ArticlePDF Available

Molecular mechanisms of neurotoxicity induced by polymyxins and chemo-prevention.

Authors:

Abstract and Figures

Polymyxins (colistin and polymyxin B) are used increasingly for the treatment of life-threatening infections caused by multidrug resistant (MDR) Gram-negative pathogens, in particular Klebsiella pneumoniae, Acinetobacter baumannii and Pseudomonas aeruginosa. Neurotoxicity is one major unwanted side-effects associated with polymyxin therapy. This review covers our current understanding of polymyxin-induced neurotoxicity, its underlying mechanisms, and the discovery of novel neuro-protective agents to limit this neurotoxicity. In recent years, an increasing body of literature supports the notion that polymyxin-induced nerve damage is largely related to oxidative stress and mitochondrial dysfunction. P53, PI3K/Akt and MAPK pathways are also involved in colistin-induced neuronal cell death. The activation of the redox homeostasis pathways such as Nrf2/HO-1 and autophagy have also been shown to play protective roles against polymyxin-induced neurotoxicity. These pathways have been demonstrated to be up-regulated by neuro-protective agents including curcumin, rapamycin and minocycline. Further research is needed towards the development of novel polymyxin formulations in combination with neuro-protective agents to ameliorate this unwanted adverse effect during polymyxins therapy in patients.
Content may be subject to copyright.
Subscriber access provided by UNIV TEXAS SW MEDICAL CENTER
is published by the American Chemical Society. 1155 Sixteenth Street N.W.,
Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Review
Molecular mechanisms of neurotoxicity
induced by polymyxins and chemo-prevention
Chongshan Dai, Xilong Xiao, Jichang Li, Giuseppe D Ciccotosto, Roberto Cappai, Susheng
Tang, Elena K. Schneider-Futschik, Daniel Hoyer, Tony Velkov, and Jianzhong Shen
ACS Chem. Neurosci., Just Accepted Manuscript • DOI: 10.1021/acschemneuro.8b00300 • Publication Date (Web): 26 Oct 2018
Downloaded from http://pubs.acs.org on October 27, 2018
Just Accepted
“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.
1
Molecular mechanisms of neurotoxicity induced by polymyxins and chemo-
prevention
Chongshan Dai1, Xilong Xiao1, Jichang Li2, Giuseppe D. Ciccotosto3, Roberto
Cappai3, Shusheng Tang1, Elena K. Schneider-Futschik3, Daniel Hoyer3,4,5, Tony
Velkov2#, Jianzhong Shen1,#
1Department of Veterinary Pharmacology and Toxicology, College of Veterinary
Medicine, China Agricultural University, No.2 Yuanmingyuan West Road, Beijing
100193, P. R. China.
2Department of Veterinary Pharmacology and Toxicology, College of Veterinary
Medicine, Northeast Agricultural University, Harbin, P. R. China;
3Department of Pharmacology & Therapeutics, School of Biomedical Sciences, Faculty
of Medicine, Dentistry and Health Sciences, The University of Melbourne, Parkville, VIC,
3010, Australia.
4The Florey Institute of Neuroscience and Mental Health, The University of Melbourne,
30 Royal Parade, Parkville, VIC, 3052, Australia.
5Department of Molecular Medicine, The Scripps Research Institute, 10550 N. Torrey
Pines Road, La Jolla, CA 92037, USA.
*Joint corresponding authors:
Tony Velkov, Telephone: +61 3 83449846. E-mail: Tony.Velkov@unimelb.edu.au OR
Jianzhong Shen, Telephone: +86 10 6273 3857; Fax: +86 10 6273 1032. E-mail:
sjz@cau.edu.cn.
Page 1 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
2
Abstract
Neurotoxicity is one major unwanted side-effects associated with polymyxins (i.e.
colistin and polymyxin B) therapy. Clinically, colistin neurotoxicity is characterized by
neurological symptoms including dizziness, visual disturbances, vertigo, confusion,
hallucinations, seizures, ataxia, facial and peripheral paresthesias. Pathologically,
colistin-induced neurotoxicity is characterized by cell injury and death in neuronal cell.
This review covers our current understanding of polymyxin-induced neurotoxicity, its
underlying mechanisms, and the discovery of novel neuro-protective agents to limit this
neurotoxicity. In recent years, an increasing body of literature supports the notion that
polymyxin-induced nerve damage is largely related to oxidative stress and mitochondrial
dysfunction. P53, PI3K/Akt and MAPK pathways are also involved in colistin-induced
neuronal cell death. The activation of the redox homeostasis pathways such as
Nrf2/HO-1 and autophagy have also been shown to play protective roles against
polymyxin-induced neurotoxicity. These pathways have been demonstrated to be up-
regulated by neuro-protective agents including curcumin, rapamycin and minocycline.
Further research is needed towards the development of novel polymyxin formulations in
combination with neuro-protective agents to ameliorate this unwanted adverse effect
during polymyxins therapy in patients.
Keywords: Polymyxins; Neurotoxicity; Mitochondria dysfunction; Apoptosis; Autophagy;
Neuroprotective agents, Chemo-prevention.
Page 2 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
3
1. Introduction
Antimicrobial resistance is currently a major public health concern; notably The
World Health Organisation (WHO) has identified antibiotic resistance as one of the three
greatest threats to human health.1 During the past two decade, there has been a
dramatic decline in discovery and development of novel classes of antibacterial agents
and, unfortunately, this has been concomitantly accompanied by an increase in the
incidence of infections by multi-drug resistant (MDR) Gram-negative pathogens.2-7 The
situation is especially worrying for MDR Klebsiella pneumoniae, Acinetobacter
baumannii and Pseudomonas aeruginosa, against which no new antibiotics will be
available for many years to come.2 Infections caused by these MDR bacteria, often do
not respond to conventional therapy and result in a longer duration of illness and higher
risk of death. Coughing WHO has placed these three problematic pathogens at the top
of its ‘2017 Antibiotic-resistant Priority Pathogens List’.1 This unfortunate situation has
led to a resurgence in the clinical use of polymyxins (polymyxin B and E syn. colistin),
as a last line treatments against MDR Gram-negative pathogens.8-9
Polymyxins are lipopeptide antibiotics that were first discovered in 1947.10 Unlike
polymyxin B, which is available in the clinic as the sulphate salt, colistin is administered
to patients as an inactive prodrug, colistin methanesulphonate (CMS). 11-12 Polymyxin B
and colistin sulfate are for intravenous, oral and topical use, and colistimethate sodium
(sodium colistin methanesulphonate, colistin sulfomethate sodium) for parenteral use
(Figure 1); both can be delivered by inhalation.13 Since 1959, intravenous polymyxins
have been used for the treatment of Gram-negative bacterial infections in clinic.8 The
clinical use of polymyxins waned in the 1970s, following the early experience in the
Page 3 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
4
1960s which led to numerous cases of nephro- and neurotoxicity.8 A study that
assessed the safety of intramuscular administration of CMS during 317 courses of
therapy to 288 hospitalized patients revealed 83% of these patients experienced
neurotoxicity during the first four days.14 Patients receiving intravenous CMS, have been
reported to present with neurological symptoms including dizziness, visual disturbances,
vertigo, confusion, hallucinations, seizures, ataxia, facial and peripheral paresthesias.15-
17 Symptoms of mild neurotoxicity (such as paresthesias) tend to be more frequent,
especially in elderly patients, but are often ignored because of the lack of objective
assessment protocols.16, 18-20 A recent study showed that at higher polymyxin treatment
doses, clinical outcomes are improved, albeit, therpy was associated with a greater
incidence of adverse effects.21 Notably, high dose colistin exposure has been shown to
induce apoptosis of cerebral cortex neurons, behavioral abnormalities and disrupted
neurotransmitter levels in animal models.18, 22-24 These clinical manifestations indicate
that polymyxin-induced neurotoxicity can be divided into peripheral and central nervous
system (CNS) associated pathological effects. Abnormal
neuro-behavioral changes including sensory and motor dysfunction were detected when
mice were intravenously injected with higher accumulated dosages of colistin (15 mg/kg
per day) for 7 days.24
This review covers the current knowledge-base of the pathophysiology and
molecular mechanisms of polymyxin-induced neurotoxicity, as well as novel neuro-
protective agents.
Page 4 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
5
2. Clinical manifestations of polymyxin associated neurotoxicity
In their excellent clinical review, Falagas and Kasiakou surveyed the literature
from 1950 until May 2005 and concluded that there was a decrease in the reporting of
adverse neurotoxic events associated with polymyxin therapy in patients between 1990-
2005.19 On the other hand, neurotoxicity associated with colistin use was reported more
frequently during 1960-70s.14, 25 Bosso et. al., reported that neurotoxicity commonly
manifests as paresthesias and ataxia in 29% in patients that received intravenous
dosages in excess of 5 mg/kg per day (range, 5.7–8.0 mg/kg per day).20 These
temporal differences may be in part due to advancements in dosing practices with
polymyxins.26 They also noted that the neurotoxic effects of polymyxins are usually mild
and resolve promptly after discontinuation of the treatment. Wabby et. al., reported that
a patient developed rapidly progressive weakness with dyspnea after 5 days of
intravenous CMS (2.5 mg/kg every 12 h).16 Recently, a case study described a patient
with a severe New Delhi metallo--lactamase-1 producing Escherichia coli infection who
exhibited convulsions following intravenous CMS (37,500 IU/kg/8 h, equal to 1.25 mg
colistin base/kg/8 h), followed by acute respiratory muscle weakness and apnoea.17 A
report from John et. al., indicated that the administration of high doses of intravenous
polymyxin B (3-6 mg/kg/day) are coincident with increased neurotoxicity events in
patients.27 These clinical reports are in line with in vitro and animal studies that have
shown colistin-induced neurotoxicity to be dose-dependent. Given that the
cerebrospinal fluid (CSF)-to-plasma ratio of intravenous colistin is 5% (increasing to
>35% with meningeal inflammation), these adverse events are likely due to peripheral
neurotoxicity.18, 23-24, 28
Page 5 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
6
In contrast to intravenous administration, the direct delivery of polymyxins into
the central nervous system (CNS) via the intrathecal (ITH) or intraventricular (IVT)
routes appears to be more effective and safer.29 The ITH/IVT polymyxin dosages
suggested by the Infectious Diseases Society of America guidelines (2004) call for
administration of at least 125,000 IU/per day; however, in the clinical setting, the dose is
often chosen empirically and colistin IT doses of 40,000-500,000 IU/day have been
reported.30-31 Clinical pharmacokinetic data for ITH/IVT colistin suggests that doses >
65,200 IU/day are necessary to achieve sustained concentrations above the MIC of 2
μg/mL (1 mg colistin is equal to 30, 000 IU) (for susceptible Gram-negative pathogens).
32 Moreover, variability between patients is often seen due to fluctuations of intracranial
pressures. In a recent overview, the clinical literature shows that across 62 cases of
reported Gram-negative CNS infections treated with ITH colistin, the majority (83 %)
were due to A. baumannii, followed by 14% due to P. aeruginosa and 3% K.
pneumonia.31, 33-35 The mean duration of treatment was 17 days, and ranged 7-28
days.35 Neurotoxic manifestations including seizures, cauda equina syndrome, chemical
meningitis/ventriculitis were reported in some patients, albeit no nephrotoxicity was
observed.35 The co-administration of rifampicin, amikacin and tigecycline with ITH/IVT
colistin has been shown to be effective for the treatment of MDR Gram-negative
infections and should be considered if supported by synergy tests on CSF culture.
Clearly, there is a need for scientifically-based dosage recommendations for ITH/IVT
polymyxins to treat infections caused by MDR-Gram-negative pathogens in the CNS.31,
33-35
Page 6 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
7
3. Accumulation of polymyxins in the central nervous system
Elucidating the mechanisms of polymyxin uptake into the CNS and neuronal cells
is an important aspect of understanding their unwanted neurotoxic side-effects.36-38 It is
well known that only small molecules with low molecular mass (<450 Da) and high lipid
solubility can efficiently permeate the healthy blood brain barrier (BBB) by a passive
trans-cellular process.39 The tight junctions of the inter-endothelial domains restrict the
passage of large hydrophilic molecules through the para-cellular route, and this is
expected to be the main reason underlying the low BBB penetration of polymyxins
(molecular mass 1,163 Da).40 This may also account for the higher incidence of
peripheral nervous system neurotoxicity as opposed to CNS-mediated neurotoxicity
associated with polymyxin therapy.16, 19 The brain uptake of colistin following a single
intravenous dose (5 mg/kg) or subcutaneous dose (40 mg/kg) to healthy mice was
shown to be negligible.40-43 However, it has been clinically demonstrated that
intravenous administration of CMS, an inactive prodrug of colistin, leads to detectable
levels of colistin in CSF of patients with CNS infections.32 This may be due to the
compromised BBB in patients with Gram-negative CNS infections resulting from
lipopolysaccharide release; the latter has been shown to damage the BBB.40-44
Coincidently, a study from our lab demonstrated that intravenous lipopolysaccharide
administration can increase the BBB transport of colistin in mice, which suggests that
the brain concentrations of colistin can be significantly enhanced during systemic
inflammation, as might be observed in patients with CNS infections.40-43 We also
showed that the continuous administration of intravenous colistin to mice can increase
the colistin concentration in brain tissue without injury of the BBB.22
Page 7 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
8
Our in vitro studies demonstrated that colistin uptake into mouse primary cortical
neuronal cells occurs in a concentration-dependent manner (un-published data). It has
been shown that there are three receptor pathways that included the endocytic receptor
megalin (MR), polypeptide transporter 2 (PEPT2), organic cation transporter 2 (OCTN2)
and p-glycoprotein (P-gp), which appear to co-governen the uptake of colistin into
kidney cells.36, 45-47 Colistin uptake has also been showed to occur in the lung epithelial
cells in a time and concentration-dependent manner.48 Notably, MR, PEPT2 and
OCTN2 are known to be expressed in neuronal cells.49-51 PEPT2 in particular, is present
in the apical membrane of choroid plexus epithelial cells (i.e. the CSF-facing), the site of
the blood-cerebrospinal fluid barrier (BCSFB), where it facilitates substrate efflux from
CSF to blood, thus reducing substrate distribution in CSF.52 A study from Jin et. al.
showed that P-gp inhibition significantly enhanced the uptake of colistin into brain tissue
of mice.42 Therefore, it is tenable to speculate that the disposition of polymyxins in the
CNS is somewhat dependent on the neuronal expression patterns of the
aforementioned transporters.
4. Polymyxin-induced neurotoxicity involves oxidative stress and mitochondrial
dysfunction
The CNS is very susceptible to oxidative damage due to its obligatory elevated
oxygen need and high content of polyunsaturated fatty acids.53 Mitochondria are vital for
maintaining basic cellular functions such as energy metabolism, ATP production and the
induction of cellular apoptosis should these essential functions go astray.54 We have
shown that mitochondria in the cerebrum of mice treated with colistin sulfate
Page 8 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
9
(intravenous injection at 15 mg/kg/day for 7 days) showed profound pathological
changes in their ultrastructure such as disruption of cristae and extensive swelling;
these were accompanied by biochemical perturbations including increased Ca2+-
induced mitochondrial permeability transition, decreased membrane potential and
reduced succinate dehydrogenase activity.24 Mitochondrial dysfunction was also
detected in primary neuronal cells of chicks treated with colistin at 4.15 and 8.3 μg/mL
for 24 h.55 Mitochondrial dysfunction is closely related to excessive reactive oxygen
species (ROS), which leads to damage to lipids, proteins, and DNA, and ultimately
results in cell death.56 In an in vitro mouse N2a neuronal cell model, colistin treatment
(200 μM for 24 h) significantly increased intracellular ROS levels and concomitantly
decreased glutathione (GSH) levels and the activity of the antioxidant enzymes
superoxide dismutase and catalase (CAT); the latter two operate in tandem to dissipate
ROS.28 Antioxidants such as N-acetyl cysteine and ascorbic acid have been shown to
protected neurons against colistin-induced oxidative stress via their scavenging ROS
activity.28 In addition, it is well known that SOD can catalyze the dismutation of the
superoxide anion into oxygen and hydrogen peroxide (H2O2); and subsequently, CAT
catalyzes the decomposition of H2O2 to water and oxygen. Thus, the concomitant
inhibition of SOD and CAT activities may partly contribute to the toxicity of colistin in
neuronal cells.28, 57-58 Nuclear factor erythroid 2-related factor 2 (Nrf2) is a transcription
factor that regulates antioxidant genes by binding to their cognate antioxidant response
elements.59-60 Nrf2 activation promotes the expression of several phase II and
antioxidant enzymes such as heme oxygenase 1 (HO-1), SOD, and CAT.60 Not
surprisingly, we detected the activation of the Nrf2/HO-1 pathway during colistin-
Page 9 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
10
induced cell death in N2a cells and mouse cerebral cortices.24, 28, 61 Together, these
studies indicate a critical role of mitochondrial pathology and oxidative stress in
polymyxin-induced neurotoxicity.
5. Apoptotic pathways involved in polymyxin-induced neuronal cell death
Apoptosis plays a critical role in maintaining brain homeostasis in response to
drug-induced toxicity.62-64 Neuronal cell apoptosis is a notable pathological feature of
colistin-induced neurotoxicity and is dose-dependent.57-58, 61, 65-66 Our previous studies
demonstrated that colistin treatment at 50-200 μM for 24 h, can significantly induce
apoptosis in the neuronal cell lines N2a and PC12, or primary chick neurons28, 57-58, 65, 67-
70. The cytotoxicity of colistin appears to be dependent on the type and source of the
neuronal cell lines, which may relate to the expression levels of the aforementioned
receptors that mediate colistin up-take.28, 48, 65 For example, the HK-2 kidney cell line is
more susceptible to polymyxin-induced apoptosis than the A549 lung cell line, which
may be due to the differential expression of membrane transporters such as megalin,
which is prominently expressed on the apical membrane of human kidney cells, but not
in that of in lung epithelial cells.48 The uptake of polymyxins is thought to be mainly
mediated by PEPT2 in A549 cells48; whereas both PEPT2 and megalin are likely
involved in polymyxin uptake in HK-2 cells.36, 46, 48 The major pathways of colistin-
induced apoptosis that involve both the intrinsic mitochondrial and extrinsic death
receptor pathways, p53, MAPK, PI3K/Akt pathways in neuronal cells are discussed in
detail below (Figure 2).
Page 10 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
11
5.1 Extrinsic and intrinsic apoptotic pathways in polymyxin-induced neurotoxicity
During the process of colistin-induced neuronal cell apoptosis, both the death
receptor (extrinsic) and the mitochondrial (intrinsic) pathways are activated.71 The
mitochondrial pathway is regulated by the pro- and anti-apoptotic Bcl-2 family proteins,
that include Bax and Bcl-2.72 An increase in the Bax/Bcl-2 ratio and cytochrome C
(CytC) release is evident upon colistin treatment of N2a and PC12 neuronal cells, and
these events are concordant with mitochondrial outer membrane permeabilization.28, 48,
65 The release of CytC into the cytoplasm leads to the activation of caspases-3 and -9,
which in turn triggers apoptosis.73 Colistin treatment of PC12 cells also induced Fas and
Fas-L expression, followed by caspase-8 which is activated through the Fas-associated
death domain.70 Notably, caspase pan-inhibitors markedly inhibited colistin-induced
apoptosis in N2a cells, highlighting the key role of caspases across both the intrinsic
and extrinsic pathways.28 In addition, in a mouse model, colistin treatment (18 mg/kg for
7 or 14 days) markedly increased the expression of the 78-kDa glucose-regulated
protein (Grp78) in the cerebral cortices, a marker of the endoplasmic reticulum pathway
of apoptosis, which may also partly contribute to colistin-induced neurotoxicity.61
5.2 The role of the p53 pathway in polymyxin-induced neurotoxicity
There is demonstrable evidence that the p53 pathway participates in colistin-
induced apoptosis in cultured neuronal cells and in animal models.28, 48, 61, 65 Of note,
p53 is a major orchestrator of the cellular response to a broad array of stress types by
regulating apoptosis, cell cycle arrest, senescence, DNA repair and genetic stability.74
Page 11 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
12
Under normal physiological conditions, Mdm2-mediated ubiquitination and proteosomal
degradation control p53 levels. Whereas, under conditions of cellular stress, p53
undergoes reversible post-translational modifications such as acetylization,
ubiquitinoylation, phosphorylation, and sumoylation, which allow for its dissociation from
the suppressor.74 Colistin treatment of PC12 and N2a cells significantly increases p53
protein levels in the nuclear and cytoplasmic compartments as well as its mRNA levels.
In its primary role as a transcription factor, p53 primarily induces the expression of its
target genes, including DRAM, PUMA and Bax. DRAM is an effector of p53-mediated
apoptosis leads to autophagy. As a Bcl-2 homology 3 (BH3)-only family member, PUMA
has been implicated as a p53-upregulated modulator of apoptosis.28, 61, 65 It activates
Bax, which in turn initiates caspase-3 activation via the intrinsic mitochondrial apoptotic
pathway. Inhibition of p53 using an inhibitor PFT-α markedly decreased the expression
of Bax, cleaved-caspase-3 and attenuated colistin-induced cell apoptosis in PC12
neuronal cells.67 Together, it can be surmised that p53-PUMA-Bax pathway plays an
important role in colistin-induced neurotoxicity.
p53 also mediates cross-talk between the mitochondrial and death receptor
pathways by regulating caspase-8 gene expression and subsequent events concerned
with mitochondrial dysfunction.75-76 Recent studies showed that activation of p53 may
play a dual role in colistin-induced apoptosis in PC12 neuronal cells.65 The activation of
p53 inhibits the negative regulator of autophagy mTOR (mammalian target of
rapamycin) via AMP-activated protein kinase (AMPK) pathway activation, which in turn
induces autophagy, a protective response against colistin-induced neuronal cell
apoptosis.65, 67
Page 12 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
13
5.3 The role of the MAPK pathway in polymyxin-induced neurotoxicity
MAPKs are upstream modulators of apoptosis, and at least three subfamilies
have been identified: (i) extracellular signal-regulated kinases (Erks), (ii) c-Jun N-
terminal kinases (JNKs), and p38-MAPKs, which play key roles in a variety of cellular
and molecular process including apoptosis, autophagy, inflammation and immuno-
regulation.73 p38, JNK and Erk are usually considered as a biphasic signaling pathways
that can either act as pro-survival or pro-death, depending on the kinetics and duration
of activation.73 It has been shown that ERKs are important for cell survival, whereas
JNKs and p38-MAPKs are deemed stress responsive and involved in apoptosis.77
Polymyxin B activates JNK and Erk, but not p-p38, in human dermal fibroblasts.78 In a
mouse model, the activation of p-p38 and p-JNK was detected in colistin-induced
nephrotoxicity.79
Colistin treatment of PC12 neuronal cells significantly activated the expression of
p-JNK in a time-dependent manner, which cascaded into apoptosis via the
mitochondrial pathway.65 In colistin-treated N2a neuronal cells, opposing changes of p-
JNK (up-regulation) and p-Erk (down-regulation) were detected, indicating that co-
regulation of JNK and Erk contributed to colistin-induced apoptosis in this particular cell
line.61 Moreover, during colistin-induced oxidative stress, JNK activation can promote
mitochondrial ROS production by up-regulating the Bax/Bcl-2 ratio or p53 expression
(Figure 3).65
5.4 The role of the PI3K/Akt pathway in polymyxin-induced neurotoxicity
The serine/threonine protein kinase Akt/PKB, is one of the most multi-faceted
kinases in the human kinome.80 Akt/PKB plays important roles in response to growth
Page 13 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
14
factors and other extracellular stimuli to regulate several cellular functions including
nutrient metabolism, cell growth, apoptosis and survival. During its activation, Akt is
recruited to the plasma membrane, where it binds to the PI3K products, PI(3,4,5)P3 and
PI(3,4)P2, via its pleckstrin homology (PH) domain and exposes a pair of threonine-308
and serine-473 residues making them amenable to phosphorylation.81 The protein
expression levels of p-Akt are significantly decreased in response to colistin-induced
apoptosis in the RSC96 Schwann and N2a neuronal cell lines.61, 66 The pharmacological
inhibition of Akt using the inhibitor LY294002 markedly enhances colistin-induced
apoptosis, indicating that Akt plays a pro-survival role.61 In line with this finding, Akt
signaling plays an important pro-survival role in neurons exposed to oxidative stress
and against apoptosis induced by noxious xenobiotics, such as trimethyltin chloride,
acrylamide and methyl-4-phenylpyridine.82-84 Moreover, activated Akt confers cell
survival by phosphorylating its cytoplasmic targets, such as glycogen synthase kinase
(GSK)-3β, CREB, Bcl-2, Bax, and caspase-982-84. Not surprisingly, inhibition of Akt
exacerbates colistin-induced downregulation of Bcl-2 levels and mitochondrial
dysfunction in RSC96 and N2a cells.61, 66 CREB is a well-known neuronal cell pro-
survival factor. 85 Upon activation, p-CREB translocates to the nucleus and binds to
CBP; subsequently, the p-CREB-CBP complex binds to the CRE promoter to induce
synaptic remodeling and the expression of acetylcholine esterase, as well as various
neurotransmitter receptors and ion channels.86 Recently, Herkel and colleagues
reported that the activation of the Akt-CREB signaling pathway protects against stress-
induced cell death and attenuates NF-kB-mediated inflammatory response in
RAW264.7 cells.87 Our own data show that the inhibition of p-Akt by LY294002
Page 14 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
15
aggravates the loss of p-CREB in N2a cells, in cascade, to promote colistin-induced
apoptosis; this is in line with a recent report that shows activation of Akt and CREB to
play a pro-survival role against amyloid--induced neurotoxicity in rats.88Moreover, the
Akt pathway is activated by PI3K in response to growth factors such as NGF.89 NGF
was shown to stimulate auto-phosphorylation of TrkA, which activates PI3K/AKT
signaling.90 Notably, a previous study demonstrated that NGF is a key pro-survival
factor in neurons against oxidative stress and neurotoxicity induced by MPTP, MPP+, 6-
OHDA, and H2O2.91-93 Further, NGF supplementation effectively attenuates colistin-
induced neurotoxicity in a mouse model.94 Similarly, in a recent study we showed that
colistin treatment (200 μM for 24 h) significantly inhibited the expression of NGF and up-
stream receptor p75NTR as well as that of p-Akt and p-CREB protein, whereas, it
promoted the protein expression of p-TrkA.61 p75NTR can phosphorylate Akt via the
active phosphatidylinositol 3-kinase to promote cell survival.90
p-CREB interacts with the specific CREB-binding sequence in the promoter
region of NGF and induces NGF expression.95 Thus, any decrease in p-CREB protein
levels would exacerbate colistin-induced apoptosis via inhibition of NGF (Figure 4).61
Indeed, the balance and feedback regulation between NGF-p75NTR and Akt/CREB
pathways may play a critical role in colistin-induced neurotoxicity. Luo et. al., show that
Erk can parallel with Akt to co-positively regulated the nuclear trans-localization of p-
CREB.96 These findings suggest that the inhibition of p-Erk may partly participate in
colistin-induced loss of nuclear CREB. Our group’s data demonstrate that inhibition of
Akt down-regulates the expression of Nrf2 and HO-1.66 61 The PI3K/Akt pathway also
plays a key role in regulating Nrf2-ARE-dependent protection against oxidative stress
Page 15 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
16
caused by colistin. Taken together, these data suggest that the inhibition of the NGF-
p75NTR-Akt-CREB cascade promotes colistin-induced neuronal cell death.
6. The role of autophagy in polymyxin-induced neurotoxicity
Autophagy is an evolutionarily conserved catabolic degradation process by which
cellular proteins and organelles are engulfed by autophagosomes, digested in
lysosomes, and recycled in order to sustain cellular homeostasis in the face of various
stresses including nutrient deprivation, hypoxia, oxidative stress, and DNA damage. 97
Autophagy is known to be involved in the maintenance of neuronal homeostasis,
particularly in response to drug induced-oxidative stress (e.g. cisplatin, etoposide,
olanzapine and staurosporine) and mitochondrial dysfunction.63-64, 71, 98 In general, the
fate of the cell depends on the interplay between pro-apoptotic factors and autophagy,
wherein the latter blocks the induction of apoptosis or acts to delay apoptotic cell death,
and conversely pro-apoptotic caspase activation shuts off autophagy.54, 97 Our recent
study shows that colistin treatment (200 μM) of N2a cells down-regulates the expression
of mTOR and p70s6k, and up-regulates ULK1 expression, indicating that mTOR
inhibition-mediated autophagy is activated in this neuronal cell line following colistin
exposure. Notably, elevated autophagic degradation markers were observed using
transmission electron microscopic examination.61 Rapamycin treatment further inhibited
the expression of mTOR and p70s6k and up-regulated ULK1 expression, which protects
against apoptosis caused by colistin via increased autophagy.61, 68, 99 In contrast, co-
treatment with autophagy inhibitors, such as chloroquine (CQ), exacerbated colistin-
induced apoptosis.28 Furthermore, we have shown that the inhibition of apoptosis using
Page 16 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
17
the caspase inhibitor Z-VAD-FMK markedly attenuated colistin-induced autophagy in
N2a neuronal cells.28 The autophagy-lysosomal pathway plays an important protective
role in various neurodegenerative diseases.54, 100 We have shown that fluorescent
polymyxin probes do not co-localise with the lysosome specific dye LysoTracker Deep
Red, indicating a lack of co-localization of the probes within lysosomes, suggesting the
polymyxin itself may not be degraded via this pathway.101
p53 and JNK are two apoptotic-regulatory factors that are deregulated in
processes of drug, e.g β-amyloid, 3-chloro-1,2-propanediol, olaquindox,-induced nerve,
renal and liver toxicities.76, 99, 102 p53 and JNK can also directly affect autophagy by
regulating the expression of DRAM, a pro-autophagy and pro-apoptotic protein.75-76, 98
We have demonstrated that JNK activation in response to colistin treatment of PC12
neuronal cells results in phosphorylation of Bcl-2 which enhances autophagy by
disrupting the interaction between Bcl-2 and Beclin 1, an autophagy protein (Figure
3).65 Moreover, colistin treatment of PC12 cells can induce the nuclear localization of
p53, which cascades to enhance autophagy through transactivation of target genes
including DRAM (Figure 3).67 Inhibition of JNK by the specific inhibitor SP600125
markedly abolished the overexpression of p53 indiced by colistin treatment, indicating
that the JNK-p53 pathway may partly enhance colistin-induced autophagy.65 In addition,
the activation of p53 may also contribute to colistin-induced inhibition of mTOR, a direct
regulator of autophagy.67, 75 The collective literature indicate that autophagy plays a
protective role against polymyxin-induced apoptotic neuronal cell death; such that
autophagy purposes to recover cellular homeostasis via the degradation of damaged
organelles. Conversely, if polymyxin exposure has damaged the neuronal cell beyond a
Page 17 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
18
recoverable state, then the cell undergoes apoptotic cell death. Clearly, it is the inter-
play of the two pathways that determines the ultimate fate of the neuronal cell.
7. Chemo-prevention of polymyxin-induced neurotoxicity
Available population pharmacokinetic and pharmacodynamic data indicates that
the current recommended dosage regimens of polymyxins achieve suboptimal plasma
concentrations and that higher doses are needed to achieve effective bacterial killing
and to prevent the emergence of resistance.103 However, simply escalating the
polymyxin dose is not an option due to the limited window between antibacterial efficacy
and neurotoxicity. Therefore, the development of effective neuro-protective agents that
can be co-administered during polymyxin therapy is essential for effective and safe
antimicrobial chemotherapy with polymixins. Recent literature reports have provided
evidence that the co-administration of various agents such as curcumin, rapamycin and
minocycline exert a neuro-protective effect against colistin-induced neurotoxicity.57-58, 61
7.1 Curcumin
Curcumin a natural product polyphenol (diferuloylmethane), is the yellow pigment
component in the spice turmeric, extracted from the rhizome of Curcuma longa. Apart
from its tantalizing culinary applications in Indian cooking, curcumin has been reported
to possess a range of beneficial pharmacological properties such as anti-tumor, anti-
inflammatory and anti-oxidant activities.104-106 Accordingly, curcumin has been assessed
for it potential as a therapeutic agent against cancer, cardiovascular, neurological,
Page 18 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
19
autoimmune and metabolic diseases.107-109 Clearly, curcumin’s anti-inflammatory and
anti-oxidative properties hold a great deal clinical potential in a number of chronic and
neuro-protective diseases.107 To this end, our in vitro data showed that curcumin can
effectively attenuate colistin-induced neurotoxicity in N2a cells by inhibiting oxidative
stress, increasing the levels of the intracellular anti-ROS enzymes SOD, CAT and GSH
levels, and activating the Nrf2/HO-1 anti-oxidative stress pathway.58 Curcumin treatment
also down-regulates NF-kB and NF-kB-regulated genes involved in inflammation and
apoptosis in N2a cells.58 Although curcumin exhibits poor systemic availability when
administered via the oral route, various formulations for intravenous use, including
nanoparticles and liposomal encapsulation have shown promise as anti-neoplastic
agents.108-110 Available preclinical and phase I/II data suggest that curcumin is well
tolerated, and has a good safety profile.108 Encouragingly, a recent animal study
showed that rats that were orally administered curcumin (200 mg/kg/day) following
colistin treatment (IP administered 300,000 IU colistin/kg/day for 6 days), showed
markedly ameliorated histological damage and decreased markers of oxidative damage
and inflammation in their brain and kidney tissues compared to the colistin-only
treatment group.111 These data highlight the potential of curcumin as a
neuroprotective/nephroprotective agent that can be safely co-administered orally to
patients receiving polymyxin therapy.
7.2 Minocycline
Minocycline is a broad-spectrum tetracycline antibiotic reported to display
antioxidant and neuroprotective activities.112-116 Since minocycline and colistin produce
Page 19 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
20
a pharmacodynamically synergistic antibacterial effect117-118, their co-administration has
the potential to reduce colistin dosage and thus toxicity, whilst concomitantly exerting
improved bacterial killing, i.e ‘kill two birds with one stone’. Our data demonstrates that
minocycline could not only inhibit colistin-induced ROS generation, but also enhance
the antioxidant capacity of N2a cells by up-regulating the activities of SOD and CAT,
and thus attenuating colistin-induced mitochondrial dysfunction and apoptosis.57 The
potent antioxidant activity of minocycline is believed to be related to its ability to chelate
mitochondrial iron which catalyzes toxic hydroxyl radical formation.100 It can be
surmised from the above that intravenous minocycline-polymyxin combination therapy
overcomes many of the limitations of polymyxin monotherapy and holds great clinical
potential for treatment of CNS infections due to MDR Gram-negatives; notably, (i) The
favorable pharmacokinetic profile of intravenous minocycline119, (ii) antibacterial synergy
with polymyxins117-118, 120-121, 122 neuroprotective/nephroprotective properties122 and (iv)
stability against many tetracycline bacterial resistance enzymes.122
7.3 Salidroside
A recent study from our group shows that salidroside, a medicinal ingredient of
Rhodiola rosea, exerts a potent protective effect against colistin-induced cytotoxicity
and apoptosis in RSC96 Schwann cells.66 Previous studies have shown that salidroside
has neuroprotective effects against glutamate-induced apoptosis in rat primary
hippocampal neurons, as well as alleviating amyloid--induced cell death of human SH-
SY5Y neuroblastoma cells.123-124 Our study demonstrated that salidroside attenuates
Page 20 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
21
colistin-induced oxidative stress and modulates the PI3K/Akt/mitochondrial signaling
pathway.66
7.4 Rapamycin
mTORC1 is the mechanistic target for rapamycin and is part of a regulatory-
associated complex responsible for regulating cell growth in response to nutrients, such
as amino acids, oxygen, and cellular energy levels.125 mTORC1 and activated p70S6K
(a downstream target of mTOR) have been shown to be markedly elevated in the
cerebral cortex and hippocampus in animal models of Alzheimer’s disease and other
neurological diseases such as tuberous sclerosis.126 Rapamycin is a potent and specific
inhibitor of mammalian/mechanistic target of rapamycin (mTOR).127 A growing body of
evidence suggests that rapamycin possesses a neuroprotective role in many
neurodegenerative diseases.81, 88 Rapamycin pre-treatment inhibits the expression of p-
mTOR and p-p70s6k and up-regulates ULK1 expression, which induces autophagy to
help recover N2a neurons from colistin-induced cell death via inhibition of apoptosis.61
Our recent study shows that rapamycin administration at 2.5 mg/kg/day for 14 days can
effectively attenuate colistin-induced neurotoxicity in a mouse model.61 Rapamycin was
also shown to perturb the GSH levels and increased the transcription of anti-oxidant
genes mediated by cap-n-collar (the Drosophila orthologue of Nrf2), inhibiting colistin-
induced caspase-9 and -3 mediated apoptosis in N2a cells and mouse cerebral
cortex.61 In addition, we observed that colistin-induced ROS accumulation can activate
JNK and inhibit Erk1/2, also contributing to neuronal apoptosis.61 Xu et. al., showed that
rapamycin inhibited cadmium-induced phosphorylation of Erk1/2 and JNK and cleavage
Page 21 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
22
of caspase-3 in PC12 cells, SH-SY5Y cells and mouse primary neuronal cells.128-129
Indeed, rapamycin pretreatment can inhibit the expression of p-JNK, which may partly
contribute to inhibit colistin-induced oxidative stress and mitochondrial dysfunction.
Recently, Herkel and colleagues reported that the activation of the Akt-CREB signaling
pathway protects against stress-induced cell death and attenuates the NF-kB-mediated
inflammatory response in Raw264.7 cells.87 Our own data showed that rapamycin pre-
treatment recovered p-Akt and p-CREB to normal levels, in line with a recent report that
shows activation of Akt and CREB to partly contribute to the neuroprotective activity of
rapamycin against colistin-induced neurotoxicity in N2a cells.61, 88 Consistent with our
own data, previous studies have shown that rapamycin can increase Akt
phosphorylation through the inhibition of a negative feedback loop involving
mTORC1.130 Lehigh et. al., reported that target-derived NGF could mediate circuit
formation and synapse maintenance through TrkA endosome signaling within dendrites
to promote aggregation of post-synaptic protein complexes.131 Rapamycin pre-treatment
markedly up-regulated the expression of NGF and its downstream receptor protein p-
TrkA in N2a cells, which may partly contribute to the activation of Akt/CREB pathway
induced by rapamycin, to attenuate neuronal cell death.61 These finding highlights
rapamycin as a promising agent for prevention of colistin-induced oxidative stress and
neurotoxicity, similarly to minocycline, curcumin or salidroside.
8. Conclusions
There is increasing evidence supporting the potential development of novel combination
formulations of polymyxins with neuro-protective agents to reduce polymyxin-induced
neurotoxicity. It is notable that oxidative stress is one of the toxicological hallmarks of
Page 22 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
23
polymyxin-induced neurotoxicity. Mitochondria, as the major generator of ROS, appear
to play a prominent role in colistin-induced neuronal cell death by mediating apoptotic
pathways. Therefore, alleviating mitochondrial dysfunction is a potential target for the
prevention of polymyxin-induced neurotoxicity. Cellular autophagy also plays an
important role in the process of polymyxin-induced neurotoxicity. Data from animal
models and in vitro cell culture demonstrate that autophagy is markedly induced during
polymyxin treatment of neurons in a rescue attempt against apoptotic cell death.
Colistin-induced apoptosis and autophagy is also regulated by complex signal pathways,
such as PI3K/Akt, MAPK, NF-kB, p53, mTOR and Nrf2/HO-1, all of which represent
potential pharmacological targets. To this end drugs and supplements (i.e minocycline,
rapamycin, salidroside and curcumin) already approved for human use represent the
most tangible resource, as these ‘off the shelf’ agents are safe and effective for human
use. It appears that these compounds operate via diverse mechanisms to either inhibit
ROS production or stimulate cellular salvage and repair pathways to recover neuronal
homeostasis. On key criteria all of these compounds need to satisfy in order to be
clinically effective is the ability to accumulate in sufficient quantities in the CNS, which
where the advantages of ITH/IVT administration come into play. Clinical trials are clearly
warranted with these agents in patients receiving polymyxin therapy.
Page 23 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
24
Funding
This study was supported by Key Projects in Chinese National Science and Technology
Pillar Program during the 12th Five-year Plan Period (2015BAD11B03, to C.D., S.T.). T.
V. is supported by the Australian National Health and Medical Research Council
(NHMRC).
Transparency declarations
None to declare
Author Contributions
C.D, X.X, J.L, G.D.C, R.C, S.T, and E.K.S-F, wrote, edited and proof read the
manuscript. D.H, T.V, and J.S conceived the and proof read the manuscript.
Page 24 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
25
References
1. World Heath Organization, Global priority list of antibiotic-resistant bacteria to
guide research, discovery, and development of new antibiotics 2017.
2. Braine, T., Race against time to develop new antibiotics. Bull World Health Organ
2011, 89 (2), 88-9.
3. Zowawi, H. M.; Harris, P. N.; Roberts, M. J.; Tambyah, P. A.; Schembri, M. A.;
Pezzani, M. D.; Williamson, D. A.; Paterson, D. L., The emerging threat of multidrug-
resistant Gram-negative bacteria in urology. Nat Rev Urol 2015, 12 (10), 570-84.
4. Demiraslan, H.; Cevahir, F.; Berk, E.; Metan, G.; Cetin, M.; Alp, E., Is
surveillance for colonization of carbapenem-resistant gram-negative bacteria important
in adult bone marrow transplantation units? Am J Infect Control 2017, 45 (7), 735-739.
5. Exner, M.; Bhattacharya, S.; Christiansen, B.; Gebel, J.; Goroncy-Bermes, P.;
Hartemann, P.; Heeg, P.; Ilschner, C.; Kramer, A.; Larson, E.; Merkens, W.; Mielke, M.;
Oltmanns, P.; Ross, B.; Rotter, M.; Schmithausen, R. M.; Sonntag, H. G.; Trautmann,
M., Antibiotic resistance: What is so special about multidrug-resistant Gram-negative
bacteria? GMS Hyg Infect Control 2017, 12, Doc05.
6. Oliveira, V. D.; Rubio, F. G.; Almeida, M. T.; Nogueira, M. C.; Pignatari, A. C.,
Trends of 9,416 multidrug-resistant Gram-negative bacteria. Rev Assoc Med Bras
(1992) 2015, 61 (3), 244-9.
7. Perez, F.; Adachi, J.; Bonomo, R. A., Antibiotic-resistant gram-negative bacterial
infections in patients with cancer. Clin Infect Dis 2014, 59 Suppl 5, S335-9.
8. Li, J.; Nation, R. L.; Turnidge, J. D.; Milne, R. W.; Coulthard, K.; Rayner, C. R.;
Paterson, D. L., Colistin: the re-emerging antibiotic for multidrug-resistant Gram-
negative bacterial infections. Lancet Infect Dis 2006, 6 (9), 589-601.
9. Nation, R. L.; Li, J.; Cars, O.; Couet, W.; Dudley, M. N.; Kaye, K. S.; Mouton, J.
W.; Paterson, D. L.; Tam, V. H.; Theuretzbacher, U.; Tsuji, B. T.; Turnidge, J. D.,
Framework for optimisation of the clinical use of colistin and polymyxin B: the Prato
polymyxin consensus. The Lancet Infectious Diseases 2015, 15 (2), 225-234.
10. Velkov, T.; Thompson, P. E.; Nation, R. L.; Li, J., Structure--activity relationships
of polymyxin antibiotics. J Med Chem 2010, 53 (5), 1898-916.
11. Nation, R. L.; Velkov, T.; Li, J., Colistin and polymyxin B: peas in a pod, or chalk
and cheese? Clin Infect Dis 2014, 59 (1), 88-94.
12. Nation, R. L.; Garonzik, S. M.; Thamlikitkul, V.; Giamarellos-Bourboulis, E. J.;
Forrest, A.; Paterson, D. L.; Li, J.; Silveira, F. P., Dosing guidance for intravenous
colistin in critically-ill patients. Clin Infect Dis 2017, 64 (5), 565-571.
13. Velkov, T.; Abdul Rahim, N.; Zhou, Q. T.; Chan, H. K.; Li, J., Inhaled anti-
infective chemotherapy for respiratory tract infections: successes, challenges and the
road ahead. Adv Drug Deliv Rev 2015, 85, 65-82.
14. Koch-Weser, J.; Sidel, V. W.; Federman, E. B.; Kanarek, P.; Finer, D. C.; Eaton,
A. E., Adverse effects of sodium colistimethate. Manifestations and specific reaction
rates during 317 courses of therapy. Ann Intern Med 1970, 72 (6), 857-68.
15. Weinstein, L.; Doan, T. L.; Smith, M. A., Neurotoxicity in patients treated with
intravenous polymyxin B: Two case reports. Am J Health Syst Pharm 2009, 66 (4), 345-
7.
Page 25 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
26
16. Wahby, K.; Chopra, T.; Chandrasekar, P., Intravenous and inhalational colistin-
induced respiratory failure. Clin Infect Dis 2010, 50 (6), e38-40.
17. Honore, P. M.; Jacobs, R.; Lochy, S.; De Waele, E.; Van Gorp, V.; De Regt, J.;
Martens, G.; Joannes-Boyau, O.; Boer, W.; Spapen, H. D., Acute respiratory muscle
weakness and apnea in a critically ill patient induced by colistin neurotoxicity: key
potential role of hemoadsorption elimination during continuous venovenous
hemofiltration. Int J Nephrol Renovasc Dis 2013, 6, 107-11.
18. Dai, C.; Tang, S.; Li, J.; Wang, J.; Xiao, X., Effects of colistin on the sensory
nerve conduction velocity and F-wave in mice. Basic & clinical pharmacology &
toxicology 2014, 115 (6), 577-80.
19. Falagas, M. E.; Kasiakou, S. K., Toxicity of polymyxins: a systematic review of
the evidence from old and recent studies. Crit Care 2006, 10 (1), R27.
20. Bosso, J. A.; Liptak, C. A.; Seilheimer, D. K.; Harrison, G. M., Toxicity of colistin
in cystic fibrosis patients. DICP 1991, 25 (11), 1168-70.
21. Nelson, B. C.; Eiras, D. P.; Gomez-Simmonds, A.; Loo, A. S.; Satlin, M. J.;
Jenkins, S. G.; Whittier, S.; Calfee, D. P.; Furuya, E. Y.; Kubin, C. J., Clinical outcomes
associated with polymyxin B dose in patients with bloodstream infections due to
carbapenem-resistant Gram-negative rods. Antimicrob Agents Chemother 2015, 59
(11), 7000-6.
22. Wang, J.; Yi, M.; Chen, X.; Muhammad, I.; Liu, F.; Li, R.; Li, J.; Li, J., Effects of
colistin on amino acid neurotransmitters and blood-brain barrier in the mouse brain.
Neurotoxicol Teratol 2016, 55, 32-7.
23. Dai, C. S.; Li, J. C.; Lin, W.; Li, G. X.; Sun, M. C.; Wang, F. X.; Li, J.,
Electrophysiology and ultrastructural changes in mouse sciatic nerve associated with
colistin sulfate exposure. Toxicol Mech Method 2012, 22 (8), 592-596.
24. Dai, C.; Li, J.; Li, J., New insight in colistin induced neurotoxicity with the
mitochondrial dysfunction in mice central nervous tissues. Exp Toxicol Pathol 2013, 65
(6), 941-8.
25. Fekety, F. R., Jr.; Norman, P. S.; Cluff, L. E., The treatment of gram-negative
bacillary infections with colistin. The toxicity and efficacy of large doses in forty-eight
patients. Ann Intern Med 1962, 57, 214-29.
26. Nation, R. L.; Garonzik, S. M.; Li, J.; Thamlikitkul, V.; Giamarellos-Bourboulis, E.
J.; Paterson, D. L.; Turnidge, J. D.; Forrest, A.; Silveira, F. P., Updated US and
European Dose Recommendations for Intravenous Colistin: How Do They Perform?
Clin Infect Dis 2016, 62 (5), 552-558.
27. John, J. F.; Falci, D. R.; Rigatto, M. H.; Oliveira, R. D.; Kremer, T. G.; Zavascki,
A. P., Severe Infusion-Related Adverse Events and Renal Failure in Patients Receiving
High-Dose Intravenous Polymyxin B. Antimicrob Agents Chemother 2018, 62 (1).
28. Dai, C.; Tang, S.; Velkov, T.; Xiao, X., Colistin-Induced Apoptosis of
Neuroblastoma-2a Cells Involves the Generation of Reactive Oxygen Species,
Mitochondrial Dysfunction, and Autophagy. Mol Neurobiol 2016, 53 (7), 4685-700.
29. Laxmi, S.; Tunkel, A. R., Healthcare-associated bacterial meningitis. Curr Infect
Dis Rep 2011, 13 (4), 367-73.
30. Tucker, M. J., Carcinogenic Action of Quinoxaline 1,4-Dioxide in Rats. J Natl
Cancer I 1975, 55 (1), 137-146.
Page 26 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
27
31. Bargiacchi, O.; De Rosa, F. G., Intrathecal or intraventricular colistin: a review.
Infez Med 2016, 24 (1), 3-11.
32. Imberti, R.; Cusato, M.; Accetta, G.; Marino, V.; Procaccio, F.; Del Gaudio, A.;
Iotti, G. A.; Regazzi, M., Pharmacokinetics of colistin in cerebrospinal fluid after
intraventricular administration of colistin methanesulfonate. Antimicrob Agents
Chemother 2012, 56 (8), 4416-21.
33. Bargiacchi, O.; Rossati, A.; Car, P.; Brustia, D.; Brondolo, R.; Rosa, F.; Garavelli,
P. L.; De Rosa, F. G., Intrathecal/intraventricular colistin in external ventricular device-
related infections by multi-drug resistant Gram negative bacteria: case reports and
review. Infection 2014, 42 (5), 801-9.
34. Karaiskos, I.; Galani, L.; Baziaka, F.; Giamarellou, H., Intraventricular and
intrathecal colistin as the last therapeutic resort for the treatment of multidrug-resistant
and extensively drug-resistant Acinetobacter baumannii ventriculitis and meningitis: a
literature review. Int J Antimicrob Agents 2013, 41 (6), 499-508.
35. Ng, J.; Gosbell, I. B.; Kelly, J. A.; Boyle, M. J.; Ferguson, J. K., Cure of
multiresistant Acinetobacter baumannii central nervous system infections with
intraventricular or intrathecal colistin: case series and literature review. J Antimicrob
Chemother 2006, 58 (5), 1078-81.
36. Lu, X.; Chan, T.; Xu, C.; Zhu, L.; Zhou, Q. T.; Roberts, K. D.; Chan, H. K.; Li, J.;
Zhou, F., Human oligopeptide transporter 2 (PEPT2) mediates cellular uptake of
polymyxins. J Antimicrob Chemother 2016, 71 (2), 403-12.
37. Yun, B.; Azad, M. A. K.; Nowell, C. J.; Nation, R. L.; Thompson, P. E.; Roberts,
K. D.; Velkov, T.; Li, J., Cellular Uptake and Localization of Polymyxins in Renal Tubular
Cells Using Rationally Designed Fluorescent Probes. Antimicrob Agents Ch 2015, 59
(12), 7489-7496.
38. Nation, R. L.; Li, J.; Cars, O.; Couet, W.; Dudley, M. N.; Kaye, K. S.; Mouton, J.
W.; Paterson, D. L.; Tam, V. H.; Theuretzbacher, U.; Tsuji, B. T.; Turnidge, J. D.,
Framework for optimisation of the clinical use of colistin and polymyxin B: the Prato
polymyxin consensus. Lancet Infect Dis 2015, 15 (2), 225-34.
39. Hitchcock, S. A., Blood-brain barrier permeability considerations for CNS-
targeted compound library design. Curr Opin Chem Biol 2008, 12 (3), 318-23.
40. Jin, L.; Nation, R. L.; Li, J.; Nicolazzo, J. A., Species-Dependent Blood-Brain
Barrier Disruption of Lipopolysaccharide: Amelioration by Colistin In Vitro and In Vivo.
Antimicrob Agents Chemother 2013, 57 (9), 4336-4342.
41. Jin, L.; Li, J.; Nation, R. L.; Nicolazzo, J. A., Effect of systemic infection induced
by Pseudomonas aeruginosa on the brain uptake of colistin in mice. Antimicrob Agents
Chemother 2012, 56 (10), 5240-6.
42. Jin, L.; Li, J.; Nation, R. L.; Nicolazzo, J. A., Impact of p-glycoprotein inhibition
and lipopolysaccharide administration on blood-brain barrier transport of colistin in mice.
Antimicrob Agents Chemother 2011, 55 (2), 502-7.
43. Jin, L.; Li, J.; Nation, R. L.; Nicolazzo, J. A., Brain penetration of colistin in mice
assessed by a novel high-performance liquid chromatographic technique. Antimicrob
Agents Chemother 2009, 53 (10), 4247-51.
44. Banks, W. A.; Gray, A. M.; Erickson, M. A.; Salameh, T. S.; Damodarasamy, M.;
Sheibani, N.; Meabon, J. S.; Wing, E. E.; Morofuji, Y.; Cook, D. G.; Reed, M. J.,
Lipopolysaccharide-induced blood-brain barrier disruption: roles of cyclooxygenase,
Page 27 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
28
oxidative stress, neuroinflammation, and elements of the neurovascular unit. J
Neuroinflammation 2015, 12, 223.
45. Visentin, M.; Gai, Z.; Torozi, A.; Hiller, C.; Kullak-Ublick, G. A., Colistin is
substrate of the carnitine/organic cation transporter 2 (OCTN2, SLC22A5). Drug Metab
Dispos 2017, 45 (12), 1240-1244.
46. Suzuki, T.; Yamaguchi, H.; Ogura, J.; Kobayashi, M.; Yamada, T.; Iseki, K.,
Megalin contributes to kidney accumulation and nephrotoxicity of colistin. Antimicrob
Agents Chemother 2013, 57 (12), 6319-24.
47. Lee, S. H.; Kim, J. S.; Ravichandran, K.; Gil, H. W.; Song, H. Y.; Hong, S. Y., P-
Glycoprotein Induction Ameliorates Colistin Induced Nephrotoxicity in Cultured Human
Proximal Tubular Cells. PLoS One 2015, 10 (8), e0136075.
48. Ahmed, M. U.; Velkov, T.; Lin, Y. W.; Yun, B.; Nowell, C. J.; Zhou, F.; Zhou, Q.
T.; Chan, K.; Azad, M. A. K.; Li, J., Potential Toxicity of Polymyxins in Human Lung
Epithelial Cells. Antimicrob Agents Chemother 2017, 61 (6).
49. Daniel, H.; Kottra, G., The proton oligopeptide cotransporter family SLC15 in
physiology and pharmacology. Pflugers Arch 2004, 447 (5), 610-8.
50. Spuch, C.; Ortolano, S.; Navarro, C., LRP-1 and LRP-2 receptors function in the
membrane neuron. Trafficking mechanisms and proteolytic processing in Alzheimer's
disease. Front Physiol 2012, 3, 269.
51. Courousse, T.; Bacq, A.; Belzung, C.; Guiard, B.; Balasse, L.; Louis, F.; Le
Guisquet, A. M.; Gardier, A. M.; Schinkel, A. H.; Giros, B.; Gautron, S., Brain organic
cation transporter 2 controls response and vulnerability to stress and GSK3beta
signaling. Mol Psychiatry 2015, 20 (7), 889-900.
52. Smith, D. E.; Johanson, C. E.; Keep, R. F., Peptide and peptide analog transport
systems at the blood-CSF barrier. Adv Drug Deliv Rev 2004, 56 (12), 1765-91.
53. Ikonomidou, C.; Kaindl, A. M., Neuronal death and oxidative stress in the
developing brain. Antioxid Redox Signal 2011, 14 (8), 1535-50.
54. Lee, J.; Giordano, S.; Zhang, J., Autophagy, mitochondria and oxidative stress:
cross-talk and redox signalling. The Biochemical journal 2012, 441 (2), 523-40.
55. Dai, C.; Zhang, D.; Li, J.; Li, J., Effect of colistin exposure on calcium
homeostasis and mitochondria functions in chick cortex neurons. Toxicol Mech Methods
2013, 23 (4), 281-8.
56. Stowe, D. F.; Camara, A. K., Mitochondrial reactive oxygen species production in
excitable cells: modulators of mitochondrial and cell function. Antioxid Redox Signal
2009, 11 (6), 1373-414.
57. Dai, C.; Ciccotosto, G. D.; Cappai, R.; Wang, Y.; Tang, S.; Xiao, X.; Velkov, T.,
Minocycline attenuates colistin-induced neurotoxicity via suppression of apoptosis,
mitochondrial dysfunction and oxidative stress. J Antimicrob Chemother 2017, 72 (6),
1635-1645.
58. Dai, C.; Ciccotosto, G. D.; Cappai, R.; Tang, S.; Li, D.; Xie, S.; Xiao, X.; Velkov,
T., Curcumin Attenuates Colistin-Induced Neurotoxicity in N2a Cells via Anti-
inflammatory Activity, Suppression of Oxidative Stress, and Apoptosis. Mol Neurobiol
2018, 55 (1), 421-434.
59. Barcelos, R. P.; Bresciani, G.; Rodriguez-Miguelez, P.; Cuevas, M. J.; Soares, F.
A.; Barbosa, N. V.; Gonzalez-Gallego, J., Diclofenac pretreatment effects on the toll-like
Page 28 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
29
receptor 4/nuclear factor kappa B-mediated inflammatory response to eccentric
exercise in rat liver. Life Sci 2016, 148, 247-53.
60. Dai, C.; Tang, S.; Deng, S.; Zhang, S.; Zhou, Y.; Velkov, T.; Li, J.; Xiao, X.,
Lycopene attenuates colistin-induced nephrotoxicity in mice via activation of the
Nrf2/HO-1 pathway. Antimicrob Agents Chemother 2015, 59 (1), 579-85.
61. Dai, C.; Ciccotosto, G. D.; Cappai, R.; Wang, Y.; Tang, S.; Hoyer, D.; Schneider,
E. K.; Velkov, T.; Xiao, X., Rapamycin Confers Neuroprotection against Colistin-Induced
Oxidative Stress, Mitochondria Dysfunction, and Apoptosis through the Activation of
Autophagy and mTOR/Akt/CREB Signaling Pathways. ACS Chem Neurosci 2018, 9 (4),
824-837.
62. Fu, X. Y.; Yang, M. F.; Cao, M. Z.; Li, D. W.; Yang, X. Y.; Sun, J. Y.; Zhang, Z.
Y.; Mao, L. L.; Zhang, S.; Wang, F. Z.; Zhang, F.; Fan, C. D.; Sun, B. L., Strategy to
Suppress Oxidative Damage-Induced Neurotoxicity in PC12 Cells by Curcumin: the
Role of ROS-Mediated DNA Damage and the MAPK and AKT Pathways. Mol Neurobiol
2014.
63. Xie, B. S.; Zhao, H. C.; Yao, S. K.; Zhuo, D. X.; Jin, B.; Lv, D. C.; Wu, C. L.; Ma,
D. L.; Gao, C.; Shu, X. M.; Ai, Z. L., Autophagy inhibition enhances etoposide-induced
cell death in human hepatoma G2 cells. International journal of molecular medicine
2011, 27 (4), 599-606.
64. Ha, J. Y.; Kim, J. S.; Kim, S. E.; Son, J. H., Simultaneous activation of mitophagy
and autophagy by staurosporine protects against dopaminergic neuronal cell death.
Neuroscience letters 2014, 561, 101-6.
65. Lu, Z.; Chen, C.; Wu, Z.; Miao, Y.; Muhammad, I.; Ding, L.; Tian, E.; Hu, W.; Ni,
H.; Li, R.; Wang, B.; Li, J., A Dual Role of P53 in Regulating Colistin-Induced Autophagy
in PC-12 Cells. Front Pharmacol 2017, 8, 768.
66. Lu, Z.; Jiang, G.; Chen, Y.; Wang, J.; Muhammad, I.; Zhang, L.; Wang, R.; Liu,
F.; Li, R.; Qian, F.; Li, J., Salidroside attenuates colistin-induced neurotoxicity in RSC96
Schwann cells through PI3K/Akt pathway. Chem Biol Interact 2017, 271, 67-78.
67. Zhang, L.; Xie, D.; Chen, X.; Hughes, M. L.; Jiang, G.; Lu, Z.; Xia, C.; Li, L.;
Wang, J.; Xu, W.; Sun, Y.; Li, R.; Wang, R.; Qian, F.; Li, J.; Li, J., p53 Mediates Colistin-
Induced Autophagy and Apoptosis in PC-12 Cells. Antimicrob Agents Chemother 2016,
60 (9), 5294-301.
68. Zhang, L.; Zhao, Y.; Ding, W.; Jiang, G.; Lu, Z.; Li, L.; Wang, J.; Li, J.; Li, J.,
Autophagy regulates colistin-induced apoptosis in PC-12 cells. Antimicrob Agents
Chemother 2015, 59 (4), 2189-97.
69. Jiang, H.; Lv, P.; Li, J.; Wang, H.; Zhou, T.; Liu, Y.; Lin, W., Baicalin inhibits
colistin sulfate-induced apoptosis of PC12 cells. Neural Regen Res 2013, 8 (28), 2597-
604.
70. Jiang, H.; Li, J.; Zhou, T.; Wang, C.; Zhang, H.; Wang, H., Colistin-induced
apoptosis in PC12 cells: involvement of the mitochondrial apoptotic and death receptor
pathways. International journal of molecular medicine 2014, 33 (5), 1298-304.
71. Fulda, S.; Debatin, K. M., Extrinsic versus intrinsic apoptosis pathways in
anticancer chemotherapy. Oncogene 2006, 25 (34), 4798-811.
72. Tait, S. W.; Green, D. R., Mitochondria and cell death: outer membrane
permeabilization and beyond. Nat Rev Mol Cell Biol 2010, 11 (9), 621-32.
Page 29 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
30
73. Chang, L.; Karin, M., Mammalian MAP kinase signalling cascades. Nature 2001,
410 (6824), 37-40.
74. Kruse, J. P.; Gu, W., Modes of p53 regulation. Cell 2009, 137 (4), 609-22.
75. Akeno, N.; Miller, A. L.; Ma, X.; Wikenheiser-Brokamp, K. A., p53 suppresses
carcinoma progression by inhibiting mTOR pathway activation. Oncogene 2015, 34 (5),
589-99.
76. Akhter, R.; Sanphui, P.; Das, H.; Saha, P.; Biswas, S. C., The regulation of p53
up-regulated modulator of apoptosis by JNK/c-Jun pathway in beta-amyloid-induced
neuron death. J Neurochem 2015, 134 (6), 1091-103.
77. Wada, T.; Penninger, J. M., Mitogen-activated protein kinases in apoptosis
regulation. Oncogene 2004, 23 (16), 2838-49.
78. Sugiura, Y.; Hiramatsu, K.; Hamauzu, R.; Motoki, T.; Miyazaki, M.; Uto, H.;
Tsubouchi, H.; Tanaka, S.; Gohda, E., Mitogen-activated protein kinases-dependent
induction of hepatocyte growth factor production in human dermal fibroblasts by the
antibiotic polymyxin B. Cytokine 2012, 60 (1), 205-11.
79. Dai, C.; Li, J.; Tang, S.; Li, J.; Xiao, X., Colistin-induced nephrotoxicity in mice
involves the mitochondrial, death receptor, and endoplasmic reticulum pathways.
Antimicrob Agents Chemother 2014, 58 (7), 4075-85.
80. Wadhwa, B.; Makhdoomi, U.; Vishwakarma, R.; Malik, F., Protein kinase B:
emerging mechanisms of isoform-specific regulation of cellular signaling in cancer.
Anticancer Drugs 2017, 28 (6), 569-580.
81. Calap-Quintana, P.; Soriano, S.; Llorens, J. V.; Al-Ramahi, I.; Botas, J.; Molto, M.
D.; Martinez-Sebastian, M. J., TORC1 Inhibition by Rapamycin Promotes Antioxidant
Defences in a Drosophila Model of Friedreich's Ataxia. PLoS One 2015, 10 (7),
e0132376.
82. Zhao, W.; Pan, X.; Li, T.; Zhang, C.; Shi, N., Lycium barbarum Polysaccharides
Protect against Trimethyltin Chloride-Induced Apoptosis via Sonic Hedgehog and
PI3K/Akt Signaling Pathways in Mouse Neuro-2a Cells. Oxid Med Cell Longev 2016,
2016, 9826726.
83. Sun, G.; Wang, X.; Li, T.; Qu, S.; Sun, J., Taurine attenuates acrylamide-induced
apoptosis via a PI3K/AKT-dependent manner. Hum Exp Toxicol 2018,
960327118765335.
84. Hu, M.; Li, F.; Wang, W., Vitexin protects dopaminergic neurons in MPTP-
induced Parkinson's disease through PI3K/Akt signaling pathway. Drug Des Devel Ther
2018, 12, 565-573.
85. Hu, M.; Liu, Z.; Lv, P.; Wang, H.; Zhu, Y.; Qi, Q.; Xu, J., Autophagy and
Akt/CREB signalling play an important role in the neuroprotective effect of nimodipine in
a rat model of vascular dementia. Behav Brain Res 2017, 325 (Pt A), 79-86.
86. Singh, A. K.; Kashyap, M. P.; Tripathi, V. K.; Singh, S.; Garg, G.; Rizvi, S. I.,
Neuroprotection Through Rapamycin-Induced Activation of Autophagy and
PI3K/Akt1/mTOR/CREB Signaling Against Amyloid-beta-Induced Oxidative Stress,
Synaptic/Neurotransmission Dysfunction, and Neurodegeneration in Adult Rats. Mol
Neurobiol 2017, 54 (8), 5815-5828.
87. Herkel, J.; Schrader, J.; Erez, N.; Lohse, A. W.; Cohen, I. R., Activation of the
Akt-CREB signaling axis by a proline-rich heptapeptide confers resistance to stress-
induced cell-death and inflammation. Immunology 2017.
Page 30 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
31
88. Singh, A. K.; Kashyap, M. P.; Tripathi, V. K.; Singh, S.; Garg, G.; Rizvi, S. I.,
Neuroprotection Through Rapamycin-Induced Activation of Autophagy and
PI3K/Akt1/mTOR/CREB Signaling Against Amyloid-beta-Induced Oxidative Stress,
Synaptic/Neurotransmission Dysfunction, and Neurodegeneration in Adult Rats. Mol
Neurobiol 2016.
89. Wang, Q.; Sun, G.; Gao, C.; Feng, L.; Zhang, Y.; Hao, J.; Zuo, E.; Zhang, C.; Li,
S.; Piao, F., Bone marrow mesenchymal stem cells attenuate 2,5-hexanedione-induced
neuronal apoptosis through a NGF/AKT-dependent pathway. Sci Rep 2016, 6, 34715.
90. Clewes, O.; Fahey, M. S.; Tyler, S. J.; Watson, J. J.; Seok, H.; Catania, C.; Cho,
K.; Dawbarn, D.; Allen, S. J., Human ProNGF: biological effects and binding profiles at
TrkA, P75NTR and sortilin. J Neurochem 2008, 107 (4), 1124-35.
91. Zhao, J.; Cheng, Y. Y.; Fan, W.; Yang, C. B.; Ye, S. F.; Cui, W.; Wei, W.; Lao, L.
X.; Cai, J.; Han, Y. F.; Rong, J. H., Botanical drug puerarin coordinates with nerve
growth factor in the regulation of neuronal survival and neuritogenesis via activating
ERK1/2 and PI3K/Akt signaling pathways in the neurite extension process. CNS
Neurosci Ther 2015, 21 (1), 61-70.
92. Salinas, M.; Diaz, R.; Abraham, N. G.; Ruiz de Galarreta, C. M.; Cuadrado, A.,
Nerve growth factor protects against 6-hydroxydopamine-induced oxidative stress by
increasing expression of heme oxygenase-1 in a phosphatidylinositol 3-kinase-
dependent manner. J Biol Chem 2003, 278 (16), 13898-904.
93. Kamata, H.; Tanaka, C.; Yagisawa, H.; Hirata, H., Nerve growth factor and
forskolin prevent H2O2-induced apoptosis in PC12 cells by glutathione independent
mechanism. Neuroscience letters 1996, 212 (3), 179-82.
94. Dai, C. X. X. L. J., Effects of nerve growth factor on colistin induced peripheral
neurotoxicity in mice. In China Animal Husbandry and Veterinary Society Veterinary
Pharmacology and Toxicology Branch Twelfth Symposium, Zhengzhou, China, 2013.
95. Finkbeiner, S., CREB couples neurotrophin signals to survival messages. Neuron
2000, 25 (1), 11-4.
96. Luo, D.; Zhao, J.; Cheng, Y.; Lee, S. M.; Rong, J., N-Propargyl Caffeamide
(PACA) Ameliorates Dopaminergic Neuronal Loss and Motor Dysfunctions in MPTP
Mouse Model of Parkinson's Disease and in MPP(+)-Induced Neurons via Promoting
the Conversion of proNGF to NGF. Mol Neurobiol 2018, 55 (3), 2258-2267.
97. Marino, G.; Niso-Santano, M.; Baehrecke, E. H.; Kroemer, G., Self-consumption:
the interplay of autophagy and apoptosis. Nat Rev Mol Cell Biol 2014, 15 (2), 81-94.
98. Altman, B. J.; Rathmell, J. C., Autophagy: not good OR bad, but good AND bad.
Autophagy 2009, 5 (4), 569-70.
99. Klionsky, D. J. et al., Guidelines for the use and interpretation of assays for
monitoring autophagy (3rd edition). Autophagy 2016, 12 (1), 1-222.
100. Hu, J.; Kholmukhamedov, A.; Lindsey, C. C.; Beeson, C. C.; Jaeschke, H.;
Lemasters, J. J., Translocation of iron from lysosomes to mitochondria during
acetaminophen-induced hepatocellular injury: Protection by starch-desferal and
minocycline. Free Radic Biol Med 2016, 97, 418-426.
101. Yun, B.; Azad, M. A.; Nowell, C. J.; Nation, R. L.; Thompson, P. E.; Roberts, K.
D.; Velkov, T.; Li, J., Cellular Uptake and Localization of Polymyxins in Renal Tubular
Cells Using Rationally Designed Fluorescent Probes. Antimicrob Agents Chemother
2015, 59 (12), 7489-96.
Page 31 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
32
102. Li, D.; Dai, C.; Yang, X.; Li, B.; Xiao, X.; Tang, S., GADD45a Regulates
Olaquindox-Induced DNA Damage and S-Phase Arrest in Human Hepatoma G2 Cells
via JNK/p38 Pathways. Molecules 2017, 22 (1).
103. Jacobs, M.; Gregoire, N.; Megarbane, B.; Gobin, P.; Balayn, D.; Marchand, S.;
Mimoz, O.; Couet, W., Population Pharmacokinetics of Colistin Methanesulfonate and
Colistin in Critically Ill Patients with Acute Renal Failure Requiring Intermittent
Hemodialysis. Antimicrob Agents Chemother 2016, 60 (3), 1788-93.
104. Zhao, X. C.; Zhang, L.; Yu, H. X.; Sun, Z.; Lin, X. F.; Tan, C.; Lu, R. R., Curcumin
protects mouse neuroblastoma Neuro-2A cells against hydrogen-peroxide-induced
oxidative stress. Food Chem 2011, 129 (2), 387-394.
105. Sahin, K.; Orhan, C.; Tuzcu, Z.; Tuzcu, M.; Sahin, N., Curcumin ameloriates heat
stress via inhibition of oxidative stress and modulation of Nrf2/HO-1 pathway in quail.
Food Chem Toxicol 2012, 50 (11), 4035-4041.
106. Tokac, M.; Taner, G.; Aydin, S.; Ozkardes, A. B.; Dundar, H. Z.; Taslipinar, M. Y.;
Arikok, A. T.; Kilic, M.; Basaran, A. A.; Basaran, N., Protective effects of curcumin
against oxidative stress parameters and DNA damage in the livers and kidneys of rats
with biliary obstruction. Food Chem Toxicol 2013, 61, 28-35.
107. Ghosh, S.; Banerjee, S.; Sil, P. C., The beneficial role of curcumin on
inflammation; diabetes and neurodegenerative disease: A recent update. Food Chem
Toxicol 2015, 83, 111-124.
108. Aggarwal, B. B., Surh, Y.-J. and Shishodia, S., The Molecular Targets and
Therapeutic Uses of Curcumin in Health and Disease. Springer US, New York: 2007;
Vol. 595.
109. Gou, M.; Men, K.; Shi, H.; Xiang, M.; Zhang, J.; Song, J.; Long, J.; Wan, Y.; Luo,
F.; Zhao, X.; Qian, Z., Curcumin-loaded biodegradable polymeric micelles for colon
cancer therapy in vitro and in vivo. Nanoscale 2011, 3 (4), 1558-67.
110. Gupta, S. C.; Patchva, S.; Aggarwal, B. B., Therapeutic Roles of Curcumin:
Lessons Learned from Clinical Trials. Aaps J 2013, 15 (1), 195-218.
111. Edrees, N. E.; Galal, A. A. A.; Abdel Monaem, A. R.; Beheiry, R. R.; Metwally, M.
M. M., Curcumin alleviates colistin-induced nephrotoxicity and neurotoxicity in rats via
attenuation of oxidative stress, inflammation and apoptosis. Chem Biol Interact 2018,
294, 56-64.
112. Rojewska, E.; Makuch, W.; Przewlocka, B.; Mika, J., Minocycline prevents
dynorphin-induced neurotoxicity during neuropathic pain in rats. Neuropharmacology
2014, 86, 301-10.
113. Du, Y.; Ma, Z.; Lin, S.; Dodel, R. C.; Gao, F.; Bales, K. R.; Triarhou, L. C.;
Chernet, E.; Perry, K. W.; Nelson, D. L.; Luecke, S.; Phebus, L. A.; Bymaster, F. P.;
Paul, S. M., Minocycline prevents nigrostriatal dopaminergic neurodegeneration in the
MPTP model of Parkinson's disease. Proc Natl Acad Sci U S A 2001, 98 (25), 14669-
74.
114. Kraus, R. L.; Pasieczny, R.; Lariosa-Willingham, K.; Turner, M. S.; Jiang, A.;
Trauger, J. W., Antioxidant properties of minocycline: neuroprotection in an oxidative
stress assay and direct radical-scavenging activity. J Neurochem 2005, 94 (3), 819-27.
115. Schildknecht, S.; Pape, R.; Muller, N.; Robotta, M.; Marquardt, A.; Burkle, A.;
Drescher, M.; Leist, M., Neuroprotection by minocycline caused by direct and specific
scavenging of peroxynitrite. J Biol Chem 2011, 286 (7), 4991-5002.
Page 32 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
33
116. Tikka, T. M.; Vartiainen, N. E.; Goldsteins, G.; Oja, S. S.; Andersen, P. M.;
Marklund, S. L.; Koistinaho, J., Minocycline prevents neurotoxicity induced by
cerebrospinal fluid from patients with motor neurone disease. Brain 2002, 125 (Pt 4),
722-31.
117. Tan, T. Y.; Ng, L. S.; Tan, E.; Huang, G., In vitro effect of minocycline and colistin
combinations on imipenem-resistant Acinetobacter baumannii clinical isolates. J
Antimicrob Chemother 2007, 60 (2), 421-3.
118. Bowers, D. R.; Cao, H.; Zhou, J.; Ledesma, K. R.; Sun, D.; Lomovskaya, O.;
Tam, V. H., Assessment of minocycline and polymyxin B combination against
Acinetobacter baumannii. Antimicrob Agents Chemother 2015, 59 (5), 2720-5.
119. Ritchie, D. J.; Garavaglia-Wilson, A., A review of intravenous minocycline for
treatment of multidrug-resistant Acinetobacter infections. Clin Infect Dis 2014, 59 Suppl
6, S374-80.
120. Rodriguez, C. H.; Nastro, M.; Vay, C.; Famiglietti, A., In vitro activity of
minocycline alone or in combination in multidrug-resistant Acinetobacter baumannii
isolates. J Med Microbiol 2015, 64 (10), 1196-200.
121. Liang, W.; Liu, X. F.; Huang, J.; Zhu, D. M.; Li, J.; Zhang, J., Activities of colistin-
and minocycline-based combinations against extensive drug resistant Acinetobacter
baumannii isolates from intensive care unit patients. BMC Infect Dis 2011, 11, 109.
122. Zou, Y.; Smith Iii, A. B., Total synthesis of architecturally complex indole
terpenoids: strategic and tactical evolution. J Antibiot (Tokyo) 2017.
123. Li, Y.; Gao, Z. X.; Dong, F. T.; Cheng, Y. Y., Quantitative Determination of
Quinocetone and Its Metabolites in Chicken Plasma by Liquid Chromatography/Tandem
Mass Spectrometry. J Aoac Int 2008, 91 (6), 1494-1498.
124. Zhang, L.; Yu, H.; Zhao, X.; Lin, X.; Tan, C.; Cao, G.; Wang, Z., Neuroprotective
effects of salidroside against beta-amyloid-induced oxidative stress in SH-SY5Y human
neuroblastoma cells. Neurochem Int 2010, 57 (5), 547-55.
125. Efeyan, A.; Comb, W. C.; Sabatini, D. M., Nutrient-sensing mechanisms and
pathways. Nature 2015, 517 (7534), 302-10.
126. Caccamo, A.; Maldonado, M. A.; Majumder, S.; Medina, D. X.; Holbein, W.;
Magri, A.; Oddo, S., Naturally secreted amyloid-beta increases mammalian target of
rapamycin (mTOR) activity via a PRAS40-mediated mechanism. J Biol Chem 2011, 286
(11), 8924-32.
127. Maiese, K., Targeting molecules to medicine with mTOR, autophagy and
neurodegenerative disorders. Br J Clin Pharmacol 2016, 82 (5), 1245-1266.
128. Xu, C.; Zhang, H.; Liu, C.; Zhu, Y.; Wang, X.; Gao, W.; Huang, S.; Chen, L.,
Rapamycin inhibits Erk1/2-mediated neuronal apoptosis caused by cadmium.
Oncotarget 2015, 6 (25), 21452-67.
129. Xu, C.; Wang, X.; Zhu, Y.; Dong, X.; Liu, C.; Zhang, H.; Liu, L.; Huang, S.; Chen,
L., Rapamycin ameliorates cadmium-induced activation of MAPK pathway and neuronal
apoptosis by preventing mitochondrial ROS inactivation of PP2A. Neuropharmacology
2016, 105, 270-284.
130. O'Reilly, K. E.; Rojo, F.; She, Q. B.; Solit, D.; Mills, G. B.; Smith, D.; Lane, H.;
Hofmann, F.; Hicklin, D. J.; Ludwig, D. L.; Baselga, J.; Rosen, N., mTOR inhibition
induces upstream receptor tyrosine kinase signaling and activates Akt. Cancer Res
2006, 66 (3), 1500-1508.
Page 33 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
34
131. Lehigh, K. M.; West, K. M.; Ginty, D. D., Retrogradely Transported TrkA
Endosomes Signal Locally within Dendrites to Maintain Sympathetic Neuron Synapses.
Cell Rep 2017, 19 (1), 86-100.
Page 34 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
35
Figures and Captions
Figure 1: Chemical structures of polymyxin B, colistin and colistin
methanesulphonate (CMS). Thr: threonine; Leu: leucine; Phe: phenylalanine; Dab:,-
diaminobutyric acid.
Page 35 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
36
Figure 2. Schematic diagram depicting the major apoptosis pathways involved in
colistin-induced neurotoxicity. Colistin-induced apoptosis involves oxidative stress, the
activation of death receptor and mitochondrial pathways (aberration of the Bax/ Bcl2
ratio, loss of membrane potential and production of reactive oxygen species [ROS]) and
activation of caspase-3, -8, and -9. The activation of p53, c-Jun N-terminal kinases
(JNKs)/MAPK pathways and the inhibition of the PI3K/Akt pathway may contribute to
colistin-induced apoptosis. These pathways have been demonstrated to be targeted
regulated by neuro-protective agents including curcumin, rapamycin and minocycline.
Page 36 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
37
Figure 3. Proposed model for the crosstalk between p53 and JNK in the regulation of
colistin-induced apoptosis and autophagy in neuronal cells. P53 is a transcription factor
that primarily induces the expression pro-apoptotic and pro-autophagic genes including
DRAM, PUMA, AMPK and Bax. JNK activation can promote mitochondrial ROS
production by up-regulating the Bax/Bcl-2 ratio and p53 expression, which eventuates in
apoptosis. The activation of p53 and JNK can both promote autophagy, which protects
against colistin-induced apoptotic cell death.
Page 37 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
38
Figure 4. The role of the NGF and PI3K/Akt pathways in colistin-induced neurotoxicity.
In the neuronal cell, colistin could inhibit the expression of Akt, then down-regulated the
expression of transcription factor CREB (cAMP response element-binding protein),
which is a well-known neuronal cell pro-survival factor. CREB can translocate to the
nucleus and bind to cAMP response elements, expressed the target gene NGF. Colistin
potentially inhibits the NGF- p75NTR pathway, which creates a ‘perfect’ vicious cycle by
the inhibition of Akt pathway by colistin.
Page 38 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
45x17mm (300 x 300 DPI)
Page 39 of 39
ACS Paragon Plus Environment
ACS Chemical Neuroscience
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
As a potent neurotoxic agent, acrylamide (ACR) is formed in food processing at higher temperature. Taurine (TAU), a nonessential amino acid, is used to cure neurodegenerative disorders, followed by activation of the phosphatidylinositol 3-kinase/protein kinase B (PI3K/AKT) signaling pathway. In this article, we certified that antiapoptotic efficacy of TAU in vivo and vitro. ACR-treated rats received TAU by drinking water 2 weeks after ACR intoxication. The results showed that in treated rats, TAU alleviated ACR-induced neuronal apoptosis, which was associated with the activation of PI3K/AKT signaling pathway. TAU attenuated apoptosis caused by ACR through observing terminal deoxynucleotidyl transferase-mediated dUTP nick end labeling (TUNEL)-positive cells, measure of protein expression of Bcl-2, Bax, and caspase 3 activity. TAU-induced antiapoptotic effect is PI3K/AKT-dependent, which was proved in ACR-intoxicated ventral spinal cord 4.1 cells in the presence of AKT inhibitor, MK-2206. Therefore, our results demonstrated that TAU-attenuated ACR-induced apoptosis in vivo through a PI3K/AKT-dependent manner provided new sights in the molecular mechanism of TAU protection against ACR-induced neurotoxicity.
Article
Full-text available
Parkinson’s disease (PD) is a progressive neurodegenerative disease which is characterized by the degeneration of dopaminergic neurons in the substantia nigra pars compacta (SNpc). Methods In this study, the neuroprotective effect of vitexin (Vit), a flavonoid compound isolated from Crataegus pinnatifida Bunge was examined in PD models both in vitro and in vivo. Results On SH-SY5Y cells, methyl-4-phenylpyridine (MPP⁺) treatment suppressed cell viability, induced apoptosis, and increased Bax/Bcl-2 ratio and caspase-3 activity. However, Vit improved these parameters induced by MPP⁺ treatment significantly. Further study disclosed that Vit enhanced the phosphorylation of PI3K and Akt which was downregulated by MPP⁺ in SH-SY5Y cells, the effect of which could be blocked by PI3K inhibitor LY294002 and activated by PI3K activator IGF-1. Moreover, results from the pole test and traction test suggested that Vit pretreatment prevented bradykinesia and alleviated the initial lesions caused by 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) in MPTP-treated mouse PD model. Vit also enhanced the activation of PI3K and Akt and suppressed the ratio of Bax/Bcl-2 and caspase-3 activity in MPTP-treated mice. Conclusion Taken together, this study demonstrated that Vit protected dopaminergic neurons against MPP⁺/MPTP-induced neurotoxicity through the activation of PI3K/Akt signaling pathway. Our findings may facilitate the clinical application of Vit in the therapy of PD.
Article
Full-text available
This study aimed to investigate the mechanism of p53 in regulating colistin-induced autophagy in PC-12 cells. Importantly, cells were treated with 125 μg/ml colistin for 12 and 24 h after transfection with p53 siRNA or recombinant plasmid. The hallmarks of autophagy and apoptosis were examined by real-time PCR and western blot, fluorescence/immunofluorescence microscopy, and electron microscopy. The results showed that silencing of p53 leads to down-regulation of Atg5 and beclin1 for 12 h while up-regulation at 24 h and up-regulation of p62 noted. The ratio of LC3-II/I and autophagic vacuoles were significantly increased at 24 h, but autophagy flux was blocked. The cleavage of caspase3 and PARP (poly ADP-ribose polymerase) were enhanced, while PC-12-sip53 cells exposed to 3-MA showed down-regulation of apoptosis. By contrast, the expression of autophagy-related genes and protein reduced in p53 overexpressing cells following a time dependent manner. Meanwhile, there was an increase in the expression of activated caspase3 and PARP, condensed and fragmented nuclei were evident. Conclusively, the data supported that silencing of p53 promotes impaired autophagy, which acts as a pro-apoptotic induction factor in PC-12 cells treated with colistin for 24 h, and overexpression of p53 inhibits autophagy and accelerates apoptosis. Hence, it has been suggested that p53 could not act as a neuro-protective target in colistin-induced neurotoxicity.
Article
Full-text available
Very high doses of polymyxin B (PMB) have been addressed in in vitro experiments against carbapenem-resistant Gram-negative bacilli as a strategy to improve bacterial killing and suppress resistance emergence. However, the toxicities of very high doses in patients are unknown. This is a retrospective cohort study assessing patients receiving >3mg/kg/day or a total dose ≥250mg/day of PMB. The main outcomes were severe infusion-related adverse event according to Common Terminology Criteria for Adverse Criteria and renal failure category of RIFLE criteria for acute kidney injury (AKI) during treatment. A total of 222 were included for analysis of infusion-related events. The mean PMB dose was 3.61±0.97mg/kg/day (median total dose/day=268 mg). Severe infusion-related adverse events occurred in two patients determining an incidence of 0.9% (95% Confidence Interval, 0.2-3.2): one was classified as a life-threatening and one as severe adverse effect. Renal failure was analyzed in 115 patients: 25 (21.7%) patients presented renal failure (54 [47.0%] developed any degree of AKI, categorized as Risk [27.8%]; Injury [25.9%] and Failure [46.3%]). Vasoactive drug, concomitant nephrotoxic drugs and baseline creatinine clearance were independent risk factors for renal failure. Neither PMB daily dose scaled by body weight nor total daily dose were associated with renal failure. In-hospital mortality was 60% (134 patients): 26% (57 patients) occurred during treatment and none during infusion. Our data suggest that high dose schemes have acceptable safety profile and could be further tested in clinical trials assessing strategies to improve patients' outcomes and minimize the emergence of PMB resistance.
Article
Full-text available
Colistin is a polycation antibiotic used for the treatment of multidrug-resistance (MDR) gram-negative infections; nevertheless, its use is often limited by the high incidence of renal damage. The mechanism underlying colistin-induced nephrotoxicity is not known, but perhaps related to its accumulation in the renal cortex upon extensive reabsorption from the nascent urine. Because little is known about the membrane transport of colistin, the purpose of the present study was to characterize better the transport system involved in colistin renal handling by using HEK293 cells stably transfected with the main organic cation transporters expressed at the apical membrane of the proximal tubule. [(14)C]Colistin was transported by the carnitine/organic cation transporter 2 (OCTN2, SLC22A5) but not by the organic cation transporter 1 (OCT1) and N1 (OCTN1). Non-labeled colistin inhibited the OCTN2-mediated transport of [(3)H]L-carnitine in a non-competitive manner and that of [(14)C]tetraethylammonium bromide ([(14)C]TEA) in a competitive manner. Unlike that of [(3)H]L-carnitine, the [(14)C]colistin OCTN2-mediated uptake was Na(+)-independent. When endogenous OCTN2-mediated colistin transport was inhibited by co-incubation with L-carnitine, primary mouse proximal tubular cells were fully protected from colistin toxicity, suggesting that colistin toxicity occurred upon intracellular accumulation.
Article
Neurotoxicity is an unwanted side effect patient’s experience when receiving parenteral colistin therapy. The development of effective neuroprotective agents that can be co-administered during colistin therapy remains a priority area in antimicrobial chemotherapy. The present study aimed to investigate the protective effect of nerve growth factor (NGF) against colistin-induced peripheral neurotoxicity using a murine model. C57BL/6 mice were randomly divided into the following 6 groups: (i) untreated control, (ii) NGF alone 36 μg/kg/day (administered intraperitoneally), (iii) colistin alone (18 mg/kg/day administered intraperitoneally) and (iv-vi) colistin (18 mg/kg/day) plus NGF (9, 18 and 36 μg/kg/day). After treatment for 7 days, neurobehavioral and electrophysiology changes, histopathological assessments of sciatic nerve damage, and oxidative stress biomarkers were examined. The mRNA expression levels of Nrf2, HO-1, Akt, Bax, caspase-3 and -9 were assessed using quantitative RT-PCR. The results showed that across all the groups wherein NGF was co-administered with colistin, a marked attenuation of colistin-induced sciatic nerve damage and improved sensory and motor function were observed. In comparison to the colistin only treatment group, animals that received NGF displayed up-regulated Nrf2 and HO-1 mRNA expression levels and down-regulated Bax, caspase-3 and -9 mRNA expression levels. In summary, our study reveals that NGF co-administration protects against colistin-induced peripheral neurotoxicity via the activation of Akt and Nrf2/HO-1 pathways and inhibition of oxidative stress. This study highlights the potential clinical application of NGF as a neuroprotective agent for co-administration during colistin therapy.
Article
Colistin is an effective antibiotic against multidrug-resistant (MDR) gram-negative bacterial infections; however, nephrotoxic and neurotoxic effects are fundamental dose-limiting factors for this treatment. This study was conducted to assess the potential protective effects of curcumin, a phenolic constituent of turmeric, against colistin-induced nephrotoxicity and neurotoxicity, and the possible mechanisms underlying any effect. Twenty-four adult male albino rats were randomly classified into 4 equal groups; the control group (orally received saline solution), the curcumin-treated group (orally administered 200 mg curcumin/kg/day), the colistin-treated group (IP administered 300,000 IU colistin/kg/day) and the concurrent group (orally received 200 mg curcumin/kg/day concurrently with colistin injection); all rats were treated for 6 successive days. Colistin administration significantly increased serum creatinine, urea and uric acid levels as well as brain gamma butyric acid (GABA) concentrations. In renal and brain tissues, colistin significantly increased malondialdehyde (MDA), nitric oxide (NO), tumor necrosis factor-α (TNF-α), interleukin-6 (IL-6), and caspase-3 expression levels. In addition, colistin significantly decreased catalase (CAT), glutathione (GSH), and B-cell lymphoma 2 (Bcl-2) expressions. Curcumin administration in colistin-treated rats partially restored each of these altered biochemical, antioxidant, inflammatory and apoptotic markers. Histopathological changes in renal and brain tissues were also alleviated by curcumin co-treatment. Our study reveals a critical role of oxidative damage, inflammation and apoptosis in colistin-induced nephrotoxicity and neurotoxicity and showed that they were markedly ameliorated by curcumin co-administration. Therefore, curcumin could represent a promising agent for prevention of colistin-induced nephrotoxicity and neurotoxicity.
Article
Our previous studies showed colistin-induced neurotoxicity involves apoptosis and oxidative damage. The present study demonstrates a neuroprotective effect of rapamycin against colistin-induced neurotoxicity in vitro and in vivo. In a mouse model, colistin treatment (18 mg/kg/d; 14 days) produced marked neuronal mitochondria damage in the cerebral cortex and increased activation of caspase-9 and -3. Rapamycin co-treatment (2.5 mg/kg/d) effectively reduced this neurotoxic effect. In an in vitro mouse neuroblastoma-2a (N2a) cell culture model, rapamycin pre-treatment (500 nM) significantly decreased colistin (200 μM) induced cell death from ~50% to 72%. Moreover, rapamycin showed a marked neuroprotective effect in the N2a cells by decreasing intracellular reactive oxygen species (ROS) production and by up-regulating the activities of the anti-ROS enzymes superoxide dismutase and catalase, and recovering GSH levels to normal. Moreover, rapamycin pre-treatment protected against colistin-induced mitochondrial dysfunction, caspase activation and subsequent apoptosis by up-regulating autophagy and activating the Akt/CREB, NGF and Nrf2 pathways, while inhibiting p53 signaling. Taken together, this is the first study to demonstrate that rapamycin protects against colistin-induced neurotoxicity by activating autophagy, inhibiting oxidative stress, mitochondria dysfunction and apoptosis. Our data highlight that regulating autophagy to rescue neurons from apoptosis may become a new targeted therapy to relieve the adverse neurotoxic effects associated with colistin therapy.
Book
Curcumin is derived from the root of the plant Curcuma longa (also called turmeric) and its medicinal uses have been described for over 5000 years. More than 1500 papers published within last half a century has revealed that curcumin has a potential in the treatment of wide variety of inflammatory diseases including cancer, diabetes, cardiovascular diseases, arthritis, Alzheimer, psoriasis etc, through modulation of numerous molecular targets. This is the first monograph to deal specifically with this subject.
Article
Indole terpenes have attracted the interests of synthetic chemists due to their complex architectures and potent biological activities. Examples of total syntheses of several indole terpenes were reviewed in this article to honor Professor KC Nicolaou.