Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2019 Dec 1.
Published in final edited form as: Cell Calcium. 2018 Sep 12;76:23–35. doi: 10.1016/j.ceca.2018.09.003

Data-driven modeling of mitochondrial dysfunction in Alzheimer’s disease

Patrick Toglia 1, Angelo Demuro 2, Don-On Daniel Mak 3, Ghanim Ullah 1,*
PMCID: PMC6289702  NIHMSID: NIHMS1507845  PMID: 30248575

Abstract

Intracellular accumulation of oligomeric forms of β amyloid (Aβ) are now believed to play a key role in the earliest phase of Alzheimers disease (AD) as their rise correlates well with the early symptoms of the disease. Extensive evidence points to impaired neuronal Ca2+ homeostasis as a direct consequence of the intracellular Aβ oligomers. However, little is known about the downstream effects of the resulting Ca2+ rise on the many intracellular Ca2+-dependent pathways. Here we use multiscale modeling in conjunction with patchclamp electrophysiology of single inositol 1,4,5-trisphosphate (IP3) receptor (IP3R) and fluorescence imaging of whole-cell Ca2+ response, induced by exogenously applied intracellular Aβ42 oligomers to show that Aβ42 inflicts cytotoxicity by impairing mitochondrial function. Driven by patch-clamp experiments, we first model the kinetics of IP3R, which is then extended to build a model for the whole-cell Ca2+ signals. The whole-cell model is then fitted to fluorescence signals to quantify the overall Ca2+ release from the endoplasmic reticulum by intracellular Aβ42 oligomers through G-protein-mediated stimulation of IP3 production. The estimated IP3 concentration as a function of intracellular Aβ42 content together with the whole-cell model allows us to show that Aβ42 oligomers impair mitochondrial function through pathological Ca2+ uptake and the resulting reduced mitochondrial inner membrane potential, leading to an overall lower ATP and increased production of reactive oxygen species and H2O2. We further show that mitochondrial function can be restored by the addition of Ca2+ buffer EGTA, in accordance with the observed abrogation of Aβ42 cytotoxicity by EGTA in our live cells experiments.

Keywords: Alzheimer’s disease, intracellular β amyloid, Ca2+ dyshomeostasis, mitochon-drial dysfunction

Introduction

Alzheimers disease (AD) is associated with increased production and/or impaired clearance of self-aggregating forms of β-amyloid (Aβ) [1, 2]. Strong evidence indicates that soluble oligomeric Aβ aggregates represent the major toxic species in the etiology of AD, leading to uncontrolled elevation of cytosolic Ca2+ [311]. Proposed mechanisms of action of Aβ oligomers include formation of self-aggregating Ca2+-permeable pores in the plasma membrane (PM) [1216], alteration of the physicochemical properties of membrane lipids and proteins [1719], and direct interaction with endogenous Ca2+-permeable receptor/channels [2022].

While most studies on Aβ toxicity to date have focused on the effect of extracellular Aβ oligomers, understanding the effect of intracellular Aβ on cell’s function is becoming increasingly relevant, substantiated by the following observations: (i) intracellular Aβ accumulation precedes extracellular deposition [2325]; (ii) intracellular Aβ are likely to contribute in the earliest phase of the pathogenesis of AD [23, 26, 27]; (iii) the endoplasmic reticulum (ER) of neurons has been identified as the specific site of Aβ production [28]; (iv) accumulation of intracellular Aβ have been linked to profound deficits of long-term potentiation and cognitive dysfunction in AD mice models [29, 30]; and (v) the ER membrane is the site of action for fundamental Ca2+ release due to opening of inositol 1,4,5-trisphosphate (IP3) receptors (IP3Rs) and ryanodine receptors (RyRs), regulating numerous cell’s functions. Furthermore, there is compelling evidence that intracellular Aβ evokes the liberation of Ca2+ from intracellular stores [8, 3133].

We have previously shown that intracellular injection of Aβ42 oligomers stimulates Gprotein-mediated IP3 production and consequently liberate Ca2+ from the ER through activation of IP3Rs [32]. Here we synergistically combine multiscale modeling, patch-clamp electrophysiology of single IP3R, and total internal reflection fluorescence (TIRF) microscopy of global Ca2+ signaling to show that exogenously applied intracellular Aβ42 oligomers inflict cytotoxicity by impairing mitochondrial function. We achieve this by first developing a model for the kinetics of IP3R, driven by single-channel patch-clamp data obtained from type 1 IP3R in isolated nuclei of Xenopus laevis oocytes. This simple four state model accurately reproduces the gating of IP3R as a function of its two ligands, Ca2+ and IP3. The single channel model is used to build a preliminary model for whole-cell Ca2+ signaling. The whole-cell model is then fitted to fluorescence traces representing global Ca2+ signals recorded in Xenopus laevis oocytes in response to exogenously applied intracellular injection of Aβ42 oligomers. This allowed us to quantify and model the overall G-protein-mediated IP3 production and consequently the Ca2+ released from the ER as functions of intracellular Aβ42 concentration ([Aβ42]). The resulting data-driven models for IP3 and cytosolic Ca2+ signals were then combined with a detailed model for mitochondrial Ca2+ signaling and bioenergetics to show that intracellular Aβ42 impairs mitochondrial function due to pathological Ca2+ uptake and diminished mitochondrial inner membrane potential (△Ψm), resulting in overall lower ATP availability to the cell and increased production of reactive oxygen species (ROS) and [H2O2]. This dual theory-experiment approach allows us to investigate the extent of Ca2+ signaling disruption and mitochondrial dysfunction as functions of cell’s intracellular [Aβ42]. We further show that mitochondrial function is restored by the addition of Ca2+ buffer ethylene glycol-bis(β-aminoethyl ether)-N,N,N’,N’-tetraacetic acid (EGTA), in agreement with our experimental observations where Aβ42 cytotoxicity was abrogated by cytosolic injection of EGTA. In addition to unifying observations from two different experimental techniques across different spatiotemporal scales, from single channel to whole-cell, our study provides a testable hypothesis for Aβ42 cytotoxicity. Moreover, this to our knowledge is the first demonstration of extracting parameters related to single channel kinetics from whole-cell Ca2+ signals.

Methods

Computational Model

In the following, we describe the steps to model the intracellular Ca2+ signaling pathways affected by intracellular Aβ42 (Fig. S1). Intracellular Aβ42 oligomers stimulate the production of IP3 through PLC, which binds to IP3Rs to release Ca2+ from the ER. Ca2+ is also released from the ER through leak channels, pumped back into the ER through sarco/ER Ca2+-ATPase (SERCA), and buffered by Ca2+ sensitive dye and other buffers. Ca2+ released from the ER is also buffered by mitochondria that affects the production of ATP and reactive species among other things. Models for mitochondrial function and Ca2+ buffers are based on our previous work [34, 35] and the functional forms for Ca2+ leak and SERCA fluxes are adopted from [36]. To complete the computational framework for all the pathways shown in Fig. S1, we first develop a model for the gating of type-1 IP3R in Xenopus Laevis oocytes as a function Ca2+ and IP3 concentrations, based on the observations from patch-clamp electrophysiology of the channel in the absence of Aβ42. A kinetic scheme closely reproducing the gating of IP3R is crucial for estimating the amount of IP3 generated by Aβ42 as the open probability (Po) of the receptor encodes information about the available IP3 concentration. Since the changes in cytosolic Ca2+ due to Aβ42-induced IP3 generation are observed as fluctuations in the fluorescence of Ca2+-sensitive dye through TIRF microscopy, we next quantify the amount of Ca2+ represented by the fluorescence changes. Once the relationship between the changes in fluorescence and cytosolic Ca2+ concentration is determined, we estimate the amount of IP3 generated by Aβ42. Finally, we put all these components together to estimate the amount of IP3 generated by exogenously applied intracellular Aβ42 and investigate its downstream effects.

Single Ca2+ channel model

Our single-channel IP3R model is an extension of our previous work [35] and uses a previously developed method that ensures the law of mass action and detailed balance [3739]. We refer the interested reader to these papers for details about the modeling procedure for the single IP3R. The study in Ref. [35] modeled the gating of IP3R as a function of Ca2+ only at fixed IP3 concentration of 10μM. While we proposed a seven-state Markov chain that could model both the Ca2+ and IP3 dependence of IP3R gating, we did not derive the rate equations for the model in our previous study [35]. Since IP3 in this study is a dynamic variable that depends on the intracellular concentration of Aβ42, the model developed here takes into account the explicit dependence of the channel’s Po on Ca2+ and IP3 as seen in our patchclam experiments on type-1 IP3R in Xenopus Laevis oocytes [40, 41] (Fig. 1). Furthermore, the four-state model developed in this work is a much simpler model that reproduces the experimental data with the same accuracy as the more complex models proposed previously. Note that each point in Fig. 1 is an average of multiple experiments for the same ligand concentrations of intracellular Ca2+ ([Ca2+]i) and IP3 ([IP3]). We first write the Po of IP3R in terms of occupancies of gating states as

Po([Ca2+]i,[IP3])=ZOZO+ZC, (1)

where

ZO=m=0mmaxn=0nmaxKOmn[Ca2+]im[IP3]n,

and

ZC=m=0mmaxn=0nmaxKCmn[Ca2+]im[IP3]n

are the total occupancies of all open and all close states respectively. KOmn[Ca2+]im[IP3]n and KCmn[Ca2+]im[IP3]n are the unnormalized occupancies of an open and a close state with m Ca2+ and n IP3 bound respectively. As shown below, KOmn and KCmn turn out to be functions of [IP3]. We perform an exhaustive search, testing millions of combinations of states with 0 to 5 Ca2+ and 0 to 4 IP3 bound, searching for the combination so that eq. (1) gives the best fit to the Po of the channel under optimal [IP3]=10 μM (Fig. 1, black circles) according to Bayesian information criterion [42]. A model with four states gives us the best fit to the data. These states are a rest state (R) with no Ca2+ bound, an active state (A) with 2 Ca2+ bound, an open state (O) with 2 Ca2+ bound, and an inhibited state (I) with 5 Ca2+ bound. Thus, Po([Ca2+]i,[IP3])=KO2n[Ca2+]i2[IP3]n/(KC0n[Ca2+]i0[IP3]n+KO2n[Ca2+]i2[IP3]n+KC2n[Ca2+]i2[IP3]n+KC5n[Ca2+]i5[IP3]n). All states have the same number of IP3 bound, which is a consequence of constant [IP3]. For clarity, we drop the subscripts from occupancy parameters and represent KC0n,KC2n,KC5n, and KO2n by KR,KA,KI, and KO respectively. Where KR, KA, KO, and KI are proportional to the occupancies of state R, A, O, and I respectively. We take R to be a reference state with KR = 1. Thus,

Po([Ca2+]i,[IP3])=KO[Ca2+]i21+KO[Ca2+]i2+KA[Ca2+]i2+KI[Ca2+]i5. (2)

We then fit eq. (2) to the Po data taken at [IP3] = 33nM (Fig. 1, blue), 20nM (Fig. 1, red), and 10nM (Fig. 1, green), leaving KA, KO, and KI free each time to obtain the values of these parameters as functions of [IP3]. Thus, we get four values for each one of these three parameters corresponding to the four [IP3] values. These values are then fitted with the following formulas to get the functional forms of the three parameters.

KA=a1(a3a2)[IP3]a2+a3a2+a4, (3)
KI=a5(a7a6)[IP3]a6+a7a6+a8, (4)
KO=a9[IP3]a10[IP3]a10+a11a10. (5)

Constants a1 − a11 given by the fits are listed in Table 1. We would like to remark that this approach of writing the rate constants (or parameters related to the transition rates) as functions of IP3 and/or Ca2+ has been frequently used in literature (see for example [36, 43, 44]) as it significantly simplifies the Markov chain model for the channel. For example, without such simplification a model for IP3R in Xenopus Laevis oocytes will have a minimum of seven states [35] while those in other cells could have anywhere from eight to eighteen states [37, 41, 4547].

Fig. 1:

Fig. 1:

Equilibrium Po of the single IP3R channel in Xenopus laevis oocytes as a function of Ca2+ and IP3 concentrations. Po at [IP3] = 10μM (black), 33nM (blue), 20nM (red), and 10nM (green) as we vary [Ca2+]i. The symbols and lines represent experimental results [40, 41] and model fits respectively.

Table 1:

Constants involved in the functions for occupancy variables. There parameters were obtained from fits to the single channel Po data recorded at different [Ca2+]i and [IP3] values using patch-clamp on type-1 IP3R.

Parameter Value
a1 76.87μM−2
a2 4.98
a3 0.013μM
a4 4.26μM−2
a5 1.45μM−5
a6 17.76
a7 0.025μM
a8 0.000086μM−5
a9 17.66μM−2
a10 1.59
a11 0.017μM

The Po data can only be used to infer the number of states and their occupancies in the model. For the connectivity of these states and the related transition rates, we use the information from rapid perfusion experiments on type-1 IP3R [48]. The kinetic data suggests a direct link between states O and I but no direct link between states R and O. These experiments also point to a direct link between states R and I. All considerations from the kinetic experiments lead to the network shown in Fig. 2. Further details about the connectivity between states and the derivation of transitions rates are given in the next section. Model fits to the experimental data as we vary [Ca2+]i and [IP3] are shown by solid lines in Fig. 1.

Fig. 2:

Fig. 2:

Schematic of the four-state model for single IP3R Ca2+ channel. KXY represents the transition rate from state X to Y where X,Y =, A, O, I , and R respectively.

Transition rates for the single-channel model

In the kinetic experiments, we measured the response of the channel to rapid jumps in both [IP3] and [Ca2+]i. To measure the inhibition latency (the time for the channel to go from actively gating to inhibited), we jumped [Ca2+]i from optimal ([Ca2+]i = 2μM where the Po is maximal or near maximal ) to inhibiting ([Ca2+]i = 300μM where Po is near zero) at fixed [IP3] = 10μM. The optimal and inhibiting values of [Ca2+]i are estimated from the experimental observations about the Po shown in this paper and our previous studies on type-1 IP3R. If the channel had needed to traverse some sequence of states to get from the open state to the inhibited state (or states) there would not exist very short inhibition latencies. At least with the resolution afforded by our experiments, we found non-zero values in the distribution of times to inhibition (fi(ti)) at ti = 0ms. Thus, there appears to be no states intermediate to O and I. We therefore conclude that the open state is directly connected to the inhibited state. There appears to be nothing in the data that demands a link between A and I. We did report a deficit of short activation latencies when we jumped [Ca2+]i from < 10 nM to optimal. This implies that R is not directly connected to O.

In another set of experiments, we jumped [Ca2+]i from near 0 ([Ca2+]i <10 nM) to inhibiting value of 300μM and found that the channel failed to burst (open) in 9 out of 103 experiments. Similarly, in experiments where we jumped [Ca2+]i from 300μM to < 10 nM, 88 out of 94 times the channel deactivated without bursting. Both these observations suggest a direct link between R and I states. Thus, the kinetic data on type-1 IP3R appears to be consistent with a network topology shown in Fig. 2.

Next, we use our novel occupancy/flux formalism [49] and the information from these kinetic experiments to derive expressions for the mean transition times between different states. As discussed in detail previously [37], we prefer this approach over parameterizing the model with reaction rates because we find it simpler and more intuitive. It separates thermodynamic quantities (equilibrium occupancies) from kinetic quantities (reaction fluxes), or equivalently, it separates reaction rates from equilibrium constants. Occupancy and flux parameters are used also because occupancy determination is far more robust than state lifetime measurements. Failure to detect brief events (channel opening and closing) can have a large impact on estimates of mean lifetimes of the states but does not affect state occupancy estimates significantly. Furthermore, this approach ensures detailed balance and microscopic reversibility.

According to this formalism, the mean transition time between two states X and Y where the channel has m and n ligands ([L]) bound respectively is given as

TXY=KX[L]m(1kmn[L]q),TYX=KY[L]n(1kmn[L]q).

Where KX[L]m and KY [L]n is occupancy of state X and Y respectively, kmn[Ca2+]iq is the probability flux between the two states, and q = max(m, n).

Note that in our model there are transitions that involves binding of more than one Ca2+. For example, the channel has 2 and 5 Ca2+ bound in states O and I respectively. However, the direct link between O and I does not necessarily mean that the channel binds 3 Ca2+ simultaneously. It rather means that the states with 3 and 4 Ca2+ bound have very low occupancy and are not required to be included in the model. However, these transition states act as speed-bumps for the probability flux and their effect can be incorporated in the mean transition times. Thus, if there is a low-occupancy transition state Z between states X and Y with o ligands bound, the mean transition times between states X and Y become

TXY=KX[L]m(1kmo[L]r+1kon[L]s),TYX=KY[L]n(1kmo[L]r+1kon[L]s),

where r = max(m, o) and s = max(o, n).

The transition rate from state X to Y (KXY) is simply the inverse of mean time to transition from state X and Y and vice versa.

Using this formalism, the mean transition times between different states in our single-channel model (Fig. 2) are given as

TRA=KR[Ca2+]i0(1k01[Ca2+]i+1k12[Ca2+]i2), (6)
TAR=KA[Ca2+]i2(1k01[Ca2+]i+1k12[Ca2+]i2), (7)
TAO=KA[Ca2+]i2(1k22[Ca2+]i2), (8)
TOA=KO[Ca2+]i2(1k22[Ca2+]i2), (9)
TOI=KO[Ca2+]i2(1k23[Ca2+]i3+1k34[Ca2+]i4+1k45[Ca2+]i5), (10)
TIO=KI[Ca2+]i5(1k23[Ca2+]i3+1k34[Ca2+]i4+1k45[Ca2+]i5), (11)
TRI=KR(1k˜01[Ca2+]i2+1k˜12[Ca2+]i2+1k˜23[Ca2+]i3+1k˜34[Ca2+]i4+1k˜45[Ca2+]i5), (12)
TIR=KI[Ca2+]i5(1k˜01[Ca2+]i+1k˜12[Ca2+]i2+1k˜23[Ca2+]i3+1k˜34[Ca2+]i4+1k˜45[Ca2+]i5). (13)

We assume that the probability flux between two intermediate states is very high so that we can use relative large values for k34, k˜12,k˜23, and k˜34. Thus, the terms involving these constants can be ignored without affecting the mean transition times.

The mean open time of type-1 IP3R under fixed [Ca2+]i of 2μM is 10ms, i.e. TOA ∼ 10ms. Using eq. 9, we get

k22=KO10ms. (14)

The remaining six flux parameters are determined from the rapid-perfusion experiments described above. Observations from more than three thousands of these kinetic experiments are summarized as follows (see [48] for further details).

Mean activation time (changing [Ca2+]i from < 10 nM to 2 μM) = 40 ± 3 ms.

Mean de-activation time (changing [Ca2+]i from 2 μM to < 10nM) = 160 ± 20 ms.

Mean inhibition time (changing [Ca2+]i from 2 μM to 300μM) = 290 ± 40 ms.

Mean inhibition-recovery time (changing [Ca2+]i from 300 μM to 2 μM ) = 2.4 ± 0.2s. Changing [Ca2+]i from < 10 nM to 300μM 9103 experiments failed to cause bursts. Changing [Ca2+]i from 300 μM to < 10 nM, 694 times the channel bursts before getting de-activated.

Thus we can write:

TRA|[Ca2+]i=2μM=40ms, (15)
TAR|[Ca2+]i=10nM=160ms, (16)
TOI|[Ca2+]i=300μM=290ms, (17)
TIO|[Ca2+]i=2μM=2.4s, (18)
TRATRI|[Ca2+]i=300μM=9103, (19)
TIRTIO|[Ca2+]i=10nM=694. (20)

Using identities (15) and (16), we extracted k01 and k12. Identities (17) and (18) give us k23 and k45. k˜01 and k˜45 are extracted from (19) and (20). The values of these six probability flux parameters are given in Table 2.

Table 2:

Probability flux parameters used in the transition rates between the four states of single-channel model. These parameters were estimated from the rapid perfusion experiments measuring the response of type-1 IP3R to rapid switch of ligands Ca2+ and IP3 between different values using patch-clamp electrophysiology [48].

Parameter Value
k01 0.0162/μMms
k12 0.0127/μM2ms
k23 2.1651 × 10−4/ μM3ms
k45 3.5935 × 10−8/ μM5ms
k˜01 0.0014127/μMms
k˜45 5.6297 × 10−7/μM5ms

Thus, the kinetics of IP3Rs in the cell, based on the scheme shown in Fig. 2, are given by the following rate equations that are nonlinear in [Ca2+]i and [IP3] (through rates KXY).

dAdt=KRAR+KOAO(KAR+KAO)A, (21)
dOdt=KIOI+KAOA(KOI+KOA)O, (22)
dIdt=KRIR+KIOO(KIO+KIR)I. (23)

The fraction of channels in state R is given by the conservation of total probability, i.e.

R=1AOI. (24)

Whole-cell Ca2+ signaling model

The single-channel model developed above is used to build the whole-cell model. IP3R releases Ca2+ from the ER through IP3R to the cytoplasm when gating in state O (Jipr). The background Ca2+ leak (Jleak) from the ER is another source of Ca2+ release. Ca2+ is relocated to ER by Sarco-ER Ca2+-ATPase (Js) and buffered by fluorescence dye Fluo-4 ([dye]). In some simulations, we also incorporate mobile Ca2+ buffer EGTA that can uptake cytosolic Ca2+. Some of these fluxes depend nonlinearly on [Ca2+]i and/or [IP3], resulting in a nonlinear set of rate equations [Ca2+]i, dye ([dye]), and EGTA ([EGTA]) concentrations in the whole-cell model as given below.

d[Ca2+]idt=Jipr+JleakJs+(kdyer(Bdye[dye])kdyef[Ca2+]i[dye])+(kEGTAr(BEGTA[EGTA])kEGTAf[Ca2+]i[EGTA]). (25)

Where Bdye and BEGTA is the total concentration of Fluo-4 and EGTA, kdyef and kdyer are the binding and unbinding rates of Ca2+ to Fluo-4, and kEGTAf and kEGTAr are the binding and unbinding rates of Ca2+ to EGTA respectively. The functional form of the three fluxes are adopted from [36] and given as

Jleak=kleak([Ca2+]ER[Ca2+]i), (26)
Js=Vs[Ca2+]insKsns+[Ca2+]ins, (27)
Jipr=kiprO([Ca2+]ER[Ca2+]i). (28)

Ca2+ concentration in the ER ([Ca2+]ER) is given by the conservation of total Ca2+ ([Ca2+]t) in the cell, i.e.

[Ca2+]ER=γ([Ca2+]t[Ca2+]i). (29)

Rate equations for Fluo-4 and EGTA are given as

d[dye]dt=kdyer(Bdye[dye])kdyef[Ca2+]i[dye], (30)
d[EGTA]dt=kEGTAr(BEGTA[EGTA])kEGTAf[Ca2+]i[EGTA]. (31)

Parameter values used in the whole-cell Ca2+ model are given in Table 3.

Table 3:

Parameter values for Ca2+ and buffer rate equations

Parameter Description Value Reference
kleak ER leak flux coefficient 0.0032 s−1 [36]
kipr IP3R flux coefficient 0.2379 s−1 fit to the data
Γ the cytoplasmic-to-ER volume ratio 5.405 [130]
Vs Maximum capacity of SERCA 23.17 μM/s fit to the data
ks SERCA half-maximal activating [Ca2+]i 4.146 μM fit to the data
ns Hill coefficient for SERCA 1.084 fit to the data
[Ca2+]t Total intracellular Ca2+ 50 μM Estimated
Bdye Total fluorescent dye 40 μM Our experimental value
BEGTA Total buffer EGTA 0–3000 μM Our experimental values
kdyer reverse rate of dye 450 s−1 [131]
kdyef forward rate of dye 150μM−1s−1 [131]
kEGTAr reverse rate of EGTA 0.75 s [132]
kEGTAf forward rate of EGTA 5 μM−1s−1 [132]
kdiff Ca2+ diffusional flux coefficient 10 s−1 [36]
γ1 the cytoplasmic-to-microdomain volume ratio 100 [36]

Converting Ca2+-bound dye to fluorescence units

TIRF microscope measures changes in [Ca2+]i in terms of fluorescence changes (ΔF/Fo) as Ca2+ binds and unbinds to indicator dye Fluo-4. Where F0 and ΔF is background fluorescence and change in fluorescence in response to Ca2+ binding to Fluo-4. Thus, we convert Ca2+-bound dye from concentration units to ΔF/Fo, which is then used in fitting the model to fluorescence signals from TIRF microscopy.

A single channel opening on average increases the fluorescence over a 1.2μm × 1.2μm area around the channel relative to the background signal by 0.11 ± 0.01 (i.e. ΔF/Fo = 0.11 ± 0.01) in TIRF microscopy experiments using Fluo-4 [50]. We use this information to convert [Ca2+]-bound [dye] to ΔF/Fo in the model. We simulate a single channel, placed at the center of a 40μm × 40μm area (equal to the area scanned in the TIRF microscopy experiments in [50] as well as in our experiments) and allow it to open for 10ms (the mean open time of IP3R). The [Ca2+]i and [dye] equations described above are modified to include diffusion of these two species with the widely accepted diffusion coefficients of 223μm2/s and 200μm2/s for free Ca2+ (DCa) and Fluo-4 (Ddye) respectively. Furthermore, to side-step the requirement of using extremely small spatial grid and stay consistent with the spatial resolution of 0.3μm of the TIRF microscopy experiments in [50], we partially adopt the procedure in [36], which simulates Ca2+ release and uptake at the microdomain around the channel separately from the rest of the simulation area. The movement of Ca2+ from the microdomain to the nearest gird point (the central grid point in our case) and vice versa is given by diffusion. Also, to remain consistent with the TIRF microscopy experiments that estimated the mean fluorescence change during single channel opening, no EGTA was included in these simulations. With these modifications, the rate equation for [Ca2+]i becomes

d[Ca2+]idt=DCa2[Ca2+]i+Jdiff+JleakJs+(kdyer(Bdye[dye])kdyef[Ca2+]i[dye]). (32)

Where the first and second terms represent the diffusion of Ca2+ between neighboring grid points and the transfer of Ca2+ from the microdomain around IP3R to the grid point at the center of simulating area respectively. The functional form of Jdiff is adopted from [36] where it is modeled with a function similar to Fick’s first law. That is, Jdiff = kdiff([Ca2+]b−[Ca2+]i). For all other grid points, Jdiff = 0. [Ca2+]b is Ca2+ concentration in the microdomain and is given by the rate equation

d[Ca2+]bdt=γ1(JiprJdiff). (33)

Jipr for the single channel is governed by the current through the channel, i.e.

Jipr=I2×F×δV. (34)

Where I = 0.05pA is the observed current through IP3R [51], F is Faraday’s constant, and δV is the volume of a hemisphere over the channel with a radius of 10nm [52, 53].

With the inclusion of the microdomain in the model, the expression for [Ca2+]ER changes accordingly, i.e.

[Ca2+]ER=γ([Ca2+]t[Ca2+]i[Ca2+]b/γ1). (35)

With the inclusion of diffusion, the rate equation for [dye] changes to

d[dye]dt=DCa2[dye]+kdyer(Bdye[dye])kdyef[Ca2+]i[dye]. (36)

Initial and boundary conditions used for different equations are given in section “Numerical and Experimental Methods” below. The above equations are simulated using forward difference method and the peak change in Ca2+-bound dye (Δ[dyeCa2+]) with respect to resting level ([dyeCa2+]o) is recorded, where [dyeCa2+] = Bdye − [dye]. A single channel opening of 10ms resulted in Δ[dyeCa2+]/[dyeCa2+]o = 0.1756 when averaged over 1.2μm × 1.2μm area around the channel. This together with the experimentally observed mean ΔF/Fo during a single IP3R opening is used to convert Ca2+-bound dye to ΔF/Fo.

Estimating IP3 production due to exogenously applied intracellular Aβ42

Injection of different concentrations of Aβ42 oligomers into the oocytes stimulates IP3 production through a G-protein coupled mechanism [32]. IP3 activates IP3R on the ER membrane, releasing Ca2+ into the cytoplasm. The global [Ca2+]i is then represented by ΔF/Fo, imaged through TIRF microscopy. As shown in Fig. 3, after injecting Aβ42, [Ca2+]i first increases to some fixed value and then decays slowly to base level. The rise time, peak, and decay rate of [Ca2+]i depends on the amount of Aβ42 injected into the cell. Guided by the shape of these traces we write the following function for IP3 generated by exogenously applied intracellular Aβ42.

[IP3]=p1[11+exp(p2tp3)]exp(tp4). (37)

Note that in addition to generation by Aβ42, eq. 37 also reflects the subsequent dephosphorylation and phosphorylation of IP3 into inositol bisphosphate and inositol tetrakisphosphate respectively.

Fig. 3:

Fig. 3:

Intracellular injection of Aβ42 oligomers leads to IP3 generation and consequently Ca2+ release from the ER into the cytoplasm. Time-traces of ΔF/Fo indicating [Ca2+]i rise from experiment (circles) and model fits (solid lines) due to injection of 30μg/ml (blue), 10μg/ml (black), 3μg/ml (red), and 1μg/ml (green) [Aβ42] oligomers (A) and corresponding IP3 production (B). The experimental traces represent the mean time course of Ca2+dependent fluorescence recorded from 5 (30 μg/ml), 4 (10 μg/ml), 4 (3 μg/ml), and 6 (1 μg/ml) oocytes respectively [32].

Next, the whole-cell model (eqs. 2130 together with eq. 37 and Ca2+-bound [dye] to ΔF/Fo conversion) is fitted to the fluorescence trace recorded at Aβ42 = 30μg/ml (blue circles in Fig. 3). To perform the fit, we write a Matlab code that computes the least squares error (χ2) as follows. The code first solves the rate equations for the whole-cell (Eqs. 2130) and uses the conversion factor for Ca2+-bound dye to ΔF/Fo to get the whole-cell fluorescence signal given by the model (ΔF/Fo,M). Using the whole-cell fluorescence signal from TIRF microscopy experiments (ΔF/Fo,E) as observable, we can write χ2 as

χ2=1Ni=1N(ΔF/Fo,MΔF/Fo,E)2, (38)

where N is the number of data points in the observed fluorescence trace. χ2 is minimized by leaving parameters p1 −p4, Vs, ks, ns, and kipr free. Where the last four parameters represent maximum SERCA rate, SERCA half-maximal activating [Ca2+]i, Hill coefficient for SERCA, and maximum flux through IP3R.

We repeat the same procedure for fluorescence traces recorded at Aβ42 = 10,3, and 1μg/ml, with the exception that Vs, ks, ns, and kipr are fixed at the values given by the fit to the fluorescence trace recorded at Aβ42 = 30μg/ml. This is due to the fact that we are not aware of any evidence showing that these parameters change as Aβ42 is changed. Parameter values for p1p4 from these fits are given in Table 4 as functions of [Aβ42] and can be closely approximated by the following equations.

p1=0.019[Aβ42]1.15[Aβ42]]1.15+0.00251.15, (39)
p2=26.93exp(88.06[Aβ42]), (40)
p3=4.76exp(2.85[Aβ42])+1.73, (41)
p4=3.96exp(2.28[Aβ42])+304.07. (42)

Where [Aβ42] is in units of μM. In our experiments, 10nl of 30, 10, 3, and 1μg/ml of Aβ42 were injected into the cell. Using the molecular weight of 4514.1 g/mol gives [Aβ42] in units of μM (Table 4). We remark that converting [Aβ42] unit from μg/ml to μM is not required for the model and the functions in Eqs. 3942 are valid in any case with proper unit conversion.

Table 4:

Parameter values involved in IP3 response function. These parameters were obtained by fitting the whole-cell model to the whole-cell Ca2+ signals recorded in our TIRF microscopy experiments.

[Aβ42] injected at 10nL Values
30 μg/ml p1=0.0279 μM
= p2=7.869 s
0.0127 μM p3=1.855 s
p4=24.069 s

10 μg/ml p1=0.023 μM
= p2=19.986 s
0.00423 μM p3=3.155 s
p4=49.359 s

3 μg/ml p1=0.006 μM
= p2=27.203 s
0.00127 μM p3=4.304 s
p4=117428.967 s

1 μg/ml p1=0.003 μM
= p2=25.243 s
0.000423 μM p3=5.948 s
p4=1.507 s

We would like to point out that after expressing in units of nM, the Aβ42 concentrations used in our experiments (Table 4) are well in the range normally used to investigate Aβ42 toxicity [8, 33, 54, 55]. The numbers given in Table 4 are based on the molecular weight of Aβ42 monomer. In oligomeric form, there can be tens of Aβ42 monomer in a single oligomer. Thus the oligomer concentrations used in our experiments are in reality much smaller than the numbers give in Table 4. Furthermore, the intracellular accumulation of Aβ42 can be two to three folds higher than the extracellular concentration [27, 5661].

Details about the numerical scheme used to solve the rate equations and optimize χ2 function are given in the “Numerical Methods” section below.

Mitochondrial function model

The rate equations modeling mitochondrial function are adopted from our previous work [34, 62] and are listed in Table S1. These equations are originally based on the model in [63, 64], which is an extension of the Magnus-Keizer model [65]. The model includes processes such as the tricarboxylic acid (TCA) cycle, electron transport chain (ETC), Ca2+ signaling pathways, reactive oxygen species (ROS), and [H2O2] production. The key components of this model are shown in Fig. S2 and described in SI Appendix, section S1. For further details of the mitochondrial model, we refer the interested reader to Ref. [34]. We would like to point out that different extensions of Magnus-Keizer model are applied to different cell types [6675], indicating the robustness of the model and its suitability to represent many cell types.

The models for IP3 generation by exogenously applied intracellular Aβ42 and whole-cell Ca2+ signaling are coupled with the mitochondrial model through the rate equation for mitochondrial Ca2+ concentration ([Ca2+]m in Table S1) (Fig. S1) to investigate Ca2+ signaling disruption as well as mitochondrial dysfunction as functions of intracellular Aβ42 concentration, and to identify the conditions that would restore mitochondrial function. Specifically, we will show that an increase in [Ca2+]i due to Aβ42 increases Ca2+ influx to mitochondria through mitochondrial Ca2+ uniporter (Vuni in Table S1), which affects mitochondrial ATP and ROS production. The increase in mitochondrial Ca2+ concentration ([Ca2+]m) resulting from intracellular injection of Aβ42 can be counterbalanced and mitochondrial function restored by buffering excessive cytosolic Ca2+ using EGTA.

Numerical and Experimental Methods

Fits to the single IP3R Po data were performed in Mathematica using in-built function NonlinearModelFit. Akaike information criterion (AIC) and Bayesian information criterion were used for assessing the quality of fits and model selection. The model with smallest AIC and BIC scores was selected as the best model fitting the Po data.

Simulations in section “Converting Dye to ΔF/Fo” were performed using forward difference method with a time-step of 50μs and a spatial grid size of 0.3μm, equal to the pixel size in the experiments in [50] that estimated the mean ΔF/Fo value during a single channel opening. Reducing the grid size and time-step did not change our estimates significantly. The initial value of [Ca2+]i was set to 50nM, approximately equal to the resting cytosolic Ca2+ concentration. Based on our simulations, we observed that in resting state approximately 1μM of fluorescence dye was bound to Ca2+. We thus set the initial value of [dye] = 39μM. The boundaries were fixed accordingly at steady state values of [Ca2+]i = 50nM and [dye] = 39μM.

Since there was no stimulus at the start of the experiments, IP3Rs were initially considered to be in the resting state, that is, the initial values of R, A, O, and I were set at 1, 0, 0, and 0 respectively. In simulations involving EGTA, we observed that approximately 1μM of EGTA was bound to Ca2+ in the resting conditions. We thus set the initial value of [EGTA] = BEGTA − 1μM, where BEGTA is the total EGTA concentration. In all cases, simulations were allowed to reach steady state before applying any stimulus, so selecting slightly different initial conditions would not change the final results.

To fit the whole-cell model to the TIRF signals, the rate equations were solved in Matlab using the in-built function “ode15s”. χ2 function was minimized using the in-built Matlab function “fminsearch”. The functions in Eqs. 35 and 3942 were optimized by using inbuilt Matlab function “lsqcurvefit”. Using other Matlab optimization functions such as “fminsearch”, “fmincon”, or “nlinfit” did not change the quality of fits.

Numerical integration of the full model equations (2124, 25, 30, 31, 37, and equations for mitochondrial Ca2+ dynamics and bioenergetics) was performed with Intel Fortran compiler (Intel Corporation, Santa Clara, CA). ODEs were solved using RK4 method. Code producing key results in the paper is available upon request from authors.

Details about the experimental methods are given in the SI Appendix, Section S2.

Results

Our main goal in this study is to use data-driven modeling to estimate the amount of IP3 generated and Ca2+ released from the ER as a function of exogenously applied intracellular Aβ42 concentration. The estimated [IP3] and [Ca2+]i result in a good model for IP3 and Ca2+ dynamics in the cell that paves the way for investigating the cytotoxicity of intracellular Aβ42. Specifically, we show that the observed cytotoxicity of intracellular Aβ42 is due to mitochondrial dysfunction.

Exogenously applied intracellular Aβ42 oligomers release Ca2+ from the ER into the cytoplasm through IP3 generation

Previously reported findings [32] indicate that intracellular Aβ42 oligomers but not monomers, fibrillar, or scrambled Aβ42 peptides leads to significant increase in cytosolic Ca2+ signals after a few tens of seconds of injection. Fig. 3A shows representative traces illustrating increase in ΔF/Fo in response to 10 nl injection of Aβ42 oligomers at concentrations of 1 (6 oocytes), 3 (4 oocytes), 10 (4 oocytes), and 30 (5 oocytes) μg/ml as indicated by the symbols. The amplitude of Ca2+ signals increases, while latency to the onset and peak of the responses shortens progressively with increasing concentration of Aβ42.

To rule out any contribution due to extracellular Ca2+ influx, all these experiments were performed in a Ca2+-free bathing solution including 2 mM EGTA [32]. Intracellular injection of heparine, an IP3R inhibitor, completely abolished the observed Ca2+ signals. Caffeine that acts as a reversible membrane-permeant antagonist of IP3R [7680], also blocked transient global Ca2+ signals. Aβ42-induced Ca2+ signals were also suppressed by treatment of oocytes with PTX to inhibit G-protein mediated activation of PLC and lithium to deplete membrane inositol lipids. All these observations confirm that Aβ42 induces Ca2+ release from the ER through IP3R by stimulating PLC-dependent IP3 generation. It’s worth mentioning that these and several other pathways could contribute to intracellular Ca2+ signals caused by Aβ42 in neurons in vivo. Aβ42 also form pores in the plasma membrane that are permeable to Ca2+ and would impact intracellular Ca2+ signals [12, 1416, 81]. Furthermore, the Aβ42induced Ca2+ signals in neurons could be significantly different since the Ca2+ signaling machinery in neurons is significantly different than oocytes.

To quantify the amount of IP3 generated in response to different concentrations of injected Aβ42, we fit the whole-cell model to the fluorescence traces. The lines in Fig. 3A represent the respective model fits at each concentration of Aβ42 oligomers injected into the oocytes. Fig. 3B shows the time traces of [IP3] from the fits that would evoke the respective ΔF/Fo signals in Fig. 3A. Much like the TIRF signals, the [IP3] follows similar trends of increased amplitude and shortening of onset as the concentration of Aβ42 oligomers is increased. These results suggest that more IP3 is released with increasing concentrations of intracellular Aβ42, leading to rapid changes in ΔF/Fo in the TIRF signals. Furthermore, the delayed onset of TIRF signals relative to the time of injection of Aβ42 is not due to slow diffusional spread of Aβ42 but rather reflects the time required for activation of IP3 production and subsequent accumulation of IP3 levels to evoke Ca2+-release through IP3R since the model does not incorporate any lag due to diffusion of Aβ42. Total [IP3] generated over 60 s is shown in Fig. S3A. At 30 μg/ml and 10 μg/ml we observe comparable amounts of IP3 produced over 60 s. This is because Ca2+ responses to higher doses of Aβ42 display an increasingly rapid decay which could be either due to the desensitization of IP3R or the depletion of IP3 reserves in the cell. Both these effects are not incorporated in the model. As shown in Fig. 4C, we rule out ER depletion as a reason for the rapid decay as the ER holds significant Ca2+ at the peak of cytosolic Ca2+ response. The Po of IP3R as a function of Aβ42 concentration from the fits is shown in Fig. S4, indicating that 10nM Aβ42 leads to a significant Po of the channel. We remark that this is the first demonstration of estimating the Po of IP3R as a function of intracellular Aβ42. Note that Aβ42 activate IP3R by generating IP3 and Fig. S4 does not imply a direct interaction of Aβ42 with the channel.

Fig. 4:

Fig. 4:

High [Ca2+]i in response to intracellular injection of Aβ42 oligomers leads to impaired mitochondrial function. Time-traces of [Ca2+]i (A), [Ca2+]m (B), [Ca2+]ER (C), △Ψm (D), [NADH] (E), [ATP] (F), [O2]m(G), and [H2O2] (H) responses due to injection of 30 μg/ml (blue), 10 μg/ml (black), 3 μg/ml (red), and 1 μg/ml (green) [Aβ42] oligomers. No EGTA is added in these simulations.

Mitochondrial response to evoked Ca2+

Previously, we showed that enhanced Ca2+ release from ER, induced by the gain-of-function enhancement of IP3R due to presenilin mutations found in the brains of patients with Familial AD [39, 82, 83], impaired mitochondrial function [34, 62]. To investigate whether intracellular Aβ42 oligomers also alter mitochondrial function, we couple the whole-cell Ca2+ signaling model to the model for mitochondrial function. Results from these simulations are shown in Fig. 4. At 30 μg/ml, we observe a peak [Ca2+]i of ∼ 4 μM during the 150s simulation that drops to ∼ 2 μM at 10 μg/ml and less than 1 μM at 3 μg/ml and 1 μg/ml. This change in [Ca2+]i from the steady state leads to Ca2+ uptake by mitochondria through Ca2+ uniporter that affects mitochondrial function. Traces of [Ca2+]m at different Aβ42 concentrations are shown in Fig. 4B. Total [Ca2+]i and [Ca2+]m during 60 s are shown in Fig. S3B and S3C respectively. At 30 and 10 μg/ml Aβ42 oligomers, the total amount of cytosolic Ca2+ exceeds 50 μM leading to an increase in [Ca2+]m of well over 100 μM over 60 s. Such excessive [Ca2+]m have been shown to impair mitochondrial function, leading to apoptosis [34, 62, 8486].

Changes in key variables governing mitochondrial function in response to Ca2+ release from the ER (Fig. 4C) through IP3R and entering the matrix through the uniporter are shown in Fig. 4C-H. Intracellular injection of 30μg/ml Aβ42 oligomers leads to a decrease of 24 mV in mitochondrial membrane potential (△Ψm), thereby disrupting the proton motive force (△μH) formed by respiration pumps oxidizing NADH on the ETC. While our model incorporates different enzymes and dehydrogenases involved in the TCA cycle, we primarily focus on how excessive Ca2+ alters [NADH] production (Fig. 4E) and consequently affects the synthesis of ATP and toxic agents. The large influx of Ca2+ increases the rate of oxygen consumption (oxidation of NADH) that leads to lower [NADH] and the rates of three Ca2+-sensitive enzymes in the TCA cycle (α-ketoglutarate dehydrogenase, isocitrate dehydrogenase, and malate dehydrogenase) that stimulates NADH production (Fig. S5). However, the increase in the rate of NADH oxidation exceeds that of the three enzymes in the TCA cycle combined, thereby, lowering the rate of NADH production. Between lowered [NADH] and diminished △Ψm, [ATP] production is negatively affected as increasing Aβ42 concentration leads to higher Ca2+ influx into the mitochondria (Fig. 4F). At 30 and 10 μg/ml Aβ42 oligomers, [ATP] drops by 600 μM. Even at 3 μg/ml oligomers there is almost 100 μM decrease in [ATP], dropping to 12 μM at 3 μg/ml. For smaller concentrations of Aβ42, the effect on mitochondrial function is negligible.

Fig. 4G and3H show the amount of ROS ([O2]m) and [H2O2] produced by mitochondria as we increase Aβ42 concentration. Enhanced and prolonged oxidation leads to significant ROS production and therefore [H2O2], both extremely toxic to the cell when present in excess [8789]. At Aβ42 > 3 μg/ml, we observe more than 80 μM change in ROS. Interestingly, in case of 3 μg/ml Aβ42, ROS shows a steady rise over time. This is due to the Ca2+ response observed in our experiments that did not decay over the duration of the experiment. The amount of [H2O2] decreases due to diminished △Ψm but increases later and overshoots the resting value as the ROS has reached its peak values. The overshoot is significant and longlasting in case of 30 and 10 μg/ml Aβ42. Our simulation results clearly show that intracellular Aβ42 leads to compromised mitochondrial function.

EGTA restores mitochondrial function

The experiments in [32] also showed that no cells remained viable within 37h of Aβ42 injection (Fig. 6). Interestingly, the cell viability was restored to more than 70% when injected with mobile Ca2+ buffer EGTA to a final concentration of 3 mM. We argue that EGTA significantly reduces the amount of [Ca2+]i that the mitochondrial uniporter senses, which would restore mitochondrial function and consequently will improve cell viability. Fig. 5 shows the same simulation as in Fig. 4 except that here we add 1000 μM EGTA to the cell. We see a significant decrease in the peak and total [Ca2+]i over 60 s evoked by Aβ42 oligomers (Fig. 5A and Fig. S6A), leading to a slower rise and lower peak in [Ca2+]m (Fig. 5B). The total increase in [Ca2+]m over 60 s drops to less than a third of the concentration where no EGTA is present (Fig. S6B). Similar trends are observed in other key variables governing mitochondrial function (Fig. 5D and E). The decrease in △Ψm, [NADH], and [ATP] drops to less than 2 mV, 150 μM, and 200 μM, respectively at 30 μg/ml Aβ42. A similar trend is observed for other Aβ42 concentrations. ROS and [H2O2] production are also affected by the addition of EGTA in similar fashion and their peak values drop as compared to the case with no EGTA. The restoration of the mitochondrial function due to EGTA could explain the observations in Fig. 6.

Fig. 6:

Fig. 6:

Ca2+ buffer EGTA restores cell viability. No oocytes injected with 1 μg/ml of Aβ42 oligomers remained viable after 37 hours of injection (black). Injecting 3mM of EGTA, restored the viability of oocytes to 70% [32]. Oocytes were bathed in a solution containing no added Ca2+ in order to eliminate influx from extracellular solution.

Fig. 5:

Fig. 5:

EGTA can reverse the affect of Aβ42 oligomers on Ca2+ signaling and mitochondrial function. Time-traces of [Ca2+]i (A), [Ca2+]m (B), [Ca2+]ER (C), 4Ψm (D), [NADH] (E), [ATP] (F), [O2]m(G), and [H2O2] (H) response due to injection of 30 μg/ml (blue), 10 μg/ml (black), 3 μg/ml (red), and 1 μg/ml (green) [Aβ42] oligomers with 1000 μM [EGTA] added to the cell.

To estimate the amount of [EGTA] needed to rescue mitochondrial function in varying concentration of intracellular Aβ42 oligomers, we show surface plots of the peak change in key variables from steady-state values as functions of [Aβ42] and [EGTA] in Fig. 7. For example, 30μg/ml Aβ42 oligomers leads to a 24 mV and 2 mV drop (with respect to the resting state) in △Ψm in the absence of EGTA and presence of 1000 μM EGTA respectively. Thus in the surface plot for Δ[△Ψm], a point at [Aβ42] = 30μg/ml and EGTA = 0 and 1000 μM will be equal to 24 and 2 mV respectively. We vary [Aβ42] and [EGTA] over a range of 0–30 μg/ml (0.0127 μM) and 0–3000 μM respectively and determine the peak change for various combinations of [EGTA] and [Aβ42]. Fig. 7A-F shows the peak change in [Ca2+]i, [Ca2+]m, [O2]m, ΔΨm, [H2O2], and [ATP] respectively (note that panels (D) and (F) indicate a decrease (absolute value of change) with respect to resting values). Dependence of all key variables on [Aβ42] is prevalent as the peak change increases with [Aβ42]. Predictably, increasingly higher concentration of EGTA is needed to buffer the higher [Ca2+]i due to increasing Aβ42 content and restore mitochondrial function. The key observation that can be made from Fig. 7 is that for small and intermediate values of [Aβ42], EGTA will restore mitochondrial function. However, for large concentrations of Aβ42 oligomers, restoring mitochondrial function with EGTA alone is not possible. This could explain the partial restoration of cell viability (up to 70%) by 3mM EGTA in our experiments (Fig. 6). Nevertheless, a polytherapy involving EGTA with drugs such as pertussis toxin to inhibit G-protein, lithium to deplete membrane inositol lipids, and IP3R antagonist heparin might restore mitochondrial function and hence fully abrogate cytotoxicity. Although not tested simultaneously, each one of EGTA, pertussis toxin, and lithium individually have been shown to restore cell viability to more than 70% of the control value [32]. Furthermore, higher resting [Ca2+]i has been observed in primary cortical neurons from triple transgenic mouse model of AD that exhibits intraneuronal accumulation of Aβ oligomers as compared to non-transgenic neurons [90]. Higher base levels of [Ca2+]i have been shown to significantly affect the average ATP production of mitochondria [72]. Thus a carefully devised polytherapy that would not only buffer the transient rise in Ca2+ concentration but also restore the physiological resting [Ca2+]i would be more appropriate.

Fig. 7:

Fig. 7:

EGTA rescues mitochondrial function. Surface plots showing the dependance of Aβ42 and [EGTA] on change (Δ) in values from steady state of [Ca2+]i (A), [Ca2+]m (B), [O2]m(C), △Ψm (D), [H2O2] (E), and [ATP] (F). Note that the change in panels (D) and (F) indicates a decrease with respect to resting values, while all other panels indicate an increase with respect to the resting levels.

Discussion

Understanding the downstream effects of intracellular Aβ oligomers accumulation in AD-affected neurons is crucial as they are shown to lead to profound deficits in memory and cognitive function [29, 30], and are likely to play a major role in the earliest phase of AD pathogenesis [23, 27]. As reported by different groups including us, the liberation of Ca2+ from intracellular stores is seen as a possible mechanism for the cytotoxicity of intracellular Aβ oligomers [8, 3133]. In line with this notion, we have shown that buffering the Ca2+ released from the ER in response to intracellular injection of Aβ42 oligomers restores cell viability and protects against cell death [32]. However, the intermediate pathways linking Aβ42-induced Ca2+ rise to cytotoxicity are not well understood.

During normal cell functioning, transient intracellular Ca2+ rise is restored to base level by PM Ca2+ ATPase, Na+/Ca2+ exchange, and SERCA pumps located in the ER membrane [91]. Cytosolic Ca2+ is also buffered by mitochondria through Ca2+ uniporter [92]. Prolonged and frequent occurrence of high cytosolic Ca2+ has been suggested to compromise mitochondrial function [85, 86, 93, 94]. This together with the observed Ca2+ up-regulation and mitochondrial dysfunction in several neurodegenerative diseases could provide a link between Ca2+ rise in these diseases and cytotoxicity [95103]. The studies in Refs. [99, 101] in particular show the association of intracellular Aβ with the extent of mitochondrial dysfunction both in AD and Down’s Syndrome where most patients develop AD neuropathology [104]. However, the molecular mechanism through which intracellular Aβ impairs mitochondrial function is not well understood. Moreover, other potential downstream effects of Aβ-triggered Ca2+ liberation from the ER have not been properly investigated and identified yet.

Our approach combines patch-clamp electrophysiology of single Ca2+ channel and fluorescent imaging of whole-cell Ca2+ signals into a multiscale model that allows us to closely reproduce key observations at both scales, and estimate the amount of IP3 generated and consequently the Ca2+ content liberated from the ER as a function of Aβ42 concentration during our experiments [32]. While inferring parameters related to the single-channel behavior (e.g. [IP3] and channel flux) from localized Ca2+ signals (μm-sized events called puffs) has been demonstrated before [105, 106], this to our knowledge is the first exhibition of estimating such parameters from the whole-cell Ca2+ signals. We are also not aware of any other study directly fitting a whole-cell Ca2+ model to TIRF signals. The data-driven model is then used to investigate the effect of Aβ42-induced Ca2+ liberation on mitochondrial function.

We show that intracellular Aβ42 oligomers disrupt mitochondrial function by inducing global Ca2+ signals through IP3 generation, leading to high [Ca2+]m and ROS. The enhanced [Ca2+]m also decreases △Ψm and subsequently 4μH to a point where ATP production is negatively affected (Fig. S5). These observations are well in line with the higher level of ROS, decreased △Ψm, and decreased ATP production observed in hippocampal and cortical mitochondria of AD transgenic mice as compared to non-transgenic mice [101, 107]. The decreased respiration control ratio observed in [101] was attribute to diminished △Ψm in different brain regions which is consistent with our results. The dependence of the level of mitochondrial dysfunction on Aβ42 concentration (hence Ca2+ content) would also explain the relative severe impairment of synaptic mitochondria as compared to non-synaptic ones as the Ca2+ liberation in synapses would be compounded by the higher influx from the extracellular space through Ca2+ channels in the PM [101]. Our results are also consistent with the observed oxidative stress in AD and Down’s Syndrome [104, 108, 109].

We remark that the computational framework proposed here is different than our previous study [34], where we fused experimental time-traces for cytosolic [Ca2+]i with the model for mitochondrial function. In other words, we only modeled mitochondrial function without modeling the pathways regulating [Ca2+]i, [Ca2+]ER, and [IP3]. As a simplification, the model for these variables was replaced by the experimental time-trace for cytosolic calcium. While this made modeling the effect Ca2+ signaling disruptions on mitochondrial function a lot easier, the framework is specific to the cells studied as the input was coming directly from the cell observed experimentally. The framework developed here is more general as it models all components involved in the Ca2+ signaling disruption and its downstream effects. Furthermore, Ca2+ signaling disruptions and consequently the mitochondrial dysfunction studied previously was due to Familial Alzheimer’s disease-causing mutations in presenilin. However, more than 95% cases of Alzheimer’s disease are sporadic where intracellular Aβ42, most likely through Ca2+ signaling disruptions, are believed to play a key role in the earliest phase of the disease.

Mitochondrial dysfunction due to intracellular Aβ in Ca2+-dependent manner makes a testable mechanism for neuronal toxicity in AD. For example, Simakova and Arispe demonstrated that neurons with low ATP levels exhibit higher affinity to Aβ42 binding, and are consequently vulnerable to toxicity [110]. In healthy cells, PM asymmetry is maintained by the activity of aminophospholipid translocase flippase – an ATP-dependent transmembrane transporter that shuttle phosphatidylserine from the outer to the inner leaflet of the membrane. The reverse transport of phosphatidylserine from the inner to the outer surface of PM, induced by stress and lack of ATP, is believed to be the basis for the cell-selective neurotoxicity of Aβ in certain brain regions [110]. One can envisage that as a result of Ca2+ rise due to intracellular Aβ oligomers, nearby mitochondrial ATP production drops in parts of the cell where high level of ATP is constantly required to maintain PM asymmetry through steady flippase activity. The lack of ATP and rise in ROS would impair flippase activity, allowing spontaneous translocation to the outer leaflets of the PM and leaving the cell membrane more susceptible to extracellular Aβ oligomers binding. This would lead to the exacerbation of Ca2+ dyshomeostasis and the structural and functional disruption of neuronal networks in AD [110, 111].

In addition to biogenesis, several aspects of mitochondrial dynamics are also impaired in the early stages of AD, including slower anterograde and retrograde axonal transport, decreased number of large mitochondria, lower number of mitochondria overall, and disrupted balance between fusion and fission when compared to physiological conditions [112118]. Mitochondrial transport [119122], fission, fusion, fragmentation [123, 124], and morphology [125, 126] are all disrupted by increased ROS, reduced ATP, depolarized mitochondrial membrane, and up-regulated intracellular Ca2+, all of which we show are significantly influenced by intracellular Aβ42. The fact that sustained increase of [Ca2+]i triggers intracellular Aβ42 production would further exacerbate the potential impairment of mitochondrial dynamics and function [11, 127].

The fluorescence signals in our TIRF experiments and variables in the model represent the average effect over the entire cell. That is, we treat the oocyte as a point cell and ignore the spatial aspects of various signaling pathways such as the arrangement of IP3Rs into clusters, the realistic distribution of clusters in the cell [53], and the close contacts between mitochondrion and the ER leading to privileged entry of Ca2+ released through IP3Rs into mitochondrion [128, 129]. Investigating the changes in the spatial aspects of Ca2+ signals in AD, their role in cell dysfunction, and numerous other questions related to the mitochondrial dynamics listed in the preceding paragraphs are beyond the scope of this study and the subject of our future research.

Supplementary Material

1

Highlights.

  1. A simple four state model for IP3R, driven by extensive single channel patch-clamp experiments.

  2. Estimating Ca2+ and IP3 from TIRF Signals.

  3. Estimating the amount of IP3 generated due to intracellular beta amyloid.

  4. Evaluating mitochondrial dysfunction due to intracellular beta amyloid.

  5. Estimating single channel parameters from whole-cell Ca2+ experiments.

  6. The use of Ca2+ buffer EGTA to restore viability of cells with Alzheimer’s disease pathology.

Acknowledgements

This work was supported by NIH through grants R01 AG053988 (to AD and GU) and R01 GM065830 (to DODM).

Footnotes

Conflict of Interest

Authors declare no conflict of interest.

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • [1].Hardy JA, Higgins GA, Alzheimer’s disease: the amyloid cascade hypothesis, Science 256 (1992) 184. [DOI] [PubMed] [Google Scholar]
  • [2].Haass C, Schlossmacher MG, Hung AY, Vigo-Pelfrey C, Mellon A, Ostaszewski BL, Lieberburg I, Koo EH, Schenk D, Teplow DB, et al. , Amyloid β-peptide is produced by cultured cells during normal metabolism, Nature 359 (1992) 322–325. [DOI] [PubMed] [Google Scholar]
  • [3].Khachaturian Z, Calcium hypothesis of Alzheimer0s disease and brain aging, Ann. N. Y. Acad. Sci 747 (1994) 1–11. [DOI] [PubMed] [Google Scholar]
  • [4].Alzheimer’s ACHW, Calcium hypothesis of Alzheimer′s disease and brain aging: A framework for integrating new evidence into a comprehensive theory of pathogenesis., Alzheimers Dement. 13 (2017) 178. [DOI] [PubMed] [Google Scholar]
  • [5].Green KN, LaFerla FM, Linking calcium to Aβ and Alzheimers disease, Neuron 59 (2008) 190–194. [DOI] [PubMed] [Google Scholar]
  • [6].Berridge MJ, Calcium hypothesis of Alzheimer0s disease, Pflugers Arch., EJP 459 (2010) 441–449. [DOI] [PubMed] [Google Scholar]
  • [7].Bezprozvanny I, Mattson MP, Neuronal calcium mishandling and the pathogenesis of Alzheimer’s disease, Trends Neurosci. 31 (2008) 454–463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [8].Demuro A, Mina E, Kayed R, Milton SC, Parker I, Glabe CG, Calcium dysregulation and membrane disruption as a ubiquitous neurotoxic mechanism of soluble amyloid oligomers, J. Biol. Chem 280 (2005) 17294–17300. [DOI] [PubMed] [Google Scholar]
  • [9].Demuro A, Parker I, Stutzmann GE, Calcium signaling and amyloid toxicity in Alzheimer0s disease, J. Biol. Chem 285 (2010) 12463–12468. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [10].Chakroborty S, Stutzmann GE, Calcium channelopathies and Alzheimer′s disease: Insight into therapeutic success and failures, Eur. J. Pharmacol 739 (2014) 83–95. [DOI] [PubMed] [Google Scholar]
  • [11].De Caluwé J, Dupont G, The progression towards Alzheimer’s disease described as a bistable switch arising from the positive loop between amyloids and Ca2+, Journal of theoretical biology 331 (2013) 12–18. [DOI] [PubMed] [Google Scholar]
  • [12].Arispe N, Pollard HB, Rojas E, Giant multilevel cation channels formed by Alzheimer?s disease amyloid beta-protein [Abeta P-(1–40)] in bilayer membranes, Proceedings of the National Academy of Sciences 90 (1993) 10573–10577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [13].Pollard HB, Rojas E, Arispe N, A new hypothesis for the mechanism of amyloid toxicity, based on the calcium channel activity of amyloid β protein (aβp) in phospholipid bilayer membranes, Ann. N. Y. Acad. Sci 695 (1993) 165–168. [DOI] [PubMed] [Google Scholar]
  • [14].Demuro A, Smith M, Parker I, Single-channel Ca2+ imaging implicates Aβ42 amyloid pores in Alzheimers disease pathology, J. Cell Biol (2011) jcb–201104133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [15].Lin H, Bhatia R, Lal R, Amyloid β protein forms ion channels: implications for alzheimer?s disease pathophysiology, The FASEB Journal 15 (2001) 2433–2444. [DOI] [PubMed] [Google Scholar]
  • [16].Ullah G, Demuro A, Parker I, Pearson JE, Analyzing and modeling the kinetics of amyloid beta pores associated with Alzheimer?s disease pathology, PloS one 10 (2015) e0137357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Mason RP, Jacob RF, Walter MF, Mason PE, Avdulov NA, Chochina SV, Igbavboa U, Wood WG, Distribution and fluidizing action of soluble and aggregated amyloid β-peptide in rat synaptic plasma membranes, J. Biol. Chem 274 (1999) 18801–18807. [DOI] [PubMed] [Google Scholar]
  • [18].Hertel C, Terzi E, Hauser N, Jakob-Røtne R, Seelig J, Kemp J, Inhibition of the electrostatic interaction between β-amyloid peptide and membranes prevents βamyloid-induced toxicity, Proc. Natl. Acad. Sci 94 (1997) 9412–9416. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [19].Sokolov Y, Kozak JA, Kayed R, Chanturiya A, Glabe C, Hall JE, Soluble amyloid oligomers increase bilayer conductance by altering dielectric structure, J. Gen. Physiol 128 (2006) 637–647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [20].De Felice FG, Velasco PT, Lambert MP, Viola K, Fernandez SJ, Ferreira ST, Klein WL, Aβ oligomers induce neuronal oxidative stress through an Nmethyl-D-aspartate receptor-dependent mechanism that is blocked by the Alzheimer?s drug memantine, J. Biol. Chem 282 (2007) 11590–11601. [DOI] [PubMed] [Google Scholar]
  • [21].Wang H-Y, Lee DH, D’Andrea MR, Peterson PA, Shank RP, Reitz AB, βamyloid1–42 binds to α7 nicotinic acetylcholine receptor with high affinity implications for alzheimer’s disease pathology, J. Biol. Chem 275 (2000) 5626–5632. [DOI] [PubMed] [Google Scholar]
  • [22].Texidó L, Martín-Satuée M, Alberdi E, Solsona C, Matute C, Amyloid β peptide oligomers directly activate nmda receptors, Cell Calcium 49 (2011) 184–190. [DOI] [PubMed] [Google Scholar]
  • [23].Gouras GK, Tsai J, Naslund J, Vincent B, Edgar M, Checler F, Greenfield JP, Haroutunian V, Buxbaum JD, Xu H, et al. , Intraneuronal Aβ42 accumulation in human brain, Am. J. Pathol 156 (2000) 15–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [24].D’Andrea M, Nagele R, Wang H-Y, Peterson P, Lee D, Evidence that neurones accumulating amyloid can undergo lysis to form amyloid plaques in Alzheimer′s disease, Histopathology 38 (2001) 120–134. [DOI] [PubMed] [Google Scholar]
  • [25].Wang H-Y, D’Andrea MR, Nagele RG, Cerebellar diffuse amyloid plaques are derived from dendritic Aβ42 accumulations in purkinje cells, Neurobiol. Aging 23 (2002) 213–223. [DOI] [PubMed] [Google Scholar]
  • [26].Busciglio J, Lorenzo A, Yeh J, Yankner BA, β-amyloid fibrils induce tau phosphorylation and loss of microtubule binding, Neuron 14 (1995) 879–888. [DOI] [PubMed] [Google Scholar]
  • [27].Pensalfini A, Albay R, Rasool S, Wu JW, Hatami A, Arai H, Margol L, Milton S, Poon WW, Corrada MM, et al. , Intracellular amyloid and the neuronal origin of alzheimer neuritic plaques, Neurobiology of disease 71 (2014) 53–61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [28].Hartmann T, Bieger SC, Brühl B, Tienari PJ, Ida N, Allsop D, Roberts GW, Masters CL, Dotti CG, Unsicker K, et al. , Distinct sites of intracellular production for Alzheimer’s disease Aβ40/42 amyloid peptides, Nat. Med 3 (1997) 1016–1020. [DOI] [PubMed] [Google Scholar]
  • [29].Oddo S, Caccamo A, Shepherd JD, Murphy MP, Golde TE, Kayed R, Metherate R, Mattson MP, Akbari Y, LaFerla FM, Triple-transgenic model of Alzheimer?s disease with plaques and tangles: intracellular Aβ and synaptic dysfunction, Neuron 39 (2003) 409–421. [DOI] [PubMed] [Google Scholar]
  • [30].Knobloch M, Konietzko U, Krebs DC, Nitsch RM, Intracellular Aβ and cognitive deficits precede β-amyloid deposition in transgenic arcaβ mice, Neurobiol. Aging 28 (2007) 1297–1306. [DOI] [PubMed] [Google Scholar]
  • [31].Ferreiro E, Oliveira CR, Pereira C, Involvement of endoplasmic reticulum Ca2+ release through ryanodine and inositol 1, 4, 5-triphosphate receptors in the neurotoxic effects induced by the amyloid-β peptide, J. Neurosci. Res 76 (2004) 872–880. [DOI] [PubMed] [Google Scholar]
  • [32].Demuro A, Parker I, Cytotoxicity of intracellular Aβ42 amyloid oligomers involves Ca2+ release from the endoplasmic reticulum by stimulated production of inositol trisphosphate, J. Neurosci 33 (2013) 3824–3833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [33].Jensen LE, Bultynck G, Luyten T, Amijee H, Bootman MD, Roderick HL, Alzheimer0s disease-associated peptide Aβ42 mobilizes ER Ca2+ via InsP3R-dependent and-independent mechanisms, Front Mol. Neurosci 6 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [34].Toglia P, Cheung K-H, Mak D-OD, Ullah G, Impaired mitochondrial function due to familial Alzheimer0s disease-causing presenilins mutants via Ca2+ disruptions, Cell calcium 59 (2016) 240–250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [35].Ullah G, Parker I, Mak D-OD, Pearson JE, Multi-scale data-driven modeling and observation of calcium puffs, Cell calcium 52 (2012) 152–160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [36].Cao P, Tan X, Donovan G, Sanderson MJ, Sneyd J, A deterministic model predicts the properties of stochastic calcium oscillations in airway smooth muscle cells, PLoS Comput. Biol 10 (2014) e1003783. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [37].Ullah G, Mak D-OD, Pearson JE, A data-driven model of a modal gated ion channel: The inositol 1, 4, 5-trisphosphate receptor in insect Sf9 cells, J. Gen. Physiol 140 (2012) 159–173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [38].Ullah G, Bruno WJ, Pearson JE, Simplification of reversible markov chains by removal of states with low equilibrium occupancy, J. Theor. Biol 311 (2012) 117–129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [39].Mak D-OD, Cheung K-H, Toglia P, Foskett JK, Ullah G, Analyzing and quantifying the gain-of-function enhancement of IP3 receptor gating by familial Alzheimer0s disease causing mutants in presenilins, PloS Comp. Biol 84 (2015) 143–152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [40].Mak D-OD, McBride S, Foskett JK, Inositol 1, 4, 5-tris-phosphate activation of inositol tris-phosphate receptor Ca2+ channel by ligand tuning of Ca2+ inhibition, Proc. Natl. Acad. Sci 95 (1998) 15821–15825. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [41].Shuai J, Pearson JE, Foskett JK, Mak D-OD, Parker I, A kinetic model of single and clustered IP3 receptors in the absence of Ca2+ feedback, Biophys. J 93 (2007) 1151–1162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [42].Schwarz G, et al. , Estimating the dimension of a model, Ann. Stat 6 (1978) 461–464. [Google Scholar]
  • [43].Siekmann I, Wagner LE, Yule D, Crampin EJ, Sneyd J, A kinetic model for type i and ii IP3R accounting for mode changes, Biophysical journal 103 (2012) 658–668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [44].Equations for InsP3 receptor-mediated [Ca2+]i oscillations derived from a detailed kinetic model: a Hodgkin-Huxley like formalism, author=Yue-Xian Liand John Rinzel, journal=Journal of theoretical Biology, volume=166, number=4, pages=461–473, year=1994, publisher=Elsevier (????). [Google Scholar]
  • [45].De Young GW, Keizer J, A single-pool inositol 1, 4, 5-trisphosphate-receptor-based model for agonist-stimulated oscillations in Ca2+ concentration., Proc. Natl. Acad. Sci 89 (1992) 9895–9899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [46].Dawson AP, Edward J, Irvine RF, Kinetic model of the inositol trisphosphate receptor that shows both steady-state and quantal patterns of Ca2+ release from intracellular stores, Biochemical Journal 370 (2003) 621–629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [47].Swillens S, Dupont G, Combettes L, Champeil P, From calcium blips to calcium puffs: theoretical analysis of the requirements for interchannel communication, Proceedings of the National Academy of Sciences 96 (1999) 13750–13755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [48].Mak D-OD, Pearson JE, Loong KPC, Datta S, Fernández-Mongil M, Foskett JK, Rapid ligand-regulated gating kinetics of single inositol 1, 4, 5-trisphosphate receptor Ca2+ release channels, EMBO Rep. 8 (2007) 1044–1051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [49].Yang J, Bruno WJ, Hlavacek WS, Pearson JE, On imposing detailed balance in complex reaction mechanisms, Biophysical journal 91 (2006) 1136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [50].Smith IF, Parker I, Imaging the quantal substructure of single IP3R channel activity during Ca2+ puffs in intact mammalian cells, Proc. Natl. Acad. Sci 106 (2009) 6404–6409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [51].Vais H, Foskett JK, Ullah G, Pearson JE, Mak D-OD, Permeant calcium ion feed-through regulation of single inositol 1, 4, 5-trisphosphate receptor channel gating, The Journal of general physiology 140 (2012) 697–716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [52].Taylor CW, Da Fonseca PC, Morris EP, IP3 receptors: the search for structure, Trends Biochem. Sci 29 (2004) 210–219. [DOI] [PubMed] [Google Scholar]
  • [53].Foskett JK, White C, Cheung K-H, Mak D-OD, Inositol trisphosphate receptor Ca2+ release channels, Physiol. Rev 87 (2007) 593–658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [54].Bucciantini M, Calloni G, Chiti F, Formigli L, Nosi D, Dobson CM, Stefani M, Prefibrillar amyloid protein aggregates share common features of cytotoxicity, Journal of Biological Chemistry 279 (2004) 31374–31382. [DOI] [PubMed] [Google Scholar]
  • [55].Kawahara M, Kuroda Y, Arispe N, Rojas E, Alzheimer’s β-amyloid, human islet amylin, and prion protein fragment evoke intracellular free calcium elevations by a common mechanism in a hypothalamic gnrh neuronal cell line, Journal of Biological Chemistry 275 (2000) 14077–14083. [DOI] [PubMed] [Google Scholar]
  • [56].Waters J, The concentration of soluble extracellular amyloid-β protein in acute brain slices from CRND8 mice, PLoS One 5 (2010) e15709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [57].Tomic JL, Pensalfini A, Head E, Glabe CG, Soluble fibrillar oligomer levels are elevated in alzheimer’s disease brain and correlate with cognitive dysfunction, Neurobiology of disease 35 (2009) 352–358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [58].Barykin EP, Mitkevich VA, Kozin SA, Makarov AA, Amyloid β Modification: A Key to the Sporadic Alzheimer’s Disease?, Frontiers in genetics 8 (2017) 58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [59].LaFerla FM, Green KN, Oddo S, Intracellular amyloid-β in Alzheimer’s disease, Nature Reviews Neuroscience 8 (2007) 499. [DOI] [PubMed] [Google Scholar]
  • [60].Hu X, Crick SL, Bu G, Frieden C, Pappu RV, Lee J-M, Amyloid seeds formed by cellular uptake, concentration, and aggregation of the amyloid-beta peptide, Proceedings of the National Academy of Sciences 106 (2009) 20324–20329. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [61].Hashimoto M, Bogdanovic N, Volkmann I, Aoki M, Winblad B, Tjernberg LO, Analysis of microdissected human neurons by a sensitive ELISA reveals a correlation between elevated intracellular concentrations of Aβ42 and Alzheimer?s disease neuropathology, Acta neu-ropathologica 119 (2010) 543–554. [DOI] [PubMed] [Google Scholar]
  • [62].Toglia P, Ullah G, The gain-of-function enhancement of IP3-receptor channel gating by familial Alzheimer0s disease-linked presenilin mutants increases the open probability of mitochondrial permeability transition pore, Cell calcium 60 (2016) 13–24. [DOI] [PubMed] [Google Scholar]
  • [63].Cortassa S, Aon MA, Marbán E, Winslow RL, O’Rourke B, An integrated model of cardiac mitochondrial energy metabolism and calcium dynamics, Biophys. J 84 (2003) 2734–2755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [64].Cortassa S, Aon MA, Winslow RL, O’Rourke B, A mitochondrial oscillator dependent on reactive oxygen species, Biophys. J 87 (2004) 2060–2073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [65].Magnus G, Keizer J, Minimal model of beta-cell mitochondrial Ca2+ handling, Am. J. Physiol. - Cell Physiol 273 (1997) C717–C733. [DOI] [PubMed] [Google Scholar]
  • [66].Nguyen M-HT, Jafri MS, Mitochondrial calcium signaling and energy metabolism, Ann. N. Y. Acad. Sci 1047 (2005) 127–137. [DOI] [PubMed] [Google Scholar]
  • [67].Jafri MS, Kotulska M, Modeling the mechanism of metabolic oscillations in ischemic cardiac myocytes, J. Theor. Biol 242 (2006) 801–817. [DOI] [PubMed] [Google Scholar]
  • [68].Nguyen M-HT, Dudycha SJ, Jafri MS, Effect of Ca2+ on cardiac mitochondrial energy production is modulated by Na+ and H+ dynamics, Am. J. Physiol. - Cell Physiol 292 (2007) C2004–C2020. [DOI] [PubMed] [Google Scholar]
  • [69].Fall CP, Keizer JE, Mitochondrial modulation of intracellular Ca2+ signaling, J. Theor. Biol 210 (2001) 151–165. [DOI] [PubMed] [Google Scholar]
  • [70].Oster AM, Thomas B, Terman D, Fall CP, The low conductance mitochondrial permeability transition pore confers excitability and CICR wave propagation in a computational model, J. Theor. Biol 273 (2011) 216–231. [DOI] [PubMed] [Google Scholar]
  • [71].Diekman CO, Fall CP, Lechleiter JD, Terman D, Modeling the neuroprotective role of enhanced astrocyte mitochondrial metabolism during stroke, Biophys. J 104 (2013) 1752–1763. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [72].Wacquier B, Combettes L, Van Nhieu GT, Dupont G, Interplay between intracellular Ca2+ Oscillations and Ca2+-stimulated mitochondrial metabolism, Scientific reports 6 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [73].Wacquier B, Romero Campos HE, González-Vélez V, Combettes L, Dupont G, Mitochondrial Ca2+ dynamics in cells and suspensions, The FEBS journal (2017). [DOI] [PubMed] [Google Scholar]
  • [74].Qi H, Li L, Shuai J, Optimal microdomain crosstalk between endoplasmic reticulum and mitochondria for Ca2+ oscillations, Sci. Rep 5 (2015) 7984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [75].Dupont G, Modeling the intracellular organization of calcium signaling, Wiley Interdisciplinary Reviews: Systems Biology and Medicine 6 (2014) 227–237. [DOI] [PubMed] [Google Scholar]
  • [76].Bezprozvanny I, Bezprozvannaya S, Ehrlich B, Caffeine-induced inhibition of inositol (1, 4, 5)-trisphosphate-gated calcium channels from cerebellum., Molecular Biology of the Cell 5 (1994) 97–103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [77].Brown G, Sayers L, Kirk C, Michell R, Michelangeli F, The opening of the inositol 1, 4, 5-trisphosphate-sensitive Ca2+ channel in rat cerebellum is inhibited by caffeine, Biochemical Journal 282 (1992) 309–312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [78].Demuro A, Parker I, Picomolar sensitivity to inositol trisphosphate in xenopus oocytes, Cell calcium 58 (2015) 511–517. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [79].Parker I, Ivorra I, Caffeine inhibits inositol trisphosphate-mediated liberation of intracellular calcium in xenopus oocytes., The Journal of Physiology 433 (1991) 229–240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [80].Taylor CW, Tovey SC, IP3 receptors: toward understanding their activation, Cold Spring Harbor perspectives in biology (2010) a004010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [81].Shah SI, Demuro A, Mak D-OD, Parker I, Pearson JE, Ullah G, Trace-specks: a software for automated idealization of noisy patch-clamp and imaging data, Biophysical. J 115 (2018) 9?21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [82].Cheung K-H, Shineman D, Müller M, Cardenas C, Mei L, Yang J, Tomita T, Iwatsubo T, Lee VM-Y, Foskett JK, Mechanism of Ca2+ disruption in Alzheimer0s disease by presenilin regulation of IP3R receptor channel gating, Neuron 58 (2008) 871–883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [83].Cheung K-H, Mei L, Mak D-OD, Hayashi I, Iwatsubo T, Kang DE, Foskett JK, Gain-of-function enhancement of IP3 receptor modal gating by familial Alzheimer?s disease–linked presenilin mutants in human cells and mouse neurons, Sci. Signal 3 (2010) ra22–ra22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [84].Mukherjee R, Mareninova OA, Odinokova IV, Huang W, Murphy J, Chvanov M, Javed MA, Wen L, Booth DM, Cane MC, et al. , Mechanism of mitochondrial permeability transition pore induction and damage in the pancreas: inhibition prevents acute pancreatitis by protecting production of ATP, Gut (2016) 1333–1346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [85].Rizzuto R, De Stefani D, Raffaello A, Mammucari C, Mitochondria as sensors and regulators of calcium signalling, Nat. Rev. Mol. Cell Biol 13 (2012) 566–578. [DOI] [PubMed] [Google Scholar]
  • [86].Murgia M, Giorgi C, Pinton P, Rizzuto R, Controlling metabolism and cell death:at the heart of mitochondrial calcium signalling, Journal of molecular and cellular cardiology 46 (2009) 781–788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [87].Kowaltowski AJ, Vercesi AE, Mitochondrial damage induced by conditions of oxidative stress, Free Radic. Biol. Med 26 (1999) 463–471. [DOI] [PubMed] [Google Scholar]
  • [88].Carraro M, Bernardi P, Calcium and reactive oxygen species in regulation of the mitochondrial permeability transition and of programmed cell death in yeast, Cell calcium 60 (2016) 102–107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [89].Tiwari BS, Belenghi B, Levine A, Oxidative stress increased respiration and generation of reactive oxygen species, resulting in ATP depletion, opening of mitochondrial permeability transition, and programmed cell death, Plant Physiol. 128 (2002) 1271–1281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [90].Lopez JR, Lyckman A, Oddo S, LaFerla FM, Querfurth HW, Shtifman A, Increased intraneuronal resting [Ca2+] in adult Alzheimer?s disease mice, Journal of neurochemistry 105 (2008) 262–271. [DOI] [PubMed] [Google Scholar]
  • [91].Clapham DE, Calcium signaling, Cell 131 (2007) 1047–1058. [DOI] [PubMed] [Google Scholar]
  • [92].Cárdenas C, Miller RA, Smith I, Bui T, Molgü J, Müller M, Vais H, Cheung K-H, Yang J, Parker I, et al. , Essential regulation of cell bioenergetics by constitutive Insp3 receptor Ca2+ transfer to mitochondria, Cell 142 (2010) 270–283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [93].Supnet C, Bezprozvanny I, Neuronal calcium signaling, mitochondrial dysfunction, and Alzheimers disease, J. Alzheimers Dis 20 (2010) 487–498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [94].Zampese E, Fasolato C, Kipanyula MJ, Bortolozzi M, Pozzan T, Pizzo P, Presenilin 2 modulates Endoplasmic Reticulum (ER)–mitochondria interactions and Ca2+ cross-talk, Proc. Natl. Acad. Sci 108 (2011) 2777–2782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [95].Golpich M, Amini E, Mohamed Z, Azman Ali R, Mohamed Ibrahim N, Ahmadiani A, Mitochondrial dysfunction and biogenesis in neurodegenerative diseases: pathogenesis and treatment, CNS Neurosci. Ther (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [96].Onyango IG, Lu J, Rodova M, Lezi E, Crafter AB, Swerdlow RH, Regulation of neuron mitochondrial biogenesis and relevance to brain health, BBA Mol. Bas 1802 (2010) 228–234. [DOI] [PubMed] [Google Scholar]
  • [97].Hashimoto M, Rockenstein E, Crews L, Masliah E, Role of protein aggregation in mitochondrial dysfunction and neurodegeneration in Alzheimer0s and Parkinson’s diseases, Neuromolecular Med. 4 (2003) 21–35. [DOI] [PubMed] [Google Scholar]
  • [98].Coskun PE, Wyrembak J, Derbereva O, Melkonian G, Doran E, Lott IT, Head E, Cotman CW, Wallace DC, Systemic mitochondrial dysfunction and the etiology of Alzheimers disease and down syndrome dementia, J. Alzheimers Dis 20 (2010) S293–S310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [99].Busciglio J, Pelsman A, Wong C, Pigino G, Yuan M, Mori H, Yankner BA, Altered metabolism of the amyloid β precursor protein is associated with mitochondrial dysfunction in down’s syndrome, Neuron 33 (2002) 677–688. [DOI] [PubMed] [Google Scholar]
  • [100].Li PA, Hou X, Hao S, Mitochondrial biogenesis in neurodegeneration, J. Neurosci. Res (2017). [DOI] [PubMed] [Google Scholar]
  • [101].Dragicevic N, Mamcarz M, Zhu Y, Buzzeo R, Tan J, Arendash GW, Bradshaw PC, Mitochondrial amyloid-β levels are associated with the extent of mitochondrial dysfunction in different brain regions and the degree of cognitive impairment in Alzheimers transgenic mice, J. Alzheimers Dis 20 (2010) S535–S550. [DOI] [PubMed] [Google Scholar]
  • [102].Cadonic C, Sabbir MG, Albensi BC, Mechanisms of mitochondrial dysfunction in Alzheimers disease, Mol. Neurobiol 53 (2016) 6078–6090. [DOI] [PubMed] [Google Scholar]
  • [103].Swerdlow RH, Khan SM, A mitochondrial cascade hypothesis for sporadic Alzheimer′s disease, Med. Hypotheses 63 (2004) 8–20. [DOI] [PubMed] [Google Scholar]
  • [104].Coskun PE, Busciglio J, Oxidative stress and mitochondrial dysfunction in down’s syndrome: relevance to aging and dementia, Curr. Gerontol. Geriatr. Res 2012. (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [105].Cao P, Falcke M, Sneyd J, Mapping interpuff interval distribution to the properties of inositol trisphosphate receptors, Biophys. J 112 (2017) 2138–2146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [106].Ullah G, Ullah A, Mode switching of inositol 1,4,5-trisphosphate receptor channel shapes the spatiotemporal scales of Ca2+ signals, J. Biol. Phys 42 (2016) 507–524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [107].Yao J, Irwin RW, Zhao L, Nilsen J, Hamilton RT, Brinton RD, Mitochondrial bioenergetic deficit precedes alzheimer’s pathology in female mouse model of alzheimer’s disease, Proceedings of the National Academy of Sciences 106 (2009) 14670–14675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [108].Wang X, Wang W, Li L, Perry G, Lee H.-g., Zhu X, Oxidative stress and mitochondrial dysfunction in Alzheimer′s disease, BBA Mol. Bas 1842 (2014) 1240–1247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [109].Resende R, Moreira PI, Proenc T¸Deshpande a, A., Busciglio J, Pereira C, Oliveira CR, Brain oxidative stress in a triple-transgenic mouse model of Alzheimer disease, Free Radic. Biol. Med 44 (2008) 2051–2057. [DOI] [PubMed] [Google Scholar]
  • [110].Simakova O, Arispe NJ, The cell-selective neurotoxicity of the Alzheimers Aβ peptide is determined by surface phosphatidylserine and cytosolic ATP levels. membrane binding is required for Aβ toxicity, J. Neurosci 27 (2007) 13719–13729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [111].Kuchibhotla KV, Goldman ST, Lattarulo CR, Wu H-Y, Hyman BT, Bacskai BJ, Aβ plaques lead to aberrant regulation of calcium homeostasis in vivo resulting in structural and functional disruption of neuronal networks, Neuron 59 (2008) 214–225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [112].Wang X, Su B, Siedlak SL, Moreira PI, Fujioka H, Wang Y, Casadesus G, Zhu X, Amyloid-β overproduction causes abnormal mitochondrial dynamics via differential modulation of mitochondrial fission/fusion proteins, Proc. Natl. Acad. Sci 105 (2008) 19318–19323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [113].Zhu X, Perry G, Smith MA, Wang X, Abnormal mitochondrial dynamics in the pathogenesis of Alzheimers disease, J. Alzheimers Dis 33 (2013) S253–S262. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [114].Lin MT, Beal MF, Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases, Nature 443 (2006) 787. [DOI] [PubMed] [Google Scholar]
  • [115].Iijima-Ando K, Hearn SA, Shenton C, Gatt A, Zhao L, Iijima K, Mitochondrial mislocalization underlies Aβ42-induced neuronal dysfunction in a drosophila model of Alzheimer′s disease, PLoS One 4 (2009) e8310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [116].Hirai K, Aliev G, Nunomura A, Fujioka H, Russell RL, Atwood CS, Johnson AB, Kress Y, Vinters HV, Tabaton M, et al. , Mitochondrial abnormalities in Alzheimers disease, J. Neurosci 21 (2001) 3017–3023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [117].Baloyannis SJ, Mitochondrial alterations in Alzheimers disease, J. Alzheimers Dis 9 (2006) 119–126. [DOI] [PubMed] [Google Scholar]
  • [118].Calkins MJ, Manczak M, Mao P, Shirendeb U, Reddy PH, Impaired mitochondrial biogenesis, defective axonal transport of mitochondria, abnormal mitochondrial dynamics and synaptic degeneration in a mouse model of Alzheimer′s disease, Hum. Mol. Gen 20 (2011) 4515–4529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [119].Miller KE, Sheetz MP, Axonal mitochondrial transport and potential are correlated, J. Cell Sci 117 (2004) 2791–2804. [DOI] [PubMed] [Google Scholar]
  • [120].Zhang CL, Ho PL, Kintner DB, Sun D, Chiu SY, Activity-dependent regulation of mitochondrial motility by calcium and Na/K-ATPase at nodes of ranvier of myelinated nerves, J. Neurosci 30 (2010) 3555–3566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [121].Hollenbeck PJ, Saxton WM, The axonal transport of mitochondria, J. Cell Sci 118 (2005) 5411–5419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [122].Helguera P, Seiglie J, Rodriguez J, Hanna M, Helguera G, Busciglio J, Adaptive downregulation of mitochondrial function in down syndrome, Cell Metab. 17 (2013) 132–140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [123].Van der Bliek AM, Shen Q, Kawajiri S, Mechanisms of mitochondrial fission and fusion, Cold Spring Harb. Perspect. Biol 5 (2013) a011072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [124].Knott AB, Perkins G, Schwarzenbacher R, Bossy-Wetzel E, Mitochondrial fragmentation in neurodegeneration, Nature Rev. Neurosci 9 (2008) 505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [125].Yu T, Robotham JL, Yoon Y, Increased production of reactive oxygen species in hyperglycemic conditions requires dynamic change of mitochondrial morphology, Proc. Natl. Acad. Sci 103 (2006) 2653–2658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [126].Chan DC, Mitochondria: dynamic organelles in disease, aging, and development, Cell 125 (2006) 1241–1252. [DOI] [PubMed] [Google Scholar]
  • [127].Pierrot N, Ghisdal P, Caumont A-S, Octave J-N, Intraneuronal amyloid-β1–42 production triggered by sustained increase of cytosolic calcium concentration induces neuronal death, Journal of neurochemistry 88 (2004) 1140–1150. [DOI] [PubMed] [Google Scholar]
  • [128].Rizzuto R, Brini M, Murgia M, Pozzan T, Microdomains with high Ca2+ close to IP3-sensitive channels that are sensed by neighboring mitochondria, Science 262 (1993) 744–747. [DOI] [PubMed] [Google Scholar]
  • [129].Kornmann B, Currie E, Collins SR, Schuldiner M, Nunnari J, Weissman JS, Walter P, An ER-mitochondria tethering complex revealed by a synthetic biology screen, Science 325 (2009) 477–481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [130].Ullah G, Jung P, Modeling the statistics of elementary calcium release events, Bio-phys. J 90 (2006) 3485–3495. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [131].Gee KR, Brown K, Chen WU, Bishop-Stewart J, Gray D, Johnson I, Chemical and physiological characterization of fluo-4 Ca2+-indicator dyes, Cell calcium 27 (2000) 97–106. [DOI] [PubMed] [Google Scholar]
  • [132].Dargan SL, Parker I, Buffer kinetics shape the spatiotemporal patterns of IP3-evoked Ca2+ signals, J. Physiol 553 (2003) 775–788. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1

RESOURCES