干细胞之家 - 中国干细胞行业门户第一站

 

 

搜索
朗日生物

免疫细胞治疗专区

欢迎关注干细胞微信公众号

  
查看: 442906|回复: 259
go

Caspase-mediated loss of mitochondrial function and generation of reactive oxyge [复制链接]

Rank: 7Rank: 7Rank: 7

积分
威望
0  
包包
792  
楼主
发表于 2009-3-6 09:25 |只看该作者 |倒序浏览 |打印
1 Division of Cellular Immunology, La Jolla Institute for Allergy and Immunology, San Diego, CA 92121
# l, l  Q4 |6 j: o8 {1 m0 f0 M1 N  Y
2 Division of Hematology, Department of Molecular and Experimental Medicine, The Scripps Research Institute, MEM220, La Jolla, CA 92037
' K1 E+ W/ ?  ~1 S: M1 C
6 }! p. F2 x! U) _Abstract
. Y+ W. A/ Q5 a2 d) H7 b6 H8 A2 n( i' s) y" |. ~
During apoptosis, the permeabilization of the mitochondrial outer membrane allows the release of cytochrome c, which induces caspase activation to orchestrate the death of the cell. Mitochondria rapidly lose their transmembrane potential  and generate reactive oxygen species (ROS), both of which are likely to contribute to the dismantling of the cell. Here we show that both the rapid loss of m and the generation of ROS are due to the effects of activated caspases on mitochondrial electron transport complexes I and II. Caspase-3 disrupts oxygen consumption induced by complex I and II substrates but not that induced by electron transfer to complex IV. Similarly, m generated in the presence of complex I or II substrates is disrupted by caspase-3, and ROS are produced. Complex III activity measured by cytochrome c reduction remains intact after caspase-3 treatment. In apoptotic cells, electron transport and oxygen consumption that depends on complex I or II was disrupted in a caspase-dependent manner. Our results indicate that after cytochrome c release the activation of caspases feeds back on the permeabilized mitochondria to damage mitochondrial function (loss of m) and generate ROS through effects of caspases on complex I and II in the electron transport chain.
0 @+ w3 S/ q$ O% w, _
) t0 L3 ^. A& e7 ?- ZKey Words: apoptosis; mitochondria; caspases; transmembrane potential; ROS  ^( T* z# O/ D0 @. X8 m

3 u4 Z" P; P# H  R9 M! R) t; rIntroduction7 R7 R7 J& a$ h; h
. }2 O+ i1 f* {7 \, a5 c$ m
The activation of caspase proteases is fundamental to apoptotic cell death, although their mode of action in promoting death is not fully understood. In vertebrate cells, inhibition of caspases does not necessarily prevent cell death but profoundly delays and alters the process . Activated caspases, especially caspase-3, orchestrate DNA fragmentation , nuclear condensation and membrane blebbing  through cleavage of specific substrates, but these are unlikely to fully account for the rapidity of caspase-dependent death once the proteases are activated. Here we will explore the impact of caspases on another critical compartment in dying cells, the mitochondria.2 B! H4 s, x9 a$ i* u' _

2 }# [  b! J7 S" Y& R+ VThe central role of mitochondria in the process of apoptosis has been a focus of cell death research since the observations that the antiapoptotic Bcl-2 protein localizes to the outer membrane of this organelle, a mitochondria-rich fraction was required for the induction of apoptotic changes in a cell-free system , and mitochondrial transmembrane potential is lost during an early stage of apoptosis . Over the past several years, it has become clear that a major event during apoptosis is the permeabilization of the mitochondrial outer membrane to release proteins from the intermembrane space. Several of these, including cytochrome c, AIF, Smac/DIABLO, Omi/Htra2, and EndoG, have roles in subsequent cell death . In particular, the release of cytochrome c induces the activation of caspase proteases through the induction of apoptosome formation .
1 e% W" _# T* B* i1 L
: ?6 J7 W7 }! yMitochondrial functions including protein import, ATP generation, and lipid biogenesis depend on the maintenance of m , and loss of m during apoptosis is likely to contribute to the death of the cell through loss of these functions. In addition, mitochondrial production of reactive oxygen species (ROS) also appears to play a role in cell death . The relationships between these events, release of mitochondrial proteins, and caspase activation remain controversial. Although models of mitochondrial function during apoptosis often predict hypo- or hyperpolarization of the inner membrane before outer membrane permeabilization , we have found that in the absence of caspase activation m does not necessarily change before and remains intact after this event . Single cell analysis in HeLa and other cells provided evidence that a persistent loss of m rapidly follows cytochrome c release only when caspases are activated, and otherwise this loss follows a variable (and slow) kinetics. The maintenance of m under these conditions appears to be via electron transport supported by the cytochrome c diffusely available in the cytosols of the cells that had undergone mitochondrial outer membrane permeabilization. In HeLa cells, loss of m corresponds to a rapid decline in ATP levels before cell death, and this is profoundly enhanced by caspase activation.6 D7 |+ O8 `6 V0 |

: e' D2 a3 J4 n7 d6 y6 RHere we explore the role of caspase activation in loss of m and generation of ROS during apoptosis. Although caspase-3 can cause permeabilization of the mitochondrial outer membrane, this is at least partially dependent on the function of the proapoptotic Bcl-2 family protein Bid and is blocked by Bcl-xL. However, the caspase then has a further effect on the mitochondria through disruption of the functions of complex I and II of the electron transport chain, resulting in loss of m and generation of ROS. This rapid effect of caspases on the function of the electron transport chain is therefore likely to be a major contributing factor to the process of caspase-dependent cell death.2 C* W; e& v; E4 r& C
, @2 J' ^! }4 b$ ?; X) C1 Z  h
Results9 m1 q3 C0 m4 p' B

& D5 }! |, k6 L/ FTo examine the role of caspases in mitochondrial functions during apoptosis, we first examined the effects of their inhibition on two important parameters, m (using tetramethylrhodamine ethyl ester [TMRE] ) and the generation of ROS (using dihydroethidium [2-HE] ). HeLa cells were treated with actinomycin D (ActD) or UV to induce apoptosis, which was assessed by annexin V–FITC/propidium iodide (PI) staining. Cell death, loss of m, and production of ROS showed a close correspondence in each case. Addition of the caspase inhibitor N-benzoylcarbanyl-Val-Ala-Asp-fluoro methylketone (zVAD-fmk) delayed both cell death and loss of m as observed previously . Interestingly, inhibition of caspase activation also blocked ROS production. Therefore, caspase functions appeared to be required for the rapid onset of these events during apoptosis.' v7 G: m8 R( |! ]. ~1 {
$ X) k+ b! r) t& N
fig.ommtted; @) o! X5 P, F9 L& Q
9 U: g3 I8 y5 ?% c5 c% h
Apoptosis, loss of m, and ROS production are caspase dependent. HeLa cells were treated with actinomycin D or UV (as indicated) in the presence or absence of 50 µM of zVAD-fmk, harvested, and stained with annexin V–FITC and PI to assess cell death, TMRE to measure m, and 2-HE to measure ROS production, and then analyzed by flow cytometry.0 K" d1 t2 H; s1 i  C" ]" f
6 S. N8 L  }" k/ B" [
During apoptosis, the mitochondrial outer membrane becomes permeable due to the action of pro-apoptotic Bcl-2 family proteins. Using isolated mitochondria, we asked whether such outer membrane permeabilization, with or without caspase activity, is sufficient to account for the loss of m in mitochondria during apoptosis. m was examined by uptake of TMRE  The uncoupler, carbonyl cyanide p-(trifluoromethoxy) phenylhydrazone (FCCP), caused a dissipation of m as expected. To induce outer membrane permeabilization as it occurs during apoptosis, we treated the mitochondria with the activated BH3-only protein, truncated Bid (tBid) (caspase-free; see Materials and methods), which caused a rapid release of cytochrome c ( B, inset) but did not disrupt m  This effect required that the mitochondria be maintained at high density (as in the experiment shown), since dilution of the treated organelles caused a loss of m as the cytochrome c is diluted. Although apoptosis in most cases does not depend on the action of tBid , we employed this protein as a model, since it is likely that the mitochondrial permeabilization induced by tBid is similar to that occurring during many forms of apoptosis (. Treatment of the isolated mitochondria with caspase-3 also failed to affect m . However, addition of both tBid and caspase-3 caused a loss of m, and this effect was dependent on the caspase activity as zVAD-fmk inhibited the effect. Since tBid induces the permeabilization of the mitochondrial outer membrane , it is therefore likely that it acts to permit caspase-3 access to the intermembrane space, necessary for the protease to affect m. Although the mechanism is unknown, tBid can cause the outer mitochondrial membrane to become permeable to dextrans of higher molecular weight than that of active caspase-3' I6 G  i; a# ~- G

3 F7 }. ?3 E9 t9 v$ }- W$ }( Vfig.ommtted
" u2 x) p+ Z) D# s7 M" W$ b, [) Y6 X4 V6 i' ~* @' R
Caspase-3 induces loss of m in isolated, tBid-treated mitochondria. Mouse liver mitochondria (20 µg) were incubated with FCCP (10 µM), tBid (20 µg/ml), cytochrome c (100 µM), caspase-3 (0.25 µg/ml), and/or zVAD-fmk (100 µM) as indicated, then stained with TMRE, and analyzed by flow cytometry. Low fluorescence indicates a loss of m.
4 T# \& {% x: @2 u/ g, q( h8 _5 T  j4 V0 |) U8 z' i1 ]
fig.ommtted/ k) x9 Y" F( G! \. i! T

4 _: U& Z, d+ RCaspase-3 inhibits oxygen uptake in response to substrates for complex I or II. (A) Simplified scheme of the respiratory chain. Briefly, the electron transport chain is composed of four multisubunit complexes referred to as complexes I–IV. The electrons enter the system through either complex I (NADH-ubiquinol oxidoreductase) or complex II (Succinate-ubiquinol dehydrogenase). The electrons are passed to complex III via ubiquinone (Q). Complex III (ubiquinol/cytochrome c oxido-reductase) transfers electrons to cytochrome c (C), which in turn donates them to complex IV (cytochrome oxidase). Complex IV will finally transfer 4 electrons to a molecular oxygen leading to the generation of H2O. Upon this transfer of electrons, complexes I, III, and IV pump protons out of the matrix leading to the generation of m. Complex V (F0F1-ATP synthase) uses this potential to convert ADP into ATP. OM, outer membrane; IM, inner membrane; IMS, intermembrane space. Complex substrates: malate/palmitate for complex I, succinate for complex II, TMPD/ascorbate for cytochrome c. Several respiratory inhibitors used in our studies are indicated: rotenone (complex I inhibitor), antimycin A (complex III inhibitor), KCN (complex IV inhibitor), and oligomycin (complex V inhibitor). (B–E) Mouse liver mitochondria (400 µg) were incubated in dual oxygen electrode chambers in the presence of 100 µM of cytochrome c, with in control or tBid (20 µg/ml) (B), in control or caspase-3 (20 µg/ml) (C), in tBid (20 µg/ml) with or without caspase-3 (20 µg/ml) (D), and in 100 µM Ca2  with or without caspase-3 (20 µg/ml) (E). Then drugs are added in the following order: malate/palmitate (2.5 mM and 40 µM, respectively), 2 mM ADP, 2 µM rotenone, 5 mM succinate 1 µM antimycin A, 0.4 mM TMPD with 1 mM ascorbate and 1 mM KCN. Values for oxygen consumption are represented on the curves as ng-atoms of oxygen/min/mg of protein. Western blots in , represent the amount of cytochrome c or hsp60 present in the mitochondrial pellet.
1 C2 @( _2 v- w; [9 I; {% m$ T* A3 O6 e9 M  ]
Mitochondrial m is generated by the components of the electron transport chain, which consume O2 and pump protons across the mitochondrial inner membrane to produce ATP . Addition of specific substrates for each complex together with inhibitors for upstream complexes can drive respiration that can be detected using an oxygen (Clarke) electrode. As shown in  B, untreated and tBid-treated mitochondria consumed oxygen upon addition of substrates for complex I (malate/o-palmitoyl-L-carnitine), complex II (succinate), or cytochrome c (tetrametyl-p-phenylenediamine [TMPD]/ascorbate; complex IV then consumes oxygen), and this respiration was stopped by the appropriate inhibitors in each case. Nevertheless, tBid permeabilized the mitochondrial outer membrane, seen as a loss of pellet-associated cytochrome c but not matrix HSP60 inset). Similarly, untreated and caspase-3–treated mitochondria displayed similar oxygen uptake upon stimulation . In each of these experiments (, B–E), the respiration was measured in the same mitochondria under two conditions in parallel (see Materials and methods), and therefore the pair in each figure are directly comparable (in contrast, differences in the mitochondrial preparations produced some differences in the extent of respiration between separate experiments, although the control patterns were qualitatively comparable).2 Y5 x% @5 b( X% z7 K. K$ q7 ]% F
3 m  ], L3 w0 w3 G3 F3 d
In contrast, mitochondria treated with tBid plus caspase-3 consumed no oxygen in response to complex I substrates, a difference that was more pronounced in the presence of ADP. Similarly, tBid plus caspase-3–treated mitochondria showed no increase in oxygen consumption in response to the complex II substrate succinate. Caspase-3 treatment resulted in an 88% inhibition of complex I function and a 94% inhibition of the oxygen consumption by complex II. In contrast, respiration via complex IV was similar in tBid-treated mitochondria with or without caspase-3 treatment.
) B, m6 k/ @9 }" }% ?5 {9 d, h4 b$ y. i4 g1 v) D
The effect of caspase-3 on mitochondrial respiration was not dependent on the use of tBid. Treatment of mitochondria with calcium induces a permeability transition that causes the matrix to swell and ultimately rupture the mitochondrial outer membrane , and this was seen as a decrease in pellet-associated cytochrome c but not matrix HSP60 ( inset). It has been suggested that this effect occurs during apoptosis to result in loss of m and release of intermembrane proteins from mitochondria . Unlike tBid, disruption of the mitochondrial outer membrane by calcium treatment had some inhibitory effect on respiration, particularly complex I function. This effect of calcium on complex I has been described . Nevertheless, treatment of calcium-permeabilized mitochondria with caspase-3 caused a loss of complex I and II activity. The simplest interpretation of these results is that caspase-3 enters permeabilized mitochondria and then acts to disrupt respiration by targeting proteins that are exposed to the intermembrane space.9 G/ y- E4 F" i% ^% J
& W" w& N3 J/ L' ]
Although we observed intact complex IV activity after caspase treatment of mitochondria, the reduction in oxygen consumption in response to substrates for complex I or II might nevertheless be due to a loss of function of complex III . Therefore, to assess complex III function we examined the ability of accessible complex III to reduce cytochrome c . As shown in , intact mitochondria did not reduce exogenously added cytochrome c, whereas tBid-treated mitochondria did . This effect was dependent on complex III activity, since the inhibitor antimycin A blocked cytochrome c reduction in this system. Permeabilized mitochondria (by tBid or by hypotonic lysis) treated with caspase-3 displayed full complex III activity in this assay, and thus the function of complex III (at least that of cytochrome c reduction) was not damaged by caspase-3. These results support the idea that caspases damage the function of complexes I and II without affecting those of complex III or complex IV.6 f. V; R1 x" _6 \# I

2 t4 {3 [* z- _/ c4 L0 ?$ J( dfig.ommtted
" D4 A( y& J6 v# g/ p# y+ K: R- J3 X+ W: L+ T0 [0 a
Caspase-3 does not destroy the capacity of complex III to reduce cytochrome c. Isolated mitochondria (500 µg) were incubated in the presence of tBid (25 µg/ml), caspase-3 (25 µg/ml), and/or antimycin A (1 µM) as indicated for 60 min at 37°C. As a control, 500 µg of mitochondria were incubated in water for 20 min at 4°C then incubated ± caspase-3. Reduction of exogenous cytochrome c by complex III was measured as described in Materials and methods.
; {1 C2 e; s. L& Z" l  s8 j5 k8 X* H+ c% J! {
Caspase-3 can cleave and activate Bid and in this manner trigger mitochondrial outer membrane permeabilization  and it was therefore likely that treatment of digitonin-treated cells with caspase-3 would induce cytochrome c release. Therefore, we compared Bid /  and Bid-/- cells to further assess the possible role of tBid in the dissipation of m induced by caspase-3. We observed previously that activated Bid (tBid) can induce loss of m in digitonin-permeabilized cells, and this was restored by addition of exogenous cytochrome c  consistent with the lack of effect of tBid on electron transport function we have observed here. In the experiment shown in  A, caspase-3 induced a loss of m but only in cells containing Bid; Bid-/- cells did not show a drop in m in response to caspase-3 treatment. This is consistent with our observations that caspase-3 did not act directly on isolated mitochondria to disrupt m or respiration . Addition of tBid plus caspase-3 induced a loss of m in both wild-type and Bid-/- cells, which was not restored by addition of cytochrome c. Thus, caspase-3 appeared to have two effects: activation of Bid to permeabilize the outer membrane without disrupting m, as described , and an action on the permeabilized mitochondria to disrupt m.. A8 ^" G1 ?; k. `% z
$ e1 \6 l. L+ ~7 N) u; L0 L2 M9 ?9 F# l
fig.ommtted7 }& O- ~2 \) o" V

% X- `) \. g; H0 C* [Caspase-3 treatment of permeabilized cells causes loss of m. (A) MEFs from wild-type or Bid-/- cells were permeabilized with digitonin and incubated in the presence of caspase-3 (0.5 µg/ml), tBid (20 µg/ml), and/or cytochrome c (100 µM) as indicated plus TMRE at 37°C for 30 min in the presence of succinate as substrate. (B–E) Permeabilized HeLa cells (106) were incubated with or without caspase-3 (0.5 µg/ml) in the presence of cytochrome c (100 µM) and TMRE with or without zVAD-fmk (100 µM) or BclXL-c (20 µg/ml) as indicated. (B) m generated by the incubation of permeabilized cells with the substrates. (C) Effect of caspase-3 on m in the presence of malate/palmitate. (D) Effect of caspase-3 on m in the presence of rotenone and succinate. (E) Effect of caspase-3 on m in the presence of antimycin A and TMPD/ascorbate. Substrates and inhibitors were added at the concentrations described in the legend to  B. Cells were analyzed for  by flow cytometry. In each case, the MFI for cells treated with FCCP to dissipate m was set as 0.
( W& H: `+ k$ Q6 F' l- W9 a' y$ j& y; p+ H5 _
To extend these findings, we examined m in digitonin-permeabilized cells treated with caspase-3 as in  A. However, in this case electron transport was driven by substrate/inhibitor combinations to assess the contribution to m of each complex . Complexes I, III, and IV pump protons to generate m upon electron transport, and under our conditions each complex contributes to this potential directly (complex I and IV) and/or through downstream electron transport events (complex I and II). As shown in  caspase-3 destroyed the contribution to m from complex I or complex II but had no effect on the function of complex IV. The effect of caspase-3 on the functions of complex I and II were inhibited by the addition of zVAD-fmk or/and Bcl-xL-C (. The effect of Bcl-xL is again consistent with an inability of caspase-3 to influence mitochondrial function without access to the intermembrane space, since Bcl-xL blocks mitochondrial membrane permeabilization by Bid and other proapoptotic Bcl-2 family members ./ l$ D9 W5 Z. ]  h7 s7 |

" V4 P5 x1 r+ F0 O! mOne consequence of a caspase-mediated disruption in electron transport may be the zVAD-fmk–inhibitable generation of ROS discussed above. Therefore, we examined if substrates for complex I  or complex II  drive caspase-dependent ROS generation in digitonin-permeabilized Jurkat cells. Addition of substrates for complexes I or II fueled the production of ROS in untreated mitochondria, and this was not increased by treatment with tBid. In contrast, treatment with caspase-3 (with or without addition of recombinant tBid) resulted in significant ROS production with either substrate  (but not without substrates;  A). The increase was inhibited by Bcl-xL-C, probably via inhibition of the caspase-activated, Bid-mediated permeabilization of the mitochondrial outer membrane as discussed above. Therefore, it is likely that the caspase-mediated disruption of complex I and complex II function contributes to high ROS production during apoptosis. This would account for the effect of caspase inhibition on apoptosis-associated ROS generation we observed in the experiment in .
: x# H8 f  v: g2 n  m' y  q
( U# @# v  s3 K6 o$ b$ I6 A3 mfig.ommtted
1 Z2 b$ c, J; _/ e' s6 Y" O! {3 }9 I( m. S
ROS production in the presence of malate/palmitate or succinate is enhanced by caspase-3. Jurkat cells (106) were permeabilized and incubated in the presence of dihydroethidium (2 µM) and caspase-3 (0.5 µg/ml), tBid (20 µg/ml), or Bcl-xL-C (20 µg/ml) for 30 min at 37°C. (A) Control; (B) in the presence of malate/palmitate, and (C) in the presence of succinate. Then the ROS production was monitored by flow cytometry. Note that the scales differ for each condition.; Y& E1 C, U' k! O) O, \5 k% [

" @5 e' P! w+ R2 t3 S# U" _Therefore, we sought to determine if complex I and II are targeted for caspase-mediated disruption during the process of apoptosis. HeLa cells were treated with actinomycin D to induce apoptosis, with and without zVAD-fmk to block caspase activation. Cells were then treated with digitonin in order to provide substrates with access to the mitochondria. Different substrates with or without exogenous cytochrome c were added, and m was assessed. As shown in , ActD-treated cells lost m driven by complex I, II, or IV substrates. In the absence of substrates, m was minimal (unpublished data) as in. Addition of cytochrome c restored complex IV activity in this system  but not the activities of complex I or II . This is consistent with our results with caspase-treated mitochondria and permeabilized cells, supporting the idea that mitochondrial outer membrane permeabilization and caspase activation during apoptosis disrupted the function of complex I and II. Consistent with this idea, inhibition of caspase activation protected m in each case.; s" t  a" b3 M5 j
6 j- ^) e2 ^/ C4 |5 J1 j. d; l+ L% A
fig.ommtted
" \' n2 {  h6 Y  x( H9 v6 S- u. H% c- n" H
Apoptotic cells maintain m in the presence of TMPD/ascorbate but not in the presence of malate/palmitate or succinate as substrates. HeLa cells were treated with or without 0.3 µM actinomycin D (18 h), permeabilized with digitonin, and incubated with the indicated substrates. Oligomycin (10 µg/ml) was added as indicated. Cells were stained with TMRE and analyzed by flow cytometry. (A) Malate/palmitate; (B) rotenone and succinate; (C) antimycin A and TMPD/ascorbate. Concentrations for inhibitors and substrates were the same as for the experiment in  B. Cells were analyzed for m by flow cytometry. In each case, the MFI for cells treated with FCCP to dissipate m was set as 0.
) H6 P/ ~3 d+ V1 s. g3 H
/ \5 ~* B; _  A: C. {( j/ `One way in which m can be maintained in the absence of electron transport is through ATP-dependent reversal of complex V (F0F1-ATPase) activity. Therefore, we examined the effects of the complex V inhibitor oligomycin under each of our conditions  Although oligomycin further depressed the reduced m in apoptotic cells under each of our conditions, it had no effect on m driven by complex IV substrates plus cytochrome c. This provides further support that complex IV remains functional after caspase activation in apoptotic cells. In the absence of caspase activation, addition of oligomycin did not dissipate m, consistent with our previous observations  . In addition, the observable effect of oligomycin in some cases in apoptotic cells  provides circumstantial evidence that complex V activity (at least that of the F1 component) remains intact after caspase activation; that is, ATP can drive m via proton pumping by the F1-ATPase.6 [4 ?5 G% S* A( ]" K
0 y5 a2 }1 N3 \, P  Z! ~
Oxygen consumption in apoptotic cells followed a similar pattern. Jurkat cells were treated with etoposide or staurosporine to induce apoptosis. The cells were then digitonin-permeabilized to provide access of substrates to the mitochondria. We found that oxygen consumption in the presence of complex I or complex II substrates was destroyed by the apoptotic process . This effect was caspase dependent, as it was blocked by the caspase inhibitor zVAD-fmk. In contrast, oxygen consumption by complex IV remained largely intact after caspase activation. However, there was a small but reproducible drop in complex IV activity that was seen in this assay, and this was also blocked by zVAD-fmk. This small decrease in complex IV activity (versus large decreases in those of complex I and II) was similarly seen in apoptotic HeLa cells . This small caspase-dependent effect is likely to be indirect, based on our results in isolated mitochondria  or may involve caspases other than caspase-3. Further, this may represent a small decrease in oxygen consumption without a decrease in m, since m did not decrease under the same conditions . Similar results were obtained in HeLa cells treated with staurosporine to induce apoptosis. Again, oxygen consumption in the presence of complex I or complex II substrates was destroyed by the apoptotic process, although the function of complex IV remained largely intact.
! b9 e- @" B! T4 [* f
$ |! u# M% k  t3 R3 _4 Efig.ommtted' y1 Y1 A/ U* c( S: O" {

" W3 m+ e4 `5 G# j7 S+ jFigure 8. Oxygen consumption in apoptotic cells. (A) Jurkat cells were treated for 18 h with etoposide (10 µM) or staurosporine (0.05 µM) in the presence or absence of 100 µM of zVAD-fmk as indicated. The cells were then permeabilized with digitonin. The equivalent of 400 µg of protein was loaded into respiratory chambers and oxygen consumption in the presence of different substrates and inhibitors (malate/palmitate for complex I, rotenone with succinate for complex II, antimycin A with TMPD/ascorbate for cytochrome c/complex IV) was measured as in the legend to  B. (B) HeLa cells were treated with staurosporine (0.05 µM) for 18 h, then treated with digitonin, and analyzed for oxygen consumption as in A.
5 b6 e4 `. Q2 W  b
3 x( l0 x* n" X8 hOur findings show that after mitochondrial outer membrane permeabilization and the activation of caspases, the caspases target complexes I and II of the electron transport chain. This results in a sustained loss of m and production of ROS, both of which may then contribute to the rapid dismantling of the cell.
9 b1 h3 B6 ^! n# P- _& d- |, y% A; q8 w  ], w1 I0 [1 b; e& w
Discussion
' V; _" `% ~/ u( h( |/ P( k9 j2 s
During apoptosis, proapoptotic members of the Bcl-2 family induce (and some may directly cause) permeabilization of the mitochondrial outer membrane . Although our studies often used only one of these, tBid, many other proapoptotic Bcl-2 family members have been implicated in different forms of apoptosis , and any of these is likely to trigger outer membrane permeabilization we have studied here. The release of proteins from the intermembrane space coincides with a transient drop in m, which in the absence of subsequent caspase activation can recover to normal levels for hours or even days thereafter . This maintenance of m is through electron transport, since it is inhibited by azide or cyanide (complex IV inhibitors) but not by oligomycin (complex V inhibitor) , and this depends on the cytochrome c that is now diffuse in the cell . The maintenance of m after cytochrome c release in the absence of caspase activation has been observed in diverse cell types under different conditions of apoptosis
6 b! k2 G* O0 B2 U, g( ?; J- |5 K  Z9 V, }' m6 V7 a8 @' C
Such observations suggested that caspases target mitochondrial function within minutes of cytochrome c release. In vitro, cytochrome c can induce apoptosome formation and caspase activation in ~ 10 min or less, consistent with this idea (Cohen, G., and K. Cain, personal communication). Since the highest concentration of cytosolic cytochrome c might be expected to be in the region of the mitochondria immediately after cytochrome c release, it is not unreasonable that caspase activation near mitochondria would be one of the first consequences of outer membrane permeabilization. Therefore, mitochondria may be among the earliest targets of caspase activation during apoptosis.
9 v/ S6 x' M( n' s+ h6 O4 e
1 T2 g* w6 ~7 w3 p, G1 ?  UIn the present study, we have analyzed the impact of caspases on the permeabilized mitochondria. Isolated intact mitochondria did not lose  in response to caspase-3 (Fig. 2), but mitochondria in which the outer membrane was permeable showed a disruption in complex I and II activities in response to caspase-3. In cells undergoing apoptosis via the mitochondrial pathway, mitochondrial outer membrane permeabilization occurs before caspase activation , and as noted above this event by itself does not interfere with the function of the electron transport chain unless caspases are subsequently activated. Analysis of this effect indicated that caspase-3 acts on the permeabilized mitochondria to disrupt m and respiration and induce ROS production via action on complexes I and II.
  C( }; u. c! X4 d0 |% k' S9 a8 _8 P5 ]8 _" e
In digitonin-treated cells, we observed that addition of active caspase caused a disruption of m and production of ROS, and this was blocked by Bcl-xL . This appeared to be largely dependent on the presence of Bid in the cells. Although in most cases engagement of the mitochondrial pathway is caspase independent (as noted above), Bid is activated by caspases and therefore can link other routes of caspase activation (e.g., death receptor signaling) to mitochondrial outer membrane permeabilization . Under such circumstances, caspase activation can precede cytochrome c release. However, our results would suggest that even in those cases, the disruptive effect on mitochondrial function would require both the permeabilization of the mitochondrial outer membrane followed by the action of the protease on the intermembrane space.. X; d* q. ^7 z$ l/ k! C0 S
* ~0 Q9 w( {* G2 o) K0 l1 O
Targeting of mitochondrial functions upon caspase activation has been described previously. In examining oxygen consumption during Fas-mediated apoptosis,  observed a loss of cytochrome c function without a substantial loss of function of complex IV. Although this is most easily explained by the release of cytochrome c, an examination of their data also showed an inhibition of complex I and II function as we observed. In an earlier study, cell death induced by TNF (a pathway with similarities to that of Fas) was shown to coincide with loss of complex I and II activity . These observations support our conclusions that caspase-dependent loss of mitochondrial function during apoptosis involves a disruption of complexes I and II.
8 p- q" ?) w9 g, v( t: W
0 D8 W* ?9 p- F% B3 T+ {The simplest way in which caspases can disrupt mitochondrial function is via cleavage of molecules important for electron transport. A survey of the components of the electron transport complexes reveals several potential caspase cleave sites based on known specificities of the caspases . Whether these are actual caspase substrates or not and their accessibility to caspases during apoptosis are currently unknown. Other alternative targets may be transport molecules or other systems that impact on the function of the electron transport chain.  d' g; Y0 b( P4 K7 G% h

- q% ^; O# Y) C& b, eDuring apoptosis, proteins of the intermembrane space are released, but those of the matrix are not , suggesting that the inner membrane remains intact (which is also supported by our observation that m is maintained). Without mitochondrial outer membrane permeabilization, caspase-3 had no effect on m or respiration . Therefore, the relevant caspase substrates are presumably accessible on the outside of the inner membrane (i.e., exposed to the inter-membrane space). Alternatively, activation of caspases within the mitochondria, which has been described  , may play a role here. How these would become activated upon exposure of the mitochondria to exogenous caspase-3 is, however, unclear.
# {# a. x- ?3 m  J  t9 q
% K' ^' Y$ \$ r, ^8 cOur observations that production of ROS during apoptosis can be caspase dependent  suggest that they are not required for apoptosis per se as shown using other methods by others . However, the production of ROS during apoptosis is likely to contribute to cell death . Scavenging of ROS can delay or prevent cell death during apoptosis in several systems. Therefore, caspase-induced ROS production may play roles in the dismantling of the cell after caspase activation. Similarly, the loss of electron transport activity and m would impact on all other mitochondrial functions, further contributing to the dismantling of the cell after the activation of caspases.
2 E0 s  J, B9 W$ _, O8 S) v2 g2 O! c0 J% ^% ]8 r
Materials and methods% j& Y: ^0 D3 G" k; H5 d
' n, o$ G  t4 A7 O2 \( h: d5 D* }) g/ Y
Chemicals and recombinant proteins; l0 S( H# c3 H) a8 H2 M

( h: P) m0 q# a9 i! @* _; iFCCP (10 µM), oligomycin (10 µg/ml), cytochrome c (100 µM), and all of the substrates and inhibitors of the electron transport chain (see below) were from . Recombinant Bcl-xL- and caspase-3 were produced in bacteria as previously described . Bcl-xL-C was used at 20 µg/ml.8 U! ]; r3 V( K9 H1 a5 `. G
2 q+ I! ?: g( b- i( ]& H' F
Caspase-free tBid was obtained as described . Briefly, Amino acids 57–62 were replaced by the thrombin cleavage sequence LVPRGS using site-directed mutagenesis (overlap extension method). The resulting fusion protein was activated by thrombin cleavage, producing the same COOH-terminal fragment of Bid that results from caspase-8 cleavage of wild-type full-length Bid. In addition, a 6-histidine tag was attached to the COOH terminus to facilitate purification of the active fragment. The plasmid was then transformed into Escherichia coli BL21 (DE3)  and protein expression was induced by addition of IPTG (0.5 mM). After lysis, the recombinant protein was purified using glutathione–Sepharose-4B beads  After three washes each with lysis buffer containing 0.1% Triton X-100 and PBS, the beads were incubated with 100 U of thrombin in 4 ml PBS for 2 h at 22°C to cleave off the COOH-terminal portion corresponding to tBid (aa 61–195) with a 6 x His tail. The supernatant of the cleavage reaction, containing tBid-His6, was bound to 4 ml Ni-NTA resin. This resin was loaded into a column and washed sequentially with PBS, PBS containing 300 mM additional NaCl, and finally PBS, pH 6.0, containing 300 mM NaCl. The tBid was eluted with 100 mM imidazole in PBS, pH 6.0, containing 300 mM NaCl and dialyzed against PBS containing 10% glycerol for 6 h before storage at –80°C.
( @9 k1 v4 e. [4 N2 L5 W* W* S: s! B4 [& h
Cell culture and induction of apoptosis! L* h$ R7 R6 O- X' H
( x+ D& z2 ]$ W& `; X. a
HeLa cells were cultured in DMEand Jurkat cells in RPMI-1640 supplemented with 2 mM glutamine, 200 µg/ml penicillin, 100 µg/ml streptomycin sulfate, and 10% FBS. Cells were maintained at 37°C in a humidified atmosphere of 95% air, 5% CO2. For passage, adherent cells were incubated in 0.25% trypsin , washed, and subcultured in growth medium. Jurkat cells were subcloned 1:10 when they reached 106 cells/ml.; V6 M( Q; {' |1 C. g6 x8 H6 x" R/ o

- [$ U3 ?3 a" YTo induce death, cells were preincubated or not with 50 µM zVAD-fmk (Kamiya Biomedical Company) and treated with staurosporine, actinomycin D, or with UV (as indicated) and then incubated for 18 h at 37°C. For UV treatment, cells were washed in PBS and irradiated with UV light in PBS at 37°C. The PBS was then aspired, and medium was replaced.
$ e" _" p5 i- t/ e6 d; N* r, u# J" K" i! S; E2 K7 g
Isolation of mitochondria2 Q; u/ {( O1 }5 U' a
, \  d6 u7 F8 g% s
Mitochondria were isolated as described in detail previously . The isolation procedure was performed at 4°C. Briefly, mouse liver was resuspended in 10 ml of mitochondrial isolation buffer (MIB: 220 mM mannitol, 68 mM sucrose, 10 mM Hepes-KOH, pH 7.4, 70 mM KCl, 1 mM EGTA, 1 mM PMSF, and 2 µM aprotinin) and dissociated using a 15 ml dounce with a tight fitting teflon pestle. Mitochondria were isolated by multiple steps of centrifugation in a Sorvall centrifuge with a swinging bucket rotor (HB4). The cellular lysates were centrifuged at 600 g for 10 min, and the supernatants were centrifuged at 3,500 g for 15 min. The mitochondrial pellets were resuspended in 15 ml of fresh MIB, centrifuged at 1,500 g for 5 min, and the supernatant centrifuged at 5,500 g for 10 min. The last two steps were repeated twice. The final pellets were resuspended in 400 µl of ice cold MIB.
; p- F' u$ s4 ^3 b6 K8 d% T( @6 j$ e- F3 I
Analysis of m in permeabilized cells and isolated mitochondria4 o1 k7 Y: Y8 R% S# f) a+ i
/ g8 R6 p+ K# M- ^+ m
For m analysis of isolated mitochondria , 20 µg of mitochondria was resuspended in buffer A (200 mM mannitol, 50 mM sucrose, 10 mM succinate, 10 mM Hepes-KOH, pH 7.4, 5 mM potassium phosphate pH 7.4, 5 mM DTT, and 50 nM TMRE) and incubated in the presence of tBid, caspase-3, and zVAD-fmk or FCCP as indicated, for 45 min at 37°C.) W# J7 S0 a" J6 I1 F6 o4 N3 t
& [3 v7 r6 w: M
For permeabilization, HeLa cells were trypsinized for 5 min at RT and washed with PBS. Jurkat cells were harvested and washed with PBS. The cells were resuspended in ice cold MIB containing 30 µg/ml digitonin until >95% of the cells were permeable to Trypan blue. Then the cells are washed twice in MIB (4°C).9 F7 O  t2 _! H

/ q+ c+ _5 i; @; HFor m measurement, permeabilized cells (106/ml) were incubated in buffer A  or in buffer B  (MIB   2 mM ADP   2 mM DTT   50 nM TMRE) and then incubated in the presence of substrates for the electron transport chain (see concentrations below). In , cells were first permeabilized, then incubated for 30 min at 37°C in the presence or absence of recombinant active caspase-3 (0.5 µg/ml) in buffer B containing cytochrome c (100 µM) or zVAD-fmk (100 µM), or Bcl-xL-C (20 µg/ml), or both and finally incubated in the presence of the substrates for the electron transport chain. In , oligomycin (10 µg/ml) is added when indicated. Cells and isolated mitochondria were then analyzed by flow cytometry on a FACScan  measuring TMRE fluorescence in FL-2.
. Y( o9 F  q4 i, ^
; l3 u$ x, ^. p: A0 ^- F9 L% ^6 rOxygen electrode measurement
) g3 F  A. v8 m) J! t$ v* }/ S/ n. g8 F$ ~2 `8 v( h2 Z
Two independent Clark oxygen electrodes (Instech Laboratories) with two independent thermojacketed chambers were used. This dual system allowed us to analyze two samples in parallel. For isolated mitochondria, the respiration buffer (RB) was 140 mM KCl, 10 mM MgCl2, 10 mM MOPS (pH 7.4), 5 mM KH2PO4, DTT 5 mM, 1 mM EGTA (or 0.2 mM EGTA in case of). For permeabilized cells, the RB was 250 mM sucrose, 2 mM EDTA, 30 mM KH2PO4, 5 mM MgCl2, and 50 mM Tris (pH 7.4). The volume corresponding to 400 µg of protein was injected into the chambers containing 600 µl of air-saturated RB prewarmed at 37°C. To rule out an effect of dilution of cytochrome c, all measurements were performed in the presence of 100 µM cytochrome c. Substrates and inhibitors were added in the following order and final concentration: 2.5 mM malate, 40 µM O-palmitoyl-L-carnitine, 2 mM ADP, 2 µM rotenone, 5 mM succinate, 1 µM antimycin A, 1 mM ascorbate with 0.4 mM TMPD, and 1 mM potassium cyanide (KCN). Oxygen concentration was calibrated with air-saturated buffer, assuming 390 ng-atoms of oxygen/ml of buffer . Rates of oxygen consumption are expressed as ng-atoms of oxygen/min/mg of proteins.
0 o. U" M( j' @' i+ ~+ f# h7 N5 G3 d
; B0 }9 v; J2 CMeasurement of ROS and apoptosis
8 k6 G% _5 G4 r( m0 J$ A+ b6 Z7 v0 Y" D! {+ S8 z4 B0 o+ j
HeLa cells  were treated as indicated, harvested, and washed in PBS. The pellet was resuspended in 30 µl of annexin buffer (Hepes 10 mM, NaCl 150 mM, KCl 5 mM, MgCl2 1 mM, CaCl2 1.8 mM) and then divided into three groups: one third each for m measurement (as described above), ROS measurement using 2 µM of 2-HE in MIB buffer, and cell death measurement using annexin V–FITC . Cells were incubated for 30 min at 37°C in the dark. Analysis was made by flow cytometry; m was measured in FL2, 2-HE in FL2, annexin V in FL1, and propidium iodide (PI; 0.5 µg/ml added at the last minute to the sample) in FL3. The percentage of cell death in  represents the total of annexin V–positive and annexin V/PI double positive cells. In , Jurkat cells (106/ml) were permeabilized as described above and incubated in MIB plus 2 mM ADP, 2 mM DTT, 2 µM 2-HE, and caspase-3 (0.5 µg/ml), tBid (20 µg/ml), and/or Bcl-xL-c (20 µg/ml) as indicated.1 a0 d0 {1 R# j1 D4 N

! H( ?% i8 o2 o/ o; r0 ZComplex III assay. J9 L1 X8 w: p2 j3 Y* s

( \5 q/ N6 L8 y/ b+ HMeasurements were performed as described previously . Briefly, 500 µg of mitochondria were incubated in 100 µl of buffer C (125 mM sucrose, 60 mM KCl, 20 mM Tris-HCl, pH 7.4) in the presence of tBid (25 µg/ml) or recombinant caspase-3 (25 µg/ml) (as indicated) for 60 min at 37°C, or in 1 ml of water (20 min at 4°C), then pelleted by centrifugation (5,500 rpm, 10 min) and resuspended in 100 µl of buffer C. Then the samples were mixed with 300 µl of buffer C containing 3 mM of KCN (to block oxidation by complex IV) and decyl benzoquinol (55 µM final). Finally, ferricytochrome c (80 µM) was added and the rate of cytochrome c reduction at 550 nm was integrated over 30 s. Where indicated, 1 µM of antimycin A was added.  {8 A9 ]. q! G, U9 v

' @- U8 C7 Y+ _+ w6 V3 SWestern blotting
0 D4 T4 A9 W2 V$ T$ h9 i  X* H( ^! O- Z$ F7 ]2 o% e! u
To determine mitochondrial content of hsp60 and cytochrome c, incubation aliquots were centrifuged (6,000 g, 10 min), and the pellet was resuspended in 1x loading buffer. Samples were heated at 95°C and loaded on a 15% SDS–polyacrylamide gel for electrophoresis and then transferred to nitrocellulose. Membrane were blocked 1 h in TBST (25 mM Tris, 140 mM NaCl, 27 mM KCl and 0.02% Tween 20) containing 5% nonfat dried milk. Membranes were then probed with monoclonal anti–cytochrome c (clone 7H8.2C12; PharMingen) or hsp60 (clone LK-1; Stressgen). Recognized proteins were detected using HRP-labeled secondary antibodies  and ECL
2 W' h% N  \. U% X3 F3 _  k
3 _3 U% X; A5 u4 t# w9 J$ fAcknowledgments
3 _: E; N2 W) s( n5 O& ^* z6 O. V$ V( o
We thank Nigel Waterhouse for invaluable discussion, Rick Sayen for invaluable help, and S. Korsmeyer (Harvard Medical School, Boston, MA) for Bid-/- MEF.& ]8 A6 ]- y  s5 D7 m0 y

& h- [! t& F4 ~8 g- lThis work was supported by grants AI40646 and CA69381 from the National Institutes of Health. J.-E. Ricci received fellowships from the Association contre le Cancer, Institut National de la Santé et de la Recherche Médicale, and the Philippe Foundation.6 G8 q% y4 j) _  D7 Q8 F4 N
) |% C* t! J1 |+ d1 O2 _8 Z
Submitted: 15 August 2002: i1 i' E2 I& e; h  P- A6 w

. [5 B. x* L8 X% g/ V9 Q- uRevised: 25 November 2002
" E, M6 c7 R- b- q! m2 g$ C# Q
% O# N* ~. b5 W; yAccepted: 2 December 2002
2 a2 i) T$ U& w% _# j5 Q( m) O& D- A. {: |) D
References" H6 [9 f- z. m! M
$ Y2 k  e/ I$ p, L/ Q
Amarante-Mendes, G.P., D.M. Finucane, S.J. Martin, T.G. Cotter, G.S. Salvesen, and D.R. Green. 1998. Anti-apoptotic oncogenes prevent caspase-dependent and independent commitment for cell death. Cell Death Differ. 5:298–306.
2 s- I  w5 f0 {) z/ v( W( V! F- {- s7 r( o
Bernardi, P., R. Colonna, P. Costantini, O. Eriksson, E. Fontaine, F. Ichas, S. Massari, A. Nicolli, V. Petronilli, and L. Scorrano. 1998. The mitochondrial permeability transition. Biofactors. 8:273–281.
& J: y1 W6 @3 _4 k& S
9 ~# q5 i' E' Z4 c0 c* u& r' zBossy-Wetzel, E., and D.R. Green. 1999. Caspases induce cytochrome c release from mitochondria by activating cytosolic factors. J. Biol. Chem. 274:17484–17490.# Z' q8 N7 U1 e3 s

, x  k0 ?2 Y7 q: S1 s2 TBossy-Wetzel, E., D.D. Newmeyer, and D.R. Green. 1998. Mitochondrial cytochrome c release in apoptosis occurs upstream of DEVD-specific caspase activation and independently of mitochondrial transmembrane depolarization. EMBO J. 17:37–49.
7 D% ]0 p0 ^* ?9 l& n9 d
* t. |9 O% ~7 q) M) ?Coleman, M.L., E.A. Sahai, M. Yeo, M. Bosch, A. Dewar, and M.F. Olson. 2001. Membrane blebbing during apoptosis results from caspase-mediated activation of ROCK I. Nat. Cell Biol. 3:339–345.0 ]4 H/ k: J$ K9 Q$ N( e0 Z
2 I5 M; q8 [% d: k4 u$ ^
Deshmukh, M., K. Kuida, and E.M. Johnson, Jr. 2000. Caspase inhibition extends the commitment to neuronal death beyond cytochrome c release to the point of mitochondrial depolarization. J. Cell Biol. 150:131–143.) E4 O4 D7 @% ]

, u; {7 E3 x! m( gDu, C., M. Fang, Y. Li, L. Li, and X. Wang. 2000. Smac, a mitochondrial protein that promotes cytochrome c-dependent caspase activation by eliminating IAP inhibition. Cell. 102:33–42.
0 n6 h- q8 K  g8 R' i
" Y% P+ G! S  q; `0 yEnari, M., H. Sakahira, H. Yokoyama, K. Okawa, A. Iwamatsu, and S. Nagata. 1998. A caspase-activated DNase that degrades DNA during apoptosis, and its inhibitor ICAD. Nature. 391:43–50.
; ~! W5 p+ ^# y; ~7 Q* P# G) ^6 Y5 v3 ]; E% e( ^3 Z
Farkas, D.L., M.D. Wei, P. Febbroriello, J.H. Carson, and L.M. Loew. 1989. Simultaneous imaging of cell and mitochondrial membrane potentials. Biophys. J. 56:1053–1069.
! V8 L0 \3 D1 L; S5 }
2 c# A# j4 u, y! N2 i2 D+ NFontaine, E., O. Eriksson, F. Ichas, and P. Bernardi. 1998. Regulation of the permeability transition pore in skeletal muscle mitochondria. Modulation by electron flow through the respiratory chain complex I. J. Biol. Chem. 273:12662–12668.& r5 g; F2 b1 e8 N& j% j2 p# z) L

2 W  t% C  q0 N- V, O9 f# R& f+ MGoldstein, J.C., N.J. Waterhouse, P. Juin, G.I. Evan, and D.R. Green. 2000. The coordinate release of cytochrome c during apoptosis is rapid, complete and kinetically invariant. Nat. Cell Biol. 2:156–162.
0 v  Z% a+ Z' R$ }" h1 \* c
1 T. `0 s& G# u- r/ G" v7 KGottlieb, E., M.G. Vander Heiden, and C.B. Thompson. 2000. Bcl-x(L) prevents the initial decrease in mitochondrial membrane potential and subsequent reactive oxygen species production during tumor necrosis factor alpha-induced apoptosis. Mol. Cell. Biol. 20:5680–5689.
( u, u2 |( T4 \1 T/ J1 v- W  E& t5 \: g8 y0 L9 [! H5 k
Heibein, J.A., M. Barry, B. Motyka, and R.C. Bleackley. 1999. Granzyme B-induced loss of mitochondrial inner membrane potential (Delta Psi m) and cytochrome c release are caspase independent. J. Immunol. 163:4683–4693.8 M: H5 d$ O" I! ]3 R: I
9 z% ]7 Q) ]0 J5 d; i$ }4 M* M
Jacobson, M.D., and M.C. Raff. 1995. Programmed cell death and Bcl-2 protection in very low oxygen. Nature. 374:814–816.. s2 e7 A3 C8 `% Q- k  Q/ H
( b+ L# @: L6 S) I  f( }9 P; F
Kluck, R.M., M.D. Esposti, G. Perkins, C. Renken, T. Kuwana, E. Bossy-Wetzel, M. Goldberg, T. Allen, M.J. Barber, D.R. Green, and D.D. Newmeyer. 1999. The pro-apoptotic proteins, Bid and Bax, cause a limited permeabilization of the mitochondrial outer membrane that is enhanced by cytosol. J. Cell Biol. 147:809–822.
" [4 _; i& y2 }$ m# d; e' b* w$ m/ o
Korsmeyer, S.J., M.C. Wei, M. Saito, S. Weiler, K.J. Oh, and P.H. Schlesinger. 2000. Pro-apoptotic cascade activates BID, which oligomerizes BAK or BAX into pores that result in the release of cytochrome c. Cell Death Differ. 7:1166–1173.
3 C% t+ y% I. n* N, h; w2 }: U' P# ~( h1 `7 C* z# G
Krippner, A., A. Matsuno-Yagi, R.A. Gottlieb, and B.M. Babior. 1996. Loss of function of cytochrome c in Jurkat cells undergoing fas-mediated apoptosis. J. Biol. Chem. 271:21629–21636.+ [0 |# y$ g  n

, |7 |$ l' _' N& @Kuwana, T., M.R. Marckey, G. Perkins, M.H. Ellisman, M. Latterich, R. Schneiter, D.R. Green, and D.D. Newmeyer. 2002. Bid, Bax and lipids cooperate to form supramolecular pores in the outer mitochondrial membrane. Cell. 111:331–342.# u3 y3 |; x1 R9 n. a5 E, P: Q" M& q; V

% h; E' W! h; n/ oLi, H., H. Zhu, C.J. Xu, and J. Yuan. 1998. Cleavage of BID by caspase 8 mediates the mitochondrial damage in the Fas pathway of apoptosis. Cell. 94:491–501.
6 ~/ B6 P3 B7 I" ]1 c) U0 r' X/ l# f; U: S0 l
Li, L.Y., X. Luo, and X. Wang. 2001. Endonuclease G is an apoptotic DNase when released from mitochondria. Nature. 412:95–99.
" K7 J  P0 X6 I
) M; ^8 r% `) e' {- D9 OLi, P., D. Nijhawan, I. Budihardjo, S.M. Srinivasula, M. Ahmad, E.S. Alnemri, and X. Wang. 1997. Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 complex initiates an apoptotic protease cascade. Cell. 91:479–489.
0 ~/ ?$ d- H4 x/ A  A$ T- t
5 J  y" c/ l4 O" a3 ~7 I' hMannick, J.B., C. Schonhoff, N. Papeta, P. Ghafourifar, M. Szibor, K. Fang, and B. Gaston. 2001. S-Nitrosylation of mitochondrial caspases. J. Cell Biol. 154:1111–1116.
& ~- G7 g' Q/ \5 E8 `0 D+ o5 n
% ], z( I, M) }# KMartinou, J.C., and D.R. Green. 2001. Breaking the mitochondrial barrier. Nat. Rev. Mol. Cell Biol. 2:63–67.
) K  A: [: J0 \3 V3 Y' x& f5 Q5 c! r) y- c3 z
McCarthy, N.J., M.K. Whyte, C.S. Gilbert, and G.I. Evan. 1997. Inhibition of Ced-3/ICE-related proteases does not prevent cell death induced by oncogenes, DNA damage, or the Bcl-2 homologue Bak. J. Cell Biol. 136:215–227.6 F- n( F  D: s: @3 a
0 N' N+ a8 }* k2 f9 j/ _
Newmeyer, D.D., D.M. Farschon, and J.C. Reed. 1994. Cell-free apoptosis in Xenopus egg extracts: inhibition by Bcl-2 and requirement for an organelle fraction enriched in mitochondria. Cell. 79:353–364.
5 n/ A+ b- O" e& e( v* W  D. m! N" F6 j
Nguyen, M., D.G. Millar, V.W. Yong, S.J. Korsmeyer, and G.C. Shore. 1993. Targeting of Bcl-2 to the mitochondrial outer membrane by a COOH-terminal signal anchor sequence. J. Biol. Chem. 268:25265–25268.$ e+ ]# k. i/ A  o- R8 A# f

# X6 P& W) ?  p& a* Y! sSahara, S., M. Aoto, Y. Eguchi, N. Imamoto, Y. Yoneda, and Y. Tsujimoto. 1999. Acinus is a caspase-3-activated protein required for apoptotic chromatin condensation. Nature. 401:168–173.
* W) t/ r* ]  K4 f
6 p) P3 t& q* X+ {Schulze-Osthoff, K., A.C. Bakker, B. Vanhaesebroeck, R. Beyaert, W.A. Jacob, and W. Fiers. 1992. Cytotoxic activity of tumor necrosis factor is mediated by early damage of mitochondrial functions. Evidence for the involvement of mitochondrial radical generation. J. Biol. Chem. 267:5317–5323.
5 j, z6 b5 [8 W( M$ m# t7 S) \1 `, O4 A: O/ W: R. B( e
Sebbagh, M., C. Renvoize, J. Hamelin, N. Riche, J. Bertoglio, and J. Breard. 2001. Caspase-3-mediated cleavage of ROCK I induces MLC phosphorylation and apoptotic membrane blebbing. Nat. Cell Biol. 3:346–352.
! f8 t( |4 b, j0 Y
  N8 X4 X# d) u  t3 C8 aShimizu, S., Y. Eguchi, H. Kosaka, W. Kamiike, H. Matsuda, and Y. Tsujimoto. 1995. Prevention of hypoxia-induced cell death by Bcl-2 and Bcl-xL. Nature. 374:811–813.1 f" I, c- {; u

( k  S) @3 l3 S% A  F! ]0 Z) g, HStennicke, H.R., M. Renatus, M. Meldal, and G.S. Salvesen. 2000. Internally quenched fluorescent peptide substrates disclose the subsite preferences of human caspases 1, 3, 6, 7 and 8. Biochem. J. 350:563–568., [4 L1 ^% ?* d3 e: M
9 Q; z, Y* ?& R$ v" \
Susin, S.A., H.K. Lorenzo, N. Zamzami, I. Marzo, C. Brenner, N. Larochette, M.C. Prevost, P.M. Alzari, and G. Kroemer. 1999a. Mitochondrial release of caspase-2 and -9 during the apoptotic process. J. Exp. Med. 189:381–394.
* E% J1 S1 M1 s% b) A: E& P" u+ V# G; f
Susin, S.A., H.K. Lorenzo, N. Zamzami, I. Marzo, B.E. Snow, G.M. Brothers, J. Mangion, E. Jacotot, P. Costantini, M. Loeffler, et al. 1999b. Molecular characterization of mitochondrial apoptosis-inducing factor. Nature. 397:441–446.
. Y3 g  T( b2 f5 o4 u) a& p  z5 A' J  M8 r- D) s1 w2 ^
Suzuki, Y., Y. Imai, H. Nakayama, K. Takahashi, K. Takio, and R. Takahashi. 2001. A serine protease, htra2, is released from the mitochondria and interacts with xiap, inducing cell death. Mol. Cell. 8:613–621.
( N' n! ~# _% q, H  J7 c9 J
. }& q4 ~' g: q" i  LTan, S., Y. Sagara, Y. Liu, P. Maher, and D. Schubert. 1998. The regulation of reactive oxygen species production during programmed cell death. J. Cell Biol. 141:1423–1432.
' z! D4 _- `8 g- N$ g* R! ~9 D
1 t( H4 M# G! w  l6 D: x3 ?Verhagen, A.M., P.G. Ekert, M. Pakusch, J. Silke, L.M. Connolly, G.E. Reid, R.L. Moritz, R.J. Simpson, and D.L. Vaux. 2000. Identification of DIABLO, a mammalian protein that promotes apoptosis by binding to and antagonizing IAP proteins. Cell. 102:43–53.) X) s! X& i1 @3 F, u- a) K

" y4 \5 x/ T5 v5 W, lVoisine, C., E.A. Craig, N. Zufall, O. von Ahsen, N. Pfanner, and W. Voos. 1999. The protein import motor of mitochondria: unfolding and trapping of preproteins are distinct and separable functions of matrix Hsp70. Cell. 97:565–574.- S! v  N) M0 |) X8 ^
; {& |, h2 S8 W) i; g" f; E. O
von Ahsen, O., C. Renken, G. Perkins, R.M. Kluck, E. Bossy-Wetzel, and D.D. Newmeyer. 2000. Preservation of mitochondrial structure and function after Bid- or Bax-mediated cytochrome c release. J. Cell Biol. 150:1027–1036.+ S" y. I1 |/ l# B

0 J: S- F+ M* _0 S: C1 J4 ^Waterhouse, N.J., J.C. Goldstein, R.M. Kluck, D.D. Newmeyer, and D.R. Green. 2001a. The (Holey) study of mitochondria in apoptosis. Methods Cell Biol. 66:365–391.* a' B7 Q; b* F& I; I+ b

' s1 {$ l' U% M+ `- G  MWaterhouse, N.J., J.C. Goldstein, O. von Ahsen, M. Schuler, D.D. Newmeyer, and D.R. Green. 2001b. Cytochrome c maintains mitochondrial transmembrane potential and ATP generation after outer mitochondrial membrane permeabilization during the apoptotic process. J. Cell Biol. 153:319–328.* V) n- Z: U6 c" _
. `. ~: }% g! ]5 x
Waterhouse, N.J., J.E. Ricci, and D.R. Green. 2002. And all of a sudden it's over: mitochondrial outer-membrane permeabilization in apoptosis. Biochimie. 84:113–121.* ?( D7 N1 R) t% J

  p* L$ q7 }1 _0 cXiang, J., D.T. Chao, and S.J. Korsmeyer. 1996. BAX-induced cell death may not require interleukin 1 beta-converting enzyme-like proteases. Proc. Natl. Acad. Sci. USA. 93:14559–14563.
& y7 F& N( O  }
; b& _" _' q- l' w; BYin, X.M., K. Wang, A. Gross, Y. Zhao, S. Zinkel, B. Klocke, K.A. Roth, and S.J. Korsmeyer. 1999. Bid-deficient mice are resistant to Fas-induced hepatocellular apoptosis. Nature. 400:886–891.
$ f2 V7 d" `2 |+ @8 I! f
! x. a; K' ?9 {3 vZamzami, N., P. Marchetti, M. Castedo, C. Zanin, J.L. Vayssiere, P.X. Petit, and G. Kroemer. 1995. Reduction in mitochondrial potential constitutes an early irreversible step of programmed lymphocyte death in vivo. J. Exp. Med. 181:1661–1672.* h/ [( h, x7 F- [, v, j
6 p) V7 {; \9 E% O  ~+ h1 I: A
Zamzami, N., S.A. Susin, P. Marchetti, T. Hirsch, I. Gomez-Monterrey, M. Castedo, and G. Kroemer. 1996. Mitochondrial control of nuclear apoptosis. J. Exp. Med. 183:1533–1544.
3 Q+ h( T5 u, Z1 o8 l5 L# B9 E6 D2 x! V4 [: s' w
- p& ~5 b+ I! M- Z+ Y- c' s
$ J, ^% M6 O! r" C8 f2 D

8 t5 y! k; U% d9 x3 x0 @7 ?$ R! z2 |
(Jean-Ehrland Ricci Roberta A. Gottlieb and Douglas R. Green)

Rank: 2

积分
88 
威望
88  
包包
1897  
沙发
发表于 2015-6-15 13:18 |只看该作者
既然来了,就留个脚印  

Rank: 2

积分
75 
威望
75  
包包
2118  
藤椅
发表于 2015-6-29 17:42 |只看该作者
越办越好~~~~~~~~~`  

Rank: 2

积分
136 
威望
136  
包包
1877  
板凳
发表于 2015-7-12 17:07 |只看该作者
干细胞之家微信公众号
干细胞治疗糖尿病  

Rank: 2

积分
122 
威望
122  
包包
1876  
报纸
发表于 2015-7-14 20:43 |只看该作者
这个贴不错!!!!!看了之后就要回复贴子,呵呵  

Rank: 2

积分
66 
威望
66  
包包
1790  
地板
发表于 2015-7-25 04:51 |只看该作者
说的不错  

Rank: 2

积分
69 
威望
69  
包包
1788  
7
发表于 2015-7-25 16:52 |只看该作者
不错不错.,..我喜欢  

Rank: 2

积分
66 
威望
66  
包包
1790  
8
发表于 2015-8-1 10:08 |只看该作者
肿瘤干细胞

Rank: 2

积分
64 
威望
64  
包包
1782  
9
发表于 2015-8-9 17:13 |只看该作者
不错!  

Rank: 2

积分
136 
威望
136  
包包
1877  
10
发表于 2015-8-17 05:25 |只看该作者
一楼的位置好啊..  
‹ 上一主题|下一主题
你需要登录后才可以回帖 登录 | 注册
验证问答 换一个

Archiver|干细胞之家 ( 吉ICP备2021004615号-3 )

GMT+8, 2024-5-4 08:55

Powered by Discuz! X1.5

© 2001-2010 Comsenz Inc.