large increase in

290

sEPSC event frequency, comparisons were made between GKI and WT data sets in approximate

291

1, 3, 12 and 18 month groupings (Figure 1. Supplement 2, RM-ANOVA values included); inter-

292

event interval cumulative probabilities demonstrated that frequency is higher in GKI SPNs at 1

293

and 3, but not 12 or 18 months of age.

294

A recent report by Matikainen-Ankney et al., (2016) also found a similar increase in SPN

295

event frequency in slices from similar GKI mice (aged 12 months, no such genotype interaction was

414

observed (Figure 3.B.i). The data demonstrate that early hypersensitivity of dopamine release to

415

chronic D2 activation in GKI slices is not present in slices from older animals.

416

Striatal dopamine transmission is tightly regulated by dopamine transporter (DAT)-

417

mediated reuptake; DAT levels, location and activity dictate the sphere of influence and,

418

dominating over diffusion, are the major determinant of released dopamine’s extracellular

419

lifetime in the striatum53–55. There was a significant main effect of genotype (Figure 3.B.ii. No

420

Drug; 2-way ANOVA; F(1,81)=9.73, p=0.003) upon single response decay times (Tau); decays

421

were significantly slower in slices from GKI mice aged 3m than WT controls (post-test p=0.01),

422

but similar in older slices. Slower DA transients might be expected to result from reduced DAT

423

levels or activity, but we found no genotype, age or interaction effects upon DAT (or TH)

424

staining (Figure 3. Supplement 1.A. ANOVA values included). To further probe DAT we also

425

assessed protein levels biochemically and found a significant main effect of genotype, due to

426

increased DAT protein (relative to GAPDH) in GKI brains (Figure 3. Supplement 1.B. ANOVA

427

values included).

428

Although increased DAT protein suggests slower transients are not due to a paucity of

429

DAT, it is possible that slowing of DA transients in GKI is due to a functional impairment in

430

DAT clearance after release. Therefore, in a separate set of experiments we evoked release in a

431

high concentration of the DAT blocker GBR12909 (1uM, IC50=6.63nM; Figure 3.B.ii). In the

432

absence of DAT re-uptake, as expected, WT decays were approximately twice as long (No Drug

433

442.1±38.9ms, +GBR12909 788.2±ms); however, there remained significant main genotype and

434

age effects (Figure 3.B.ii. +GBR12909; 2-way ANOVA; F(1,97)=8.55 & 15.35, p=0.004 &

14

435

0.0002, respectively). Consistent with results in the presence of DAT reuptake, when DAT was

436

blocked decays were significantly longer in slices from GKI mice than WT controls aged 3m

437

(post-test p=0.014). Decays in WT animals slowed with age (post-test p=0.002) to an extent

438

similar to both young old GKI decay times.

439

The data demonstrate augmented dopamine transmission in young GKI mice revealed by

440

repeated stimuli and extracellular lifetime of individual pulses, which is not due to a deficit in

441

DAT-dependent clearance. Moreover, an age-dependent increase in stimulated striatal dopamine

442

signalling (increased extracellular lifetime) appears to occur in GKI mice at an earlier time point.

443

As stimulated release is not impaired in older GKI slices, we conclude that the latent reduction in

444

extracellular dopamine levels observed at 12 (but not 6) months by microdialysis 26 is a result of

445

changes to the endogenous regulation of nigrostriatal dopamine release, rather than impaired

446

vesicular release per se. Interestingly, in agreement with our data here, another recent report in

447

similar knock-in mice by Longo and colleagues 56 found that DAT levels and activity are

448

increased in GKI brains from animals aged 12 (but not 3) months. In this light, synaptic DA

449

release seen here may appear similar to WT in aged GKI slices (or even reduced by

450

microdialysis 26) due to increased clearance through DAT, masking a persisting increase in DA

451

release.

452 453

Summary & Conclusions

454

Familial and idiopathic PD are now accepted as complex motor and non-motor syndromes

455

resulting from dysfunction to multiple neurotransmitter systems and cell types throughout the

456

peripheral and central nervous system 15,57. Although some treatments are effective for motor

457

dysfunction, none slow or prevent the neurodegenerative process. Many non-motor symptoms

458

antedate motor manifestations (and clinical diagnosis) by years to decades; demonstrating

459

protracted pathological processes. In this light, investigations of PD etiology in model systems,

460

especially short-lived rodents, should expect initially subtle, potentially progressive, dysfunction

461

of multiple neuronal systems, prior to (or without) overt motor dysfunction and cell loss 15,57.

462

GKI mice harbor a single point mutation, which confers the greatest genetic risk for PD

463

in humans. We find that spontaneous exploration is significantly elevated in certain tasks in

464

young mice, in broad agreement with work by other labs in similar animals 35. This exploratory

465

hyperactivity is seemingly context dependent, observed in cylinder exploration but not evident in

15

466

the open field; as such we predict this reflects changes in motivation, rather than motor function

467

per se. We expect ongoing investigations of these animals in more complex environments and

468

tasks will uncover other phenotypes. We previously observed that presynaptic glutamatergic

469

release is elevated in cortical cultures from GKI mice 27, suggesting altered synapse development

470

and / or mature function. Here we find that similar increases are observed in glutamate release

471

onto striatal neurons in brain slices from young GKI animals, without changes in synapse

472

number. This increase is not preserved as animals age, with GKI mice exhibiting a pronounced

473

down-regulation of spontaneous glutamate release. A recent independent study of similar GKI

474

animals drew similar conclusions 39, and the weight of evidence demonstrates early alterations to

475

synaptic function produced by endogenous G2019S mutations.

476

Our previous study revealed reduced extracellular dopamine levels in the striata of aged

477

(12 months), but not young (6 month) GKI mice 26. Our analyses of dopamine transmission here

478

demonstrates that reductions in vivo are unlikely to be due to impaired release or DAT reuptake,

479

and are likely a consequence of altered regulation of dopamine release and/or nigral burst firing

480

patterns in vivo. In young GKI mouse slices, dopamine release was elevated upon repeated

481

stimulation, and the extracellular lifetime of dopamine was consistently augmented. This

482

suggests that, similar to glutamate synapses, dopamine transmission is elevated in young animals

483

by the G2019S mutation, an elevation which is lost with ageing. Recent brain imaging studies

484

demonstrate that clinically manifest LRRK2 mutation carriers develop deterioration of the

485

dopamine and serotonin systems, similarly to sporadic PD patients 22,58; however, non-manifest

486

mutation carriers exhibit increased striatal dopamine turnover 22 and early increases in serotonin

487

transporter binding 58. We conclude that the GKI mouse is a valuable model in which to probe

488

the etiology and early pathophysiology of LRRK2, and potentially sporadic, parkinsonism.

489

Interventions against such pathophysiological processes in these models may provide the

490

functional neuroprotection, so desperately lacking for Parkinson’s disease and related disorders.

16

491

492 493 494 495 496 497 498 499 500 501 502 503 504 505

Figures

Figure 1. Increased exploratory behavior in young GKI mice declines with age. A) Animal body weight over time. There was no main effect of genotype upon mouse weight, but there was a significant interaction between weight and age (see text and Figure 1 supplement 1). Weights were not significantly at any age up to 500 days (i; animal n shown in ii); however, in the oldest age-group there here was a statistically significant increase in GKI mean weight (post-test *p=0.018). B) Spontaneous exploratory behavior in young and old mice. Analysis of open field exploration showed no genotype (p=0.9) or interaction (p=0.9) effects on total distance travelled (i). In the cylinder test for vertical exploration (ii) there was a significant interaction between genotype and age upon rearing events (p=0.008). At 12m old mice (ii). Upon repeated stimulation there were significant (post-test *p=0.05, **p=0.01), elevations in dopamine peak release in GKI mice, relative to WT controls at 3 months (ii). In WT and GKI slices from >12 month mice, peak dopamine release was more variable, there a was no modulation induced by repeated stimulation, as in younger slices, nor were there genotype differences. There was a significant age-dependent increase in D2 autoreceptor mediated paired-pulse depression (iii; 3m n=20(5) & 13(5), >12m n=18(6) & 30(11) for WT and GKI, respectively, post-test *** p=0.001), reflected by reduced paired-pule ratios (PPR), but there was no main effect of genotype (p=0.7) or genotype-age interaction (p=0.99). B) The D2 agonist quinpirole (10uM) equally suppressed peak evoked dopamine release in WT and GKI slices from 3m mice after a 10-minute wash-in; however, continued exposure

20

548 549 550 551 552 553 554 555 556 557 558 559 560

revealed a trend toward a main effect of genotype (p=0.07) on repeated release, and a significant genotype-time interaction (p=0.039, post-test *p=0.049 at 18min post application). This effect was not observed in >12 month slices (i). Evoked dopamine transient decay time revealed a significant main genotype effect (ii. No drug, 3m n=20(5) & 13(5), >12m n=19(6) & 33(11) for WT and GKI, respectively p=0.0025) due to significantly longer decay times (slower transients) in 3m GKI slices (post-test *p=0.027). In >12m slices, WT decay times had increased to a value similar to 3 month GKI (ii. No drug). The presence of the presynaptic dopamine transporter (DAT) blocker GBR (10uM, ii. +GBR12909) increased all transient decay times, as expected. There were significant main effects of age and genotype (3m n =17(6) & 23(8), >12m n=31(9) & 30(11), p=0.0002 & 0.0043 for WT and GKI, respectively) upon transient lifetime; in the absence of dopamine re-uptake, 3 GKI transients were still significantly slower than WT controls (post-test *p=0.014), and significant age-dependent increases in dopamine decay times were observed in WT slices (post-test *p=0.0024).

21

561 562 563 564 565 566 567 568 569

Figure 1. Supplement 1. Weights of WT, GKI and GKI homozygous animals are similar until advanced ages. Animal body weight over time. There was no main effect of genotype upon mouse weight, but there was a significant interaction between weight and age (2-way ANOVA, genotype p=0.17, interaction p=0.049). Weights were not significantly at any age up to 500 days; however, in the oldest age-group there was a statistically significant increase in GKI (Holm-Sidak post-test *p=0.018) and GKI Homo (post-test ##p=0.009) mean weight, relative to WT animals.

22

570 571 572 573 574 575 576 577

Figure 2. Supplement 1. There are no genotype-dependent alterations to SPN intrinsic membrane properties. Whole-cell recording of medium-sized spiny striatal projection neurons (SPNs) in membrane test mode and analysis of intrinsic membrane properties. While significant age-dependent changes in membrane capacitance (Cm), resistance (Rm) and decay time (Taum) were observed, there were no genotype or interaction effects. 2-way ANOVA detailed below.

23

578 579 580 581 582 583 584 585 586 587 588 589

Figure 2. Supplement 2. GKI SPN sEPSC frequency is similarly increased in 1 and 3 month slices, and similar to WT in 12 and 18 month slices. Whole-cell recording in dorsolateral striatal SPNs and analysis of sEPSCs recorded at -70mV. Higher sEPSC frequencies (as reflected by decreased inter-event intervals) underlie significant main genotype effects to similar extents in slices from 1 and 3 month animals, but there were no significant effects of genotype in 12 or 18 month slices. Details of 2-way RM-ANOVA and Holm-Sidak post-tests are shown below.

24

590 591 592 593 594 595 596 597 598 599 600 601 602

Figure 2. Supplement 3. Spontaneous EPSCs recorded in dorsolateral striatum in the coronal slice preparation are TTX insensitive. A.i) Example traces of spontaneous EPSCs (sEPSCs) in whole-cell voltage-clamp (WCVC) recordings in striatal projection neurons (SPN) of the dorsolateral striatum, prior to, and in the presence of, tetrodotoxin (TTX; 1uM) rendering any remaining responses miniature EPSCs (mEPSCs). ii) There was no significant effect of TTX upon mean sEPSC amplitude (paired t-test p=0.82) or frequency (paired t-test p=0.29) within each cell. Expressed as a percentage of the initial sEPSC mean value, neither mEPSC amplitudes nor frequencies were significantly altered by TTX application (onesample t-test against 100, p=0.7 & 0.2 respectively). The weak trend in frequency was towards increases in GKI, rather than the decreases predicted if action potential-dependent events were contributing to mean sEPSC frequency.

25

603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618

Figure 2. Supplement 4. There are no differences in the number of postsynaptic specializations, or presynaptic excitatory nerve terminal markers in the striatum of young or aged GKI mice. A) Representative bright-field images of Golgi-impregnated medium-sized spiny projection neurons (SPNs) of the dorsolateral striatum from WT and GKI mice aged 1-3m (i), with inserts showing zoom of areas marked in main image by dashed lines. There were significant increases in the density of postsynaptic protrusions as animals aged between 1-3 and >12 months, but there were no significant differences between WT and GKI dendritic protrusions at either age (WT n=57(9) & 18(3), GKI n=56(9) & 16(3), for 1-3 & >12, respectively). B) Representative confocal micrograph of VGluT1 and VGluT2 immunofluorescence staining in 2 month-old WT dorsolateral striatum, demonstrating the expected pattern of a higher density of VGluT1 than VGluT2 glutamatergic terminals and very little overlap between VGluT1 and VGluT2 (i). There were no significant effects of age or genotype upon VGluT1 or VGluT2 cluster densities (ii; WT n=8(4) & 9(4), GKI n=8(4) & 7(4), for 2m & >12m, respectively). Details of 2-way ANOVA and Holm-Sidak post-tests are shown below.

26

619 620 621 622 623 624 625 626 627 628 629

630 631 632 633 634 635 636 637

Figure 2. Supplement 5. Paired-pulse facilitation profiles in the striatum of aged GKI mice are similar & GKI mutant PPRs are insensitive to dopamine agonism and antagonism. Whole-cell patch clamp recording and eEPSC paired-pulse experiments. i) There were no significant differences in glutamatergic paired-pulse ratio at >12 months, similarly to the 1-3m age point in the absence of quinpirole (Fig.2.C). At 1m of age, when WT and GKI PPRs differ significantly in the presence of quinpirole (Fig.2.C), GKI PPRs were insensitive to both dopamine agonism (ii. 10uM quinpirole) and antagonism (iii. 10uM remoxipride). Details of 2-way ANOVA and Holm-Sidak post-tests are shown below.

Figure 2. Supplement 6. Spontaneous EPSCs recorded in GKI dorsolateral striatum SNPs are remoxipride insensitive. Spontaneous EPSCs (sEPSCs) were recorded prior to, and in the presence of, remoxipride (Remox; 10uM). A) There was no significant effect of remoxipride upon mean sEPSC amplitude (paired t-test p=0.86) or frequency (paired t-test p=0.80) within each cell. Expressed as a percentage of the initial sEPSC mean value, neither amplitude nor frequency was significantly altered by remoxipride application (one-sample t-test against 100, p=0.84 & 0.74, respectively).

27

638 639 640 641 642 643 644 645 646

Figure 2. Supplement 7. There were no differences in GKI sEPSC frequency between D1- and D2dopamine receptor expressing SPNs, and no consistent effect of quinpirole upon eEPSC peak or PPR in either cell type. Whole-cell patch clamp recording in D1 vs. D2-type SPNs in GKI slices. A) There were no differences in sEPSC mean event frequency in either cell type at young (1-3m) or old (>12m) age points. B) The D2 agonist quinpirole had no consistent modulatory effect upon D1 or D2 eEPSC peak amplitudes, or Paired-pulse ratios expressed as post-application values over pre-application (C). 2-way ANOVA and unpaired Student’s t-test values are shown.

28

647 648 649 650 651 652 653 654 655 656 657 658 659

Figure 3. Supplement 1. Early alterations in GKI dopamine release are not associated with reductions in nigrostriatal dopamine markers and persist despite genotype-dependent increases in DAT protein levels by western blot. A) Staining for presynaptic dopamine markers DAT and tyrosine hydroxylase (TH; i) showed highly specific enrichment in the striatum; however, there were no age or genotype differences (ii). Although variable, the data suggest there is no loss of either terminal marker in GKI striatum. B) Striatal DAT levels were also assessed by western blot (i); although variable results in young GKI suggest no loss of DAT. ii) A significant main genotype effect (due to increased DAT levels in aged GKI) revealed increases in striatal DAT levels in the presence of endogenous G2019S. 2-way values are shown.

29

660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710

1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12.

13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24.

Kordower, J. H. et al. Disease duration and the integrity of the nigrostriatal system in Parkinson’s disease. Brain 136, 2419–2431 (2013). Goldman, J. G. & Postuma, R. Premotor and nonmotor features of Parkinson’s disease. Curr Opin Neurol 27, 434–441 (2014). Chaudhuri, K. R. & Schapira, A. H. Non-motor symptoms of Parkinson’s disease: dopaminergic pathophysiology and treatment. Lancet Neurol. 8, 464–474 (2009). Jenner, P. et al. Parkinson’s disease--the debate on the clinical phenomenology, aetiology, pathology and pathogenesis. J Park. Dis 3, 1–11 (2013). Weintraub, D., Comella, C. L. & Horn, S. Parkinson’s disease--Part 2: Treatment of motor symptoms. Am. J. Manag. Care 14, S49-58 (2008). Weintraub, D., Comella, C. L. & Horn, S. Parkinson’s disease--Part 3: Neuropsychiatric symptoms. Am. J. Manag. Care 14, S59-69 (2008). Weintraub, D., Comella, C. L. & Horn, S. Parkinson’s disease--Part 1: Pathophysiology, symptoms, burden, diagnosis, and assessment. Am. J. Manag. Care 14, S40-8 (2008). Tolosa, E., Compta, Y. & Gaig, C. The premotor phase of Parkinson’s disease. Park. Relat Disord 13 Suppl, S2-7 (2007). Broeders, M. et al. Cognitive change in newly-diagnosed patients with Parkinson’s disease: a 5year follow-up study. J Int Neuropsychol Soc 19, 695–708 (2013). Pedersen, K. F., Larsen, J. P., Tysnes, O. B. & Alves, G. Prognosis of mild cognitive impairment in early Parkinson disease: the Norwegian ParkWest study. JAMA Neurol 70, 580–586 (2013). Litvan, I. et al. MDS Task Force on mild cognitive impairment in Parkinson’s disease: critical review of PD-MCI. Mov Disord 26, 1814–1824 (2011). Goldman, J. G., Aggarwal, N. T. & Schroeder, C. D. Mild cognitive impairment: an update in Parkinson’s disease and lessons learned from Alzheimer’s disease. Neurodegener Dis Manag 5, 425–443 (2015). Calabresi, P., Picconi, B., Tozzi, A., Ghiglieri, V. & Di Filippo, M. Direct and indirect pathways of basal ganglia: a critical reappraisal. Nat Neurosci 17, 1022–1030 (2014). Gerfen, C. R. & Surmeier, D. J. Modulation of striatal projection systems by dopamine. Annu Rev Neurosci 34, 441–466 (2011). Volta, M., Milnerwood, A. J. & Farrer, M. J. Insights from late-onset familial parkinsonism on the pathogenesis of idiopathic Parkinson’s disease. Lancet Neurol. 14, (2015). Ozelius, L. J. et al. LRRK2 G2019S as a cause of Parkinson’s disease in Ashkenazi Jews. N. Engl. J. Med. 354, 424–425 (2006). Hulihan, M. M. et al. LRRK2 Gly2019Ser penetrance in Arab-Berber patients from Tunisia: a case-control genetic study. Lancet Neurol 7, 591–594 (2008). Haugland, R. P. in Handbook of fluorescent probes and research products (ed. Gregory, J.) 554– 572 (Molecular probes Inc., 2002). Gaig, C. et al. Nonmotor symptoms in LRRK2 G2019S associated Parkinson’s disease. PLoS One 9, e108982 (2014). Adams, J. R. et al. PET in LRRK2 mutations: comparison to sporadic Parkinson’s disease and evidence for presymptomatic compensation. Brain 128, 2777–2785 (2005). Nandhagopal, R. et al. Progression of dopaminergic dysfunction in a LRRK2 kindred: a multitracer PET study. Neurology 71, 1790–1795 (2008). Sossi, V. et al. Dopamine turnover increases in asymptomatic LRRK2 mutations carriers. Mov Disord 25, 2717–2723 (2010). Helmich, R. C. et al. Reorganization of corticostriatal circuits in healthy G2019S LRRK2 carriers. Neurology 84, 399–406 (2015). Vilas, D. et al. Clinical and imaging markers in premotor LRRK2 G2019S mutation carriers. Park. Relat Disord 21, 1170–1176 (2015).

30

711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761

25. 26. 27. 28.

29.

30.

31.

32.

33. 34. 35.

36. 37.

38.

39.

40. 41. 42.

43. 44.

Thaler, A. et al. Neural correlates of executive functions in healthy G2019S LRRK2 mutation carriers. Cortex 49, 2501–2511 (2013). Yue, M. et al. Progressive dopaminergic alterations and mitochondrial abnormalities in LRRK2 G2019S knock-in mice. Neurobiol. Dis. 78, (2015). Beccano-Kelly, D. A. et al. Synaptic function is modulated by LRRK2 and glutamate release is increased in cortical neurons of G2019S LRRK2 knock-in mice. Front. Cell. Neurosci. 8, (2014). Beccano-Kelly, D. A. et al. LRRK2 overexpression alters glutamatergic presynaptic plasticity, striatal dopamine tone, postsynaptic signal transduction, motor activity and memory. Hum Mol Genet (2014). doi:10.1093/hmg/ddu543 Volta, M. et al. Chronic and acute LRRK2 silencing has no long-term behavioral effects, whereas wild-type and mutant LRRK2 overexpression induce motor and cognitive deficits and altered regulation of dopamine release. Park. Relat. Disord. 21, (2015). Beccano-Kelly, D. A. et al. LRRK2 overexpression alters glutamatergic presynaptic plasticity, striatal dopamine tone, postsynaptic signal transduction, motor activity and memory. Hum. Mol. Genet. 24, (2015). Milnerwood, A. J. et al. Early Increase in Extrasynaptic NMDA Receptor Signaling and Expression Contributes to Phenotype Onset in Huntington’s Disease Mice (DOI:10.1016/j.neuron.2010.01.008). Neuron 65, (2010). Ade, K. K., Wan, Y., Chen, M., Gloss, B. & Calakos, N. An Improved BAC Transgenic Fluorescent Reporter Line for Sensitive and Specific Identification of Striatonigral Medium Spiny Neurons. Front. Syst. Neurosci. 5, 32 (2011). Milnerwood, A. J. et al. Memory and synaptic deficits in Hip14/DHHC17 knockout mice. Proc. Natl. Acad. Sci. U. S. A. 110, (2013). Petkau, T. L. et al. Synaptic dysfunction in progranulin-deficient mice. Neurobiol. Dis. 45, 711– 722 (2012). Longo, F., Russo, I., Shimshek, D. R., Greggio, E. & Morari, M. Genetic and pharmacological evidence that G2019S LRRK2 confers a hyperkinetic phenotype, resistant to motor decline associated with aging. Neurobiol Dis 71, 62–73 (2014). Trinh, J. et al. A comparative study of Parkinson’s disease and leucine-rich repeat kinase 2 p.G2019S parkinsonism. Neurobiol. Aging 35, 1125–1131 (2014). Kachergus, J. et al. Identification of a novel LRRK2 mutation linked to autosomal dominant parkinsonism: evidence of a common founder across European populations. Am J Hum Genet 76, 672–680 (2005). Milnerwood, A. J. et al. Early Increase in Extrasynaptic NMDA Receptor Signaling and Expression Contributes to Phenotype Onset in Huntington’s Disease Mice. Neuron 65, 178–190 (2010). Matikainen-Ankney, B. A. et al. Altered Development of Synapse Structure and Function in Striatum Caused by Parkinson’s Disease-Linked LRRK2-G2019S Mutation. J Neurosci 36, 7128– 7141 (2016). Cepeda, C. et al. Differential electrophysiological properties of dopamine D1 and D2 receptorcontaining striatal medium-sized spiny neurons. Eur J Neurosci 27, 671–682 (2008). Sara, Y., Virmani, T., Deak, F., Liu, X. & Kavalali, E. T. An isolated pool of vesicles recycles at rest and drives spontaneous neurotransmission. Neuron 45, 563–573 (2005). Ramirez, D. M. O., Khvotchev, M., Trauterman, B. & Kavalali, E. T. Vti1a Identifies a Vesicle Pool that Preferentially Recycles at Rest and Maintains Spontaneous Neurotransmission. Neuron 73, 121–134 (2012). Crawford, D. C. & Kavalali, E. T. Molecular underpinnings of synaptic vesicle pool heterogeneity. Traffic 16, 338–364 (2015). Milnerwood, A. J. & Raymond, L. A. Corticostriatal synaptic function in mouse models of Huntington’s disease: early effects of huntingtin repeat length and protein load. J Physiol 585, 817–831 (2007).

31

762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792

45. 46. 47.

48. 49.

50. 51. 52. 53. 54. 55. 56. 57. 58.

Parisiadou, L. et al. LRRK2 regulates synaptogenesis and dopamine receptor activation through modulation of PKA activity. Nat Neurosci 17, 367–376 (2014). Mei, Y. et al. Adult restoration of Shank3 expression rescues selective autistic-like phenotypes. Nature (2016). doi:10.1038/nature16971 Akopian, G. & Walsh, J. P. Reliable long-lasting depression interacts with variable short-term facilitation to determine corticostriatal paired-pulse plasticity in young rats. J Physiol 580, 225– 240 (2007). Ding, J., Peterson, J. D. & Surmeier, D. J. Corticostriatal and thalamostriatal synapses have distinctive properties. J Neurosci 28, 6483–6492 (2008). Sciamanna, G., Ponterio, G., Mandolesi, G., Bonsi, P. & Pisani, A. Optogenetic stimulation reveals distinct modulatory properties of thalamostriatal vs corticostriatal glutamatergic inputs to fast-spiking interneurons. Sci. Rep. (2015). doi:10.1038/srep16742 Bamford, N. S. et al. Heterosynaptic dopamine neurotransmission selects sets of corticostriatal terminals. Neuron 42, 653–663 (2004). Kreitzer, A. C. & Malenka, R. C. Endocannabinoid-mediated rescue of striatal LTD and motor deficits in Parkinson’s disease models. Nature 445, 643–647 (2007). Phillips, P. E., Hancock, P. J. & Stamford, J. A. Time window of autoreceptor-mediated inhibition of limbic and striatal dopamine release. Synapse 44, 15–22 (2002). Rice, M. E., Patel, J. C. & Cragg, S. J. Dopamine release in the basal ganglia. Neuroscience 198, 112–137 (2011). Jones, S. R. et al. Profound neuronal plasticity in response to inactivation of the dopamine transporter. Proc Natl Acad Sci U S A 95, 4029–4034 (1998). Gainetdinov, R. R. Dopamine transporter mutant mice in experimental neuropharmacology. Naunyn Schmiedebergs Arch Pharmacol 377, 301–313 (2008). Longo, F. et al. Age-dependent dopamine transporter dysfunction and Serine129 phospho-αsynuclein overload in G2019S LRRK2 mice. Acta Neuropathol. Commun. 5, 22 (2017). Kalia, L. V. & Lang, A. E. Parkinson’s disease. Lancet 386, 896–912 (2015). Wile, D. J. et al. Serotonin and dopamine transporter PET changes in the premotor phase of LRRK2 parkinsonism: cross-sectional studies. Lancet Neurol. 16, 351–359 (2017).

32