You are on page 1of 10

Pharmacology & Therapeutics 141 (2014) 150–159

Contents lists available at ScienceDirect

Pharmacology & Therapeutics


journal homepage: www.elsevier.com/locate/pharmthera

Associate editor: S. Kennedy

Existing and potential therapeutic uses for N-acetylcysteine: The need for
conversion to intracellular glutathione for antioxidant benefits
Gordon F. Rushworth a, Ian L. Megson b,⁎
a
Highland Clinical Research Facility, Centre for Health Science, Old Perth Road, Inverness IV2 3JH, UK
b
Department of Diabetes and Cardiovascular Science, University of the Highlands and Islands, Centre for Health Science, Old Perth Road, Inverness IV2 3JH, UK

a r t i c l e i n f o a b s t r a c t

Keywords: N-acetyl-L-cysteine (NAC) has long been used therapeutically for the treatment of acetaminophen (paracetamol)
N-acetylcysteine overdose, acting as a precursor for the substrate (L-cysteine) in synthesis of hepatic glutathione (GSH) depleted
Glutathione through drug conjugation. Other therapeutic uses of NAC have also emerged, including the alleviation of clinical
Antioxidant symptoms of cystic fibrosis through cysteine-mediated disruption of disulfide cross-bridges in the glycoprotein
matrix in mucus.
More recently, however, a wide range of clinical studies have reported on the use of NAC as an antioxidant, most
notably in the protection against contrast-induced nephropathy and thrombosis. The results from these studies
are conflicting and a consensus is yet to be reached regarding the merits or otherwise of NAC in the antioxidant
setting.
This review seeks to re-evaluate the mechanism of action of NAC as a precursor for GSH synthesis in the context of its
activity as an “antioxidant”. Results from recent studies are examined to establish whether the pre-requisites for ef-
fective NAC-induced antioxidant activity (i.e. GSH depletion and the presence of functional metabolic pathways for
conversion of NAC to GSH) have received adequate consideration in the interpretation of the data. A key conclusion
is a reinforcement of the concept that NAC should not be considered to be a powerful antioxidant in its own right: its
strength is the targeted replenishment of GSH in deficient cells and it is likely to be ineffective in cells replete in GSH.
© 2013 Elsevier Inc. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
2. Biochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
3. Clinical pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
152
4. Therapeutic uses of N-acetyl-cysteine: Licenced and potential . . . . . . . . . . . . . . . . . . . . . . . 154
153
5. Discussion and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
157
Conflict of interest statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
154
Author declaration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
154
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
157

1. Introduction mucolytic agent in patients with cystic fibrosis (CF; Hurst et al., 1967).
The concept derived from the need to deliver reduced sulfhydryl moie-
N-acetyl-L-cysteine (NAC) is a drug that was first reported to have ties to effect the disruption of disulfide bridges within the glycoprotein
clinical benefit in the early 1960s, when it was shown to be an effective matrix of mucus in CF patients. The amino acid residue, L-cysteine (Cys),
represents an obvious candidate for such an agent, but unfortunately, it
is susceptible to metabolism and undergoes rapid oxidation in solution,
Abbreviations: NAC, N-acetyl-L-cysteine; GSH, glutathione; GCL, glutamate cysteine li-
gase; ROS, reactive oxygen species; GSSG, glutathione disulfide; GSH Px, glutathione generating the inactive disulfide, cystine (Cys–Cys). Acetylation of the
peroxidises; COPD, chronic obstructive pulmonary disease; NAPQI, N-aceytl-p- N-terminal of Cys was found to confer sufficient stability to the mole-
benzoquinonimine; IV, intravenous; CIN, contrast-induced nephropathy; IPF, idiopathic pul- cule to facilitate delivery of reduced sulfhydryl (thiol) moieties to
monary fibrosis; BAL, bronchoalveolar lavage; CF, cystic fibrosis; AKI, acute kidney injury. work effectively as a mucolytic agent in this clinical setting.
⁎ Corresponding author at: Department of Diabetes and Cardiovascular Science,
During the 1970s, a substantial sequence of studies involving potential
University of the Highlands and Islands, Centre for Health Science, Inverness IV2 3JH,
UK. Tel.: +44 1463 279562; fax: +44 1463 711245. sulfhydryl donor candidates was conducted in paracetamol poisoning.
E-mail address: ian.megson@uhi.ac.uk (I.L. Megson). However, other donors were either ineffective or provoked a significant

0163-7258/$ – see front matter © 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.pharmthera.2013.09.006
G.F. Rushworth, I.L. Megson / Pharmacology & Therapeutics 141 (2014) 150–159 151

number of adverse effects (Prescott et al., 1976). A new and important group would point to membrane permeation of intact NAC being
role for NAC emerged on account of subsequent studies investigating unlikely and studies with radio-labelled NAC would concur (Cotgreave
its therapeutic potential in the treatment of acetaminophen (paraceta- et al., 1987a). That said, low concentrations of NAC have been detected
mol; N-acetyl-p-aminophenol) poisoning (Prescott et al., 1977, 1979). in extracts of red blood cells from animals treated with NAC (Giustarini
The founding principle that underpinned the mechanism of action in et al., 2012), perhaps suggesting that a small proportion of NAC can per-
this setting was similar to that for CF: delivery of sulfhydryl moieties. meate the membrane, assuming that extracts were not contaminated
However, the mode of action of NAC in acetaminophen overdose was with extracellular medium. The alternative is that extracellular deacety-
thought to rely not only on the ability of NAC to offer some protection lation takes place and that Cys is taken into cells via amino acid trans-
against oxidation, but also through facilitation of rapid membrane per- porters. It remains unclear, therefore, as to the extent, means and
meability on account of the reduced polarity of the molecule compared timing of the process of de-acetylation of NAC, but the assumption is
to the parent amino acid, Cys. Cleavage of the acetyl group is thought to that, in the scenario whereby free intracellular reduced Cys is required
reveal free, reduced Cys, which is available for incorporation into the as a substrate for GSH synthesis, intact NAC penetrates the cell mem-
highly abundant intracellular antioxidant, glutathione (GSH). The ben- brane prior to hydrolysis to Cys in the intracellular environment, poten-
efit conveyed by NAC in the setting of acetaminophen overdose is to re- tially with the aid of N-deacetylases (Fig. 1). Data from an ongoing study
plenish hepatic GSH that has become depleted through the use of the (Sandilands et al., 2012) might help to clarify whether a substantial pro-
tripeptide in the drug detoxification process. That NAC remains the portion of intact NAC accesses cells instead of being cleaved in the extra-
treatment of choice for acetaminophen overdose more than 50 years cellular environment, as has been suggested previously (Cotgreave
after its first use is testament both to the importance of maintaining cel- et al., 1991). For mucolytic effects of NAC, deacetylation may not be a
lular GSH reserves and to the exceptional qualities of NAC in helping to prerequisite, given that the sulfhydryl is available for interaction with
replenish this key antioxidant when it has become acutely depleted. disulfide bridges irrespective of the presence of the acetyl group.
Despite this, the precise pharmacological mechanisms that underpin Glutathione (GSH) is a tripeptide that is synthesised and maintained
NAC activity are, as yet, not fully understood and may not even be at high (mM) concentrations in cells (Meister & Anderson, 1983).
entirely related to GSH repletion (Waring, 2012; Gosselin et al., 2013). The rate limiting step of the synthesis involves conjugation of Cys with
Since the 1980s, there has been a growing interest in the therapeutic L-glutamate (glutamate–cysteine ligase; GCL; (McPherson & Hardy,
potential of NAC in a range of diseases where oxidative stress is seen to 2012), while L-glycine is added in a subsequent synthetic step involving
be a driver and in which antioxidant effects might convey benefit. The GSH synthase (GS; Fig. 2). In its reduced form, GSH has a wide variety of
original premise for believing that NAC might be effective as an antiox- functions, from antioxidant protection to protein thiolation and drug
idant is unclear; perhaps its ability to drive synthesis of the powerful detoxification, often supported by specific enzymes (Fig. 2).
antioxidant, GSH, in hepatic cells rendered deficient through acetamin- Oxidative stress is a critical process that is implicit in the aetiology of a
ophen detoxification has been misconstrued as direct antioxidant capa- wide range of diseases, including cardiovascular disease (Le Brocq et al.,
bility, or maybe there is a general belief that all sulfhydryls will share 2008), diabetes (Stadler, 2012) and cancer (Sosa et al., 2013). GSH is a
the antioxidant power of GSH. However, the evidence regarding the an- critical intracellular antioxidant that helps to limit the impact of oxidative
tioxidant potential of NAC is that it is a relatively weak antioxidant: stress and to protect vital cellular components (lipids, proteins, DNA)
direct experiments to assess its antioxidant potential suggest that against harmful peroxidation. The antioxidant effects of GSH rely on
~10-fold more NAC is required compared to GSH to facilitate equivalent the presence of the free sulfyhydryl group as a ready source of reducing
oxygen-centred radical scavenging (Gibson et al., 2009), while the equivalents to quench radical species. As well as acting as a direct
ability of NAC to scavenge one of the major biologically relevant radical “sacrificial” scavenger of potentially harmful reactive oxygen species
species, superoxide, is not detectable (Aruoma et al., 1989). It is highly (ROS), GSH provides reducing equivalents to support the antioxidant
likely, therefore, that the vast majority of the antioxidant effects attrib-
uted directly to NAC are actually mediated by increased intracellular
GSH. The distinction is important because it means that certain condi-
tions might have to be satisfied in order for NAC to confer antioxidant
activity: first, the enzymatic machinery necessary for GSH synthesis
must be intact and expressed at sufficient levels and second, it is likely
that GSH might have to be depleted for NAC to have any beneficial
effect. Far from being impediments to the use of NAC in a range of
clinical indications, these conditions for NAC antioxidant activity should
be viewed as positive indicators for use and could open the way to strat-
ified approaches for application of NAC only in those patients likely to
benefit.
This review will provide a brief overview of the biochemistry and
clinical pharmacology associated with NAC activity, followed by an eval-
uation of the licenced uses of NAC in clinical conditions worldwide and
an appraisal of the potential of NAC in novel indications, with a bias
towards those with an antioxidant component to the suggested mode
of action.

2. Biochemistry

Irrespective of the clinical target, the role of NAC is to deliver sulfhy-


dryl moieties for utilisation in biological processes. NAC has advantages
over Cys in this respect because it is relatively resistant to oxidation
to the disulfide and was originally believed to have the capability to
cross cell membranes without the need for amino acid transporters on
account of the reduced charge imparted by the acetyl moiety. However,
residual polarity of the NAC molecule on account of the –SH and –COOH Fig. 1. N-acetyl-L-cysteine (NAC): a cysteine pro-drug.
152 G.F. Rushworth, I.L. Megson / Pharmacology & Therapeutics 141 (2014) 150–159

activity of GSH peroxidases (GPx; Pompella et al., 2003), a powerful de- therapeutic uses. In this respect, the location of ROS generation, the
fence against peroxides (R-OOH, e.g. hydrogen peroxide, HO-OH). GSH ROS species generated, the relative abundance of endogenous antioxi-
is oxidised to the thiyl radical (GS•) in the process, which rapidly dants in the locality, the rate constants of endogenous antioxidants for
dimerises to form the glutathione disulfide (GS-SG; Equation 1). the ROS generated, together with their relative concentrations, will all
be vital determinants of the success or failure of an administered antiox-
idant in helping to prevent cellular damage. Often, however, many of
these details are not known, making selection of an appropriate antiox-
idant for a specific condition difficult. Attempts to rank antioxidants
according to their known characteristics (rate constants and concentra-
tion) have been made, but are fraught with difficulties, not least because
of the variability in estimates for rate constants for reactions with radi-
cal species, as highlighted in (Winterbourne & Metodiewa, 1999). NAC,
Equation 1: Antioxidant effect of GSH/GPx on peroxides has particularly high rate constants for •OH (~1010 M−1 s−1), CO3•−,
•NO2 and HOCl (all ~107 M−1 s−1; Samuni et al., 2013), but it is fair to
GSH also reacts with another important cellular oxygen-centred free say that NAC does not rank highly amongst antioxidants in general, or
radical, superoxide (O2•−), but while the reaction oxidises GSH, it is even amongst endogenous thiols (Winterbourne & Metodiewa, 1999),
thought that superoxide is regenerated in the process (equation 2; particularly when concentration and location are taken into account.
Winterbourne & Metodiewa, 1999). In this regard, it is thought essential The case is most evident in the intracellular compartment, where the
that GSH acts in concert with superoxide dismutases (SOD), perhaps overwhelming antioxidant potential of GSH, ascorbate and specific anti-
explaining the co-location of the two antioxidants, particularly in mito- oxidant enzymes (e.g. superoxide dismutases, glutathione peroxidases,
chondria. The same process is central to the reaction of superoxide with catalase) ensure that the very low (if any) concentrations of NAC that
NAC, but not with Cys, where hydrogen peroxide is a product instead of persist inside cells are unlikely to contribute directly to antioxidant ca-
superoxide (Winterbourne & Metodiewa, 1999). pacity. The same is true in the plasma compartment in cases where
NAC is administered orally and concentrations only reach low μM; plas-
ma antioxidant capacity has been found to be unchanged by treatment
(Leelarungrayub et al., 2011). However, IV delivery might have the
capability to increase the plasma sulfhydryl load sufficiently to impact
on plasma antioxidant capacity. Plasma NAC can reach low mμM con-
centrations, compared with 3–5 μM GSH, ~250 μM cysteine, 600 μM
albumin sulfydryl (Cys 34), N50 μM ascorbate and ~400 μM urate. The
thiol pool and antioxidant potential is therefore dominated by albumin
(Turell et al., 2013) and there are few data available to indicate whether
intravenous NAC can influence plasma total antioxidant capacity. Per-
haps surprisingly, plasma antioxidant capacity with NAC is rarely mea-
sured and, when it is, the data fail to show a convincing impact
Equation 2: Reaction of thiols with superoxide: GSH is consumed, but (Buyukhatipoglu et al., 2010). The theoretical impact of NAC on plasma
superoxide is regenerated (adapted from Winterbourne & Metodiewa, antioxidant benefit is dependent on NAC remaining in the reduced state
1999). This process is the same for NAC, but different for cysteine, and on the blood being the principal source of the target ROS. Any direct
where H2O2 is generated instead of superoxide antioxidant effect could be supplemented by an indirect benefit that
might be realised through chelation of metal ions (Lodge et al., 1998;
One of the greatest strengths of GSH over most other antioxidant de- Joshi et al., 2010) capable of mediating Fenton Chemistry and formation
fences is the intracellular capability to regenerate the reduced form of hydroxyl radical, but again there are specific chelators for commonly
from the disulfide via the action of GSH reductase (GR). There is some occurring metals (e.g. iron and copper) in the blood (e.g. ferritin,
evidence to suggest that GSH can also contribute to recycling of the di- caeruloplasmin respectively) that would effectively compete with NAC
etary antioxidants vitamins C (Li et al., 2001; Montecinos et al., 2007), in this regard, unless already saturated or under-expressed. NAC has,
which has the redox potential to recycle vitamin E in turn (Halpner however, been shown in animal experiments to increase excretion
et al., 1998; May et al., 1998; Fig. 2). In this sense, GSH might be seen rates of some metal complexes (borate and chromate), but the mecha-
as the cornerstone of antioxidant defence. nism underpinning the effect is unclear (Banner et al., 1986). More
Critically, however, GSH synthesis is understood to be self-regulated comprehensive evidence is available to support a direct chelating
(Meister & Anderson, 1983; Griffith, 1999); that is to say that the rate- effect of NAC with methyl mercury to form S-conjugates (Zalups &
limiting enzyme for its synthesis (GCL) is inhibited by the product Ahmad, 2005), although at the cellular level, protection against methyl
(GSH; Fig. 2). In addition, expression of GSH-related enzymes, including mercury-induced toxicity is primarily modulated by intracellular GSH,
those involved in its synthesis, are subject to complex regulation driven which can be supplemented by NAC (Kaur et al., 2006; Becker &
by the intracellular redox state (Arrigo, 1999; Dickinson et al., 2004); Soliman, 2009). While NAC therefore has the capability to form S-
ageing and certain disease states are well-known to suppress GCL conjugates with some metals, it is unclear as to the importance of this
expression (Lu, 2009). These are important considerations with respect mechanism in driving any protective effects compared to intracellular
to the impact of NAC as a therapy: substrate might not be rate-limiting GSH acting either as a chelating agent itself or as an antioxidant to ame-
in conditions where GCL expression is suppressed. On the other hand, in liorate downstream oxidative stress. Indeed, there is some evidence to
cells where GSH is already abundant, feedback inhibition of synthesis suggest that formation of S-conjugates of mercury with NAC, Cys or ho-
could negate any benefits of NAC delivery that require GSH generation. mocysteine can facilitate uptake of mercury into renal tubular cells, with
the potential for detrimental effects (Zalups & Barfuss, 1998; Brandao
2.1. Considerations for N-acetylcysteine as a direct antioxidant et al., 2006). There are limited clinical data available to evaluate the po-
tential of NAC in treating metal toxicity (Blanusa et al., 2005); a case
The ability of a substance to act as an antioxidant in a biologically study from 1984 did suggest benefit of NAC by haemodialysis in methyl
relevant situation is a highly complex concept, but one that is central mercury poisoning (Lund et al., 1984), but the dearth of subsequent
to understanding the mode of action of NAC in many of its potential clinical data might point to a lack of clear benefit in this arena.
G.F. Rushworth, I.L. Megson / Pharmacology & Therapeutics 141 (2014) 150–159 153

Fig. 2. Impact of NAC on synthesis and utilisation pathways for GSH. De-ACase — deacetylases; GCL — glutathione cysteine ligase; GS — glutathione synthase; GPx — glutathione
peroxidase; GR — glutathione reductase; GST — glutathione-S-transferase; γ-GT — glutamyl transpeptidase.

2.2. Antioxidant and glutathione-independent effects of N-acetyl-cysteine enduring delivery mode for Cys; the initial indication for prescribing
was as a mucolytic because it was reported to break down the disulfide
A final consideration with respect to the antioxidant potential of bridges in mucus glycoproteins to reduce viscosity for those with em-
NAC is true of any antioxidant – how much is too much? The barrier physema predominant chronic obstructive pulmonary disease (COPD;
function of antioxidants dictates that higher concentrations are likely (Aitio, 2006). Oral and aerosol preparations of NAC are licenced for
to be more effective. However, it is also clear that very high concentra- use in some countries, but not worldwide e.g. UK has no licenced uses
tions of antioxidants at least have a theoretical potential to generate a for oral NAC, where carbocysteine (an established mucolytic with simi-
counter-intuitive pro-oxidant effect, probably mediated via reduction lar mechanism of action) is preferred for lung conditions. After oral ad-
of transition metal ions and enhancement of Fenton Chemistry ministration (typically 600–1200 mg pills up to 3 times daily), NAC is
(Halliwell, 1996). How important this effect is in vivo is not understood, rapidly absorbed, with peak plasma concentrations (low micromolar)
but it is likely that there is an optimal concentration for NAC in the plas- seen between 30 min and 1 h (Holdiness, 1991). Orally administered
ma, as there is thought to be for other antioxidants. NAC is absorbed in the small intestine and undergoes first-pass hepatic
Data from a wide range of in vitro studies suggest that NAC might metabolism to Cys, from which the liver synthesises GSH. Hepatic stores
have the capability to alter protein structure and/or function, perhaps of GSH are replenished before GSH is released into the plasma via a
via reduction of disulfide bridges and conformational change. Examples membrane transporter (Bartoli & Sies, 1978). Oral bioavailability is con-
include modulation of angiotensin II receptor binding in vascular sidered to be low (b10%), potentially on account of gut-wall metabolism
smooth muscle cells (Ullian et al., 2005) and altered cytokine affinity and high first-pass metabolism (Borgstrom et al., 1986; Olsson et al.,
of TNF-α receptors (Hayakawa et al., 2003). While the results from 1988). However, it is also possible that low plasma detection is due to
these studies are interesting from a theoretical perspective, the concen- rapid diffusion into cells and conversion to GSH, maintaining a constant
trations of NAC used are extremely high (typically ~10 mM and some- concentration gradient across the membrane.
times as high as ~100 mM), so their relevance to the in vivo situation The intravenous preparation of NAC bypasses first-pass and gut-wall
should be considered with some caution. metabolism, delivering sufficient concentrations (high micromolar/low
millimolar) expeditiously in acute acetaminophen overdose. It will
3. Clinical pharmacology be interesting to discover from an ongoing direct comparison of oral
versus intravenous NAC, whether intracellular GSH supplementation
While depletion of GSH is a feature of many disease states and its re- is greater with substantially higher NAC delivery via the intravenous
plenishment is desirable, administration of GSH per se is not considered route (Sandilands et al., 2012). There is some evidence to suggest that
optimal on account of its poor bioavailability and its limited ability to hepatocellular concentrations of NAC are actually increased on oral
cross the phospholipid bilayer of cells. Likewise, delivery of Cys suffers administration, compared with IV, as a result of first-pass portal circula-
from rapid oxidation to its disulfide, cystine, which has poor solubility tion (Green et al., 2013).
and renders the crucial sulfhydryl functional group at least temporarily Adverse effects experienced with NAC are somewhat dependent on
inaccessible. As a result, other means of sulfhydryl delivery have been the route of administration used. Mild reactions – such as nausea,
developed to circumvent this problem (Suzuki, 2009). NAC is the most vomiting and cutaneous reactions (pruritus and erythema) – are common
154 G.F. Rushworth, I.L. Megson / Pharmacology & Therapeutics 141 (2014) 150–159

with the IV preparation (Sandilands & Bateman, 2009). Severe systemic infection. The benefits of long and short-term administration of NAC
reactions are uncommon, but are thought to be non-immunological for bronchial hypersecretion in chronic bronchitis have been known
(non-IgE mediated) in nature and are therefore defined as anaphalactoid since the early 1980s (Multicentre Study Group, 1980; Tattersall et al.,
reactions which are reported in 20% of patients, where reactions are 1983). A subsequent systematic review, which focused specifically on
more likely to occur in patients with comparatively low acetaminophen NAC, acknowledged the beneficial effects of oral administration for
concentrations or past history of asthma (Pakravan et al., 2008; Waring treatment periods of 3–6 months for reducing exacerbations and im-
et al., 2008). Anaphalactoid symptoms include flushing, pruritus, proving symptoms of chronic bronchitis (Stey et al., 2000). However,
angiodema, bronchospasm and hypotension (Flanagan & Meredith, the results of the review were uncertain as to the long-term benefits
1991). There is an increased risk of anaphylactoid side effects with in all patients with chronic bronchitis. A subsequent systematic review
intravenous compared to oral delivery (Kanter, 2006), most likely on of mucolytics in chronic bronchitis or COPD began to stratify where mu-
account of the much higher plasma concentrations achieved. Indeed, colytics might be found to be most cost-effective, advocating their use in
the reactions are most prevalent immediately after the initial loading patients with severe COPD who are susceptible to repeated exacerba-
infusion. tions (Poole & Black, 2001).
Asthma is defined as a reversible airways obstruction where chronic
inflammation and airway hypersensitivity lead to symptoms of breath-
4. Therapeutic uses of N-acetyl-cysteine: Licenced and potential lessness and wheeze in patients. Inflammation in asthma has been in-
part shown to be mediated by ROS (Dworski, 2000). It has been postu-
4.1. Cystic fibrosis and other lung diseases lated that the actual antioxidant imbalance must be known in order to
deliver targeted exogenous antioxidant therapy (Kirkham & Rahman,
Genetic mutation to the gene encoding the cystic fibrosis transmem- 2006), suggesting that NAC therapy might not be appropriate in every
brane conductance regulator (CFTR), which functions as a cAMP chlo- case. However, in vitro studies of antioxidant use in asthma have, to-
ride channel in healthy lung cells, is responsible for cystic fibrosis date, not shown any definitive benefits; it has been postulated that this
(CF; Rowe & Clancy, 2006). This aberration leads to an alteration in lack of clinical efficacy is due to a lack of evidence base from which anti-
trans-epithelial ion transport, electrolyte balance and fluid content in oxidant type and dose were chosen from in vivo studies (Dworski, 2000).
a range of tissues. The effect is most evident in the lung, where viscous
mucus production is prolific, difficult to clear and susceptible to repeat- 4.2. Acetaminophen overdose and drug poisoning
ed chronic infection. Inflammation in response to iterative cycles of
infection manifests in lung damage and fibrosis, culminating in respira- Acetaminophen is a relatively safe and effective medication when
tory failure and mortality. used at appropriate doses. However, hepatocellular necrosis can result
The impact of NAC in the CF setting is likely to be multifactorial. First, it from either single or repeated doses of 10–15 g over 24 h in adult pa-
is believed to act as a mucolytic agent, delivering Cys residues to break tients of normal body weight in the absence of enzyme-inducing drugs
sulfhydryl bridges between glycoproteins in mucus (Sheffner et al., 1964). (Ryan, 2012). Acetaminophen is rapidly absorbed and undergoes first-
pass metabolism in the liver, one of the P450-derived by-products of
which is the potentially toxic metabolite, N-aceytl-p-benzoquinonimine
(NAPQI; Dart et al., 2006). NAPQI is normally detoxified via conjugation
to GSH in preparation for excretion in the urine, but overdose results in
hepatocellular damage on account of depletion of hepatic GSH, leading
Equation 3: Cys-mediated cleavage of disulfide bridges in proteins/ to protein binding and cellular dysfunction (Saito et al., 2010).
peptides/glycopeptides Administration of NAC in acetaminophen poisoning acts to rapidly
increase hepatic GSH synthesis and reduce protein binding (Saito
However, it is also likely that NAC might have several supplementary et al., 2010). The requirement for NAC is determined by plotting the
actions mediated by enhancement of intracellular GSH, the major anti- plasma level of acetaminophen against the time post-overdose on a
oxidant in lung tissue (Rahman & MacNee, 2000). It has been found that, nomogram; effectiveness of NAC therapy diminishes with time after
because exocytosis of GSH is thought to be dependent on the same acetaminophen overdose. The primary target is to initiate NAC treat-
channel that is dysregulated in CF (Dauletbaev et al., 2004), GSH in ment within 8 h to minimise the risk of hepatocellular damage (Gray
bronchoalveolar lavage (BAL) fluid has been found to be depressed et al., 2011), but reports indicate that administration of NAC after 10 h
(Gao et al., 1999) and after oral administration, free NAC could not be can still reduce mortality (Harrison et al., 1990). For practical reasons,
identified in BAL fluid (Cotgreave et al., 1987b). However, measures in in remote and rural areas, or when a blood sampling is unavailable,
sputum from CF patients suggest that GSH is elevated in this matrix the recommended treatment for acetaminophen overdose is methio-
and a sufficient proportion of it is in the reduced form to confer antiox- nine, although evidence of efficacy is lacking.
idant activity (Dauletbaev et al., 2004). The possibility also exists, Oral NAC (where available) has been found to be as effective as in-
however, that increased intracellular GSH on account of NAC treatment travenous NAC up to 10 h post-overdose. However, use of activated
might protect against the neutrophil-driven generation of ROS that charcoal in the accident & emergency department can decrease the
mediate the longer-term tissue damage and fibrosis in CF (Ratjen & absorption of orally administered NAC. Therefore, the preferred route
Grasemann, 2012). In addition, doses of N1.8 g/day have been shown of administration (indeed the only licenced route available in the UK)
to act to reduce the migration of neutrophils to the lungs, to restrict neu- is IV NAC (Nambiar, 2012). Despite arguments for and against the use
trophil access through narrowing of lung capillaries and to increase of IV over oral formulations of NAC, there have been no robust clinical
intra-neutrophil GSH (Tirouvanziam et al., 2006). Another small phase studies to test for the optimal route and duration of treatment (Chyka
II clinical study investigated the safety and efficacy of low-dose et al., 2000). In part, this is due to the requirement for a large sample
(700 mg/day) vs. high-dose (2800 mg/day) NAC over a 12 week peri- size to show a difference between these two routes of administration
od, where the high-dose regimen was found to be safe and efficacious (Waring, 2012).
(Dauletbaev et al., 2009). NAC has also been found to increase sputum The arrival of intravenous acetaminophen has also heralded a new set
penetration by DNA nanoparticle formulations (Suk et al., 2011). of concerns relating to potential overdose. Intravenous administration
The potential benefits of NAC have also been evaluated in a number of errors, together with a lack of understanding of dosage reductions re-
other respiratory diseases, including COPD, which may be characterised, quired for underweight patients is also thought to contribute to the po-
in part, by chronic mucus production, leading to an increased risk of tential risk of this formulation. Malnourished or perioperative patients
G.F. Rushworth, I.L. Megson / Pharmacology & Therapeutics 141 (2014) 150–159 155

who may have diminished stores of GSH have also been postulated to be tubular toxicity (Parfrey, 2005). Administration of contrast media alters
particularly susceptible to hepatocellular damage (Gray et al., 2011). the normal haemodynamic processes within the kidney. Initially, system-
ic vasodilation causes a decrease in renal blood flow which ultimately
4.3. Idiopathic pulmonary fibrosis results in vasoconstriction of the efferent arteriole to maintain renal out-
put and promote excretion of contrast. In addition to the haemodynamic
Idiopathic pulmonary fibrosis (IPF) or cryptogenic fibrosing alveoli- changes, toxic effects of contrast media in the renal tubules result in
tis is an interstitial lung disease of progressive nature. The median sur- inflammation and renal tubular necrosis. The degree of nephrotoxicity
vival from onset of dyspnoea is limited at 3–6 years (Nicholson et al., experienced after administration of contrast media varies widely from
2000). Originally, disease progression was thought to occur after an in- those who are asymptomatic to those who will require dialysis.
flammatory reaction, the termination of which results in fibrotic lesions. Ensuring adequate hydration with isotonic saline is currently the
However, this hypothesis has now been challenged and current dogma standard practice as preparation for patients about to undergo a radio-
points to cellular changes in response to chronic alveolar epithelial logical scan involving the administration of contrast media. A number
injury (Lota & Wells, 2013). of alternative pharmacological agents have been investigated for use
The presence of an imbalance between oxidants and antioxidants in in CIN. Ostensibly, agents have been trialled for this indication as a result
IPF was first reported in the 1980s. Lung epithelial cells obtained from of their vasodilatory characteristics and these include the use of calcium
BAL have been found to contain oxidants at cytotoxic levels (Cantin channel blockers, theophylline and dopamine (Jo, 2011) although none
et al., 1987) and GSH was found to be depressed within the epithelial have been shown to be efficacious.
lining fluid (Cantin et al., 1989). The results of two small studies in pa- NAC has been the subject of a series of trials in this arena, with mixed
tients with IPF found that administration of oral NAC (600 mg three results. There have been four recent trials published which demonstrate
times a day) increased the levels of GSH in BAL (Behr et al., 1997; significant reductions in CIN for those patients treated with intravenous
Meyer et al., 1994). The IFIGENIA trial reported improvements in the NAC versus control groups (Baker et al., 2003; Carbonell et al., 2010; Hsu
12 month vital capacity and carbon monoxide diffusing capacity in IPF et al., 2012; Koc et al., 2012). Another clinical study has reported a sig-
patients prescribed high-dose oral NAC (600 mg three times a day) in nificant NAC-induced decrease in serum creatinine, urinary markers of
addition to prednisolone and azathioprine (Demedts et al., 2005). How- oxidative stress and renal tubular injury (Drager et al., 2004). However,
ever, there have been some concerns over the design of the trial due to a number of studies have failed to show benefit of NAC in this setting: a
the relatively small numbers of patients it recruited (Fioret et al., 2011). study in 80 patients with chronic renal insufficiency (Goldenberg et al.,
It is also worth noting that there is little published evidence for the use 2004) failed to show a difference between oral NAC (600 mg 3 times a
of prednisolone and azathioprine in this indication (Rudd et al., 2007; day) compared to placebo for the primary end point of acute contrast-
Maher, 2011) and it has been suggested the benefits of NAC may be as induced increase in serum creatinine. A similar lack of efficacy for NAC
a result of the protective benefits of NAC with regard to azathioprine was noted in a study of 10,574 patients undergoing percutaneous
toxicity. Interestingly, when the effect of intravenous NAC administra- coronary intervention (Gurm et al., 2012). In a recent study comparing
tion at a range of doses was compared between patients with known intravenous and intrarenal administration of NAC/placebo as CIN pro-
IPF and healthy volunteers, there was no difference in the GSH levels phylaxis in patients with ST-elevation myocardial infarction undergoing
noted in the healthy volunteer group (Meyer et al., 1995), reinforcing primary angiography, there was no effect of NAC in the prevention of
the concept that NAC only elicits an effect to replenish GSH levels in CIN (Aslanger et al., 2012). It should be noted that patients were select-
tissue that is deficient in the tripeptide. ed for this study based on the index event rather than pre-existing renal
A more recent study, PANTHER-IPF, was set up to investigate the insufficiency, diabetes or chronic cardiac condition. Inappropriate dos-
safety and efficacy of the triple therapy combination of prednisolone, age and/or route of NAC administration have been cited as potential rea-
azathioprine and NAC. The study was designed with three arms: triple sons for the lack of efficacy of NAC in some studies (Shalansky et al.,
therapy, NAC-alone and a placebo group. An interim analysis was 2005), but little attention appears to have been paid to the fundamental
published in 2012 which reported an increased mortality rate (8 vs 1 premise for NAC antioxidant activity – is GSH depressed in renal tissue
death, p = 0.01) and hospitalisation (23 vs 7, p b 0.001) in those of patients? This aspect, above all, might help to explain the inconsistent
patients on the triple therapy arm: this arm was subsequently stopped results because the GSH status in renal patients is likely to be closely re-
early due to safety concerns and the authors state that the combination lated to each individual's specific disease aetiology. The relative success
should not be used in these patients (Raghu et al., 2012). of any given clinical study is likely to hinge on the proportion of patients
with depleted GSH, but with the necessary enzyme complement to
4.4. Contrast-induced nephropathy (CIN) actively convert NAC to GSH.
A recent meta-analysis on the use of intravenous NAC in CIN was in-
CIN is a serious adverse condition that develops in response to con- conclusive; the results showed a tendency towards a beneficial effect of
trast media administered to patients during radiological procedures, NAC administration, but were not significant (p = 0.06; Sun et al.,
such as computed tomography or angiography. The clinical definition 2013). The authors commented on the heterogeneity of the trials within
of CIN is a ≥25% increase in serum creatinine compared to baseline the meta-analysis and have called for a large-scale, well designed study
(Parfrey, 2005). Within secondary care settings, an increase in morbid- to investigate the potential for IV NAC in CIN.
ity and mortality has been realised as a result. The risk of developing CIN Overall, due to its low acquisition cost and relative ease of adminis-
increase in the presence of pre-existing risk factors for the development tration and despite the inconclusive evidence in favour of its use, intra-
of renal impairment, for example in those with cardiovascular or renal venous NAC as a prophylactic agent for prevention of CIN is advocated
co-morbidities (particularly diabetic nephropathy) or in the elderly by a number of centres. There is a requirement for further research in
(Jo, 2011); the risk to patients with normal renal function and without this area to determine if there is any long-term benefit of the prophylac-
risk factors for developing CIN is small (Parfrey, 2005). However, a tic treatment. An ongoing study to investigate the effect of NAC in
recently published retrospective study investigating the incidence of healthy volunteers and patients with renal failure exposed to contrast
acute kidney injury after radiological procedures either with or without medium might shed further light on the issue (Sandilands et al., 2012).
contrast was unable to find a causal link between intravenous contact
and acute kidney injury, even in groups previously considered to be at 4.5. Cardiovascular disease and diabetes
risk (McDonald et al., 2013).
The aetiology of CIN is not fully understood, but it is thought to relate Oxidative stress is a key component in the atherogenic process,
to two separate mechanisms – haemodynamic changes and renal resulting in the oxidation of lipids in low density lipoproteins (LDL)
156 G.F. Rushworth, I.L. Megson / Pharmacology & Therapeutics 141 (2014) 150–159

and rendering them recognisable to macrophages prior to ingestion decrease in the primary end points which were survival, event-free sur-
and formation of resident foam cells in the vessel wall (Griendling & vival or secondary primary tumours.
Alexander, 1997). Research has shown that that there is a depletion NAC has also been investigated as a potential agent to attenuate side
of GSH in blood vessels from atherosclerotic mice from a very early effects of platinum-based chemotherapy. One study in rats suggests that
age (Biswas et al., 2005) and an imbalance, and in some cases a com- NAC blocks oxidative stress to reduce cisplatin-induced acute renal im-
plete absence, of GPx activity, indicating the lack of an antioxidant reg- pairment (Luo et al., 2008). A recent small clinical study in 84 patients
ulator of oxidative stress in cells associated with human atherosclerotic has reported that transtympanic injections of NAC appear to prevent
plaque (Lapenna et al., 1998). Furthermore, macrophage GSH and GPx cisplatin-induced ototoxicity (Riga et al., 2013).
concentrations are inversely related to oxidation of LDL cholesterol,
whereby low concentrations of the antioxidants lead to an increase in 4.7. Human immunodeficiency virus
lipoprotein oxidation (Rosenblat & Aviram, 1998). As well as GSH deple-
tion associated with atherosclerotic plaques, there is also evidence that Two key papers reported in the late 1990s that low GSH
GSH is depressed in platelets in conditions associated with increased (Herzenberg et al., 1997) or thiol levels (Marmor et al., 1997) may be
risk of thrombosis. Principal amongst these is diabetes (both type 1 used as a predictor or decreased survival in patients with human immu-
and type 2), where hypersensitivity of platelets to thrombotic stimuli nodeficiency virus (HIV) infection. Subsequently, an 8 week double-
has been linked to low intraplatelet GSH levels (Mazzanti & Mutus, blind, placebo-controlled clinical trial identified that oral NAC could
1997). The potential for platelet GSH as a therapeutic target is supported safely replenish whole blood GSH and T-cell GSH in patients with HIV
by animal experiments in which high dose NAC was found to improve infection (De Rosa et al., 2000). This study was not designed to investi-
endothelial function (Pieper & Siebeneich, 1998). gate any potential clinical benefits in survival, but the emergence of de-
A number of clinical studies to investigate the potential for NAC as a pleted GSH as a feature in HIV infection would certainly merit further
therapeutic agent have been conducted in a type 2 diabetes population. exploration of the potential benefits of NAC in this setting.
This is a particularly important target group on account of the increased
risk of cardiovascular disease in diabetes, coupled with the lack of effi- 4.8. Neuropsychiatric disorders
cacy of aspirin in primary prevention of cardiovascular complications
amongst these patients. First, it has been shown that 6 month treat- GSH depletion is a feature of a wide range of neuropsychiatric disor-
ment with a combination of L-arginine (to prime synthesis of the ders, including Alzheimer's, Parkinson's and Huntingdon's diseases
endothelium-derived vasodilator, nitric oxide; NO) with NAC reduced (Johnson et al., 2012). Although there are subtle differences in GSH
blood pressure by ~5 mmHg in patients with type 2 diabetes and hyper- handling within the central nervous system compared to other tissues,
tension (Martina et al., 2008). Meanwhile, several other studies focused the basic concepts are the same: GSH is predominantly synthesised in
on the impact of NAC in platelet function: Gibson et al. demonstrated the cytoplasm of cells and is dependent on influx of Cys to drive the
that NAC increased intraplatelet levels of GSH, decreased ROS detection rate-limiting step of GSH synthesis. The principal player in transport
and reduced platelet activation in vitro (Gibson et al., 2011). Leading on of Cys into neurones is the excitatory amino acid transporter C1
from this observation, the antiplatelet properties of NAC were trialled (ECAAC1); astrocytes also employ cystine/glutamate (Xc−) antiporters
in vivo in a cohort of type 2 diabetes patients (Treweeke et al., 2012). to supplement the intracellular thiol pool. GSH concentrations in cere-
The study reported an inhibition of platelet–monocyte conjugation, a brospinal fluid are similar to those in blood (~5 μM) and evidence for
surrogate marker of cardiovascular risk, within 2 h of administration, transit of GSH across the blood–brain-barrier seems unlikely.
an effect that was maintained following daily self-administration over Given that Cys is central to neuronal GSH synthesis, NAC has been
a 1 week period. Crucially, however, the use of platelets in this study trialled in a number of neuropsychiatric disorders where redox imbal-
facilitated measurement of GSH in the target cells and found that NAC ance has been implicated in the aetiology. In some instances, there has
was most effective at reducing platelet–monocyte conjugation in been a suggestion that NAC might be useful in this setting, not only on
those patients with the most depleted platelet GSH; patients with plate- account of repletion of GSH, but also because NAC-derived cystine has
lets replete in GSH at baseline received no benefit from NAC treatment the potential to lead to an increase in release of glutamate from astro-
and did not experience a rise in platelet GSH. This result reinforces cytes via the Xc− antiporters, resulting in activation of neuronal gluta-
the notion that stratification of patients prior to treatment with NAC mate receptors, resulting in dopamine release (Dean et al., 2011).
is essential to ensure successful treatment: the evidence from a wide A number of small clinical studies and case reports have cited the
range of sources strongly suggests that NAC will be ineffectual in utility of NAC in a variety of neuropsychiatric conditions from addiction
those patients in whom GSH is not depleted in the target tissue. (Mardikian et al., 2007; Knackstedt et al., 2009; Gray et al., 2010) to
schizophrenia (Lavoie et al., 2008; Berk et al., 2008b), obsessive compul-
4.6. Carcinogenesis and cancer sive disorder (Lafleur et al., 2006), bipolar disorder (Berk et al., 2008a,
2011) and even as a neuroprotective agent in Alzheimer's disease
The anti-carcinogenic properties of NAC have been reported since (Adair et al., 2001). Redox imbalance and, more specifically, intracellu-
the early1980s (De Flora et al., 1984, 1985). These properties are pleio- lar GSH depletion has been implicated in autism, prompting a more
tropic, ranging from benefits of NAC in preventative, pre-neoplastic and recent randomized, placebo-controlled trial to determine the impact
cancer treatment stages. It is thought that NAC elicits its anticarcinogen- of oral NAC (900 mg self-administered once daily for 4 weeks, then
ic actions via a number of different physiological processes including the twice-daily for 4 weeks and three times daily for 4 weeks; 31 patients,
attenuation of genotoxic ROS, induction of DNA repair, and regulation of 14 randomized to NAC available for follow-up) on children with autism
carcinogen-induced apoptosis (De Flora et al., 2001). (Hardan et al., 2012). The trial found a substantial and significant
Clinical studies investigating NAC as a potential chemopreventive improvement in irritability, measured as one of several parameters
agent were first conducted in the 1990s. A phase I study reported that on the Aberrant Behaviour Checklist (ABC). NAC was not, however,
significant toxic effects were noted in subjects exposed to doses of NAC shown to increase survival in patients with amyotrophic lateral sclerosis
(after titration phase, total daily oral dose was 6400 mg/m2) which (Louwerse et al., 1995).
were deemed to be biochemically effective (Pendyala & Creaven, Clinical data are therefore beginning to accumulate to support posi-
1995). Since the publication of this study, a phase II trial, EUROSCAN, tive findings in animal studies and to suggest therapeutic potential for
conducted in patients with head and neck or lung cancer has also report- NAC in a wide range of neurological disorders in which oxidative stress
ed (van Zandwijk et al., 2000). This study used a much smaller dose of is implicated. While the studies to date are, on the whole, encouraging,
600 mg/day oral NAC for 2 years. The study failed to show any significant they are typically small and require confirmation in larger studies. Of
G.F. Rushworth, I.L. Megson / Pharmacology & Therapeutics 141 (2014) 150–159 157

particular interest will be clarification of whether NAC itself is able to References


cross the blood brain barrier or whether benefits are realised through
simply increasing the Cys pool in the blood, leading to the necessary Adair, J. C., Knoefel, J. E., & Morgan, N. (2001). Controlled trial of N-acetylcysteine for
patients with probable Alzheimer's disease. Neurology 57, 1515–1517.
increase cysteine in cerebrospinal fluid to fuel intra-neuronal GSH Aitio, M. (2006). N-acetylcysteine — passe-partout or much ado about nothing? Br J Clin
synthesis. Pharmacol 61, 5–15.
Arrigo, A. (1999). Gene expression and the thiol redox state. Free Radic Biol Med 27,
936–944.
Aruoma, O. I., Halliwell, B., Hoey, B.M., & Butler, J. (1989). The antioxidant action of
5. Discussion and conclusions N-acetylcysteine: its reaction with hydrogen peroxide, hydroxyl radical, superoxide,
and hypochlorous acid. Free Radic Biol Med 6, 593–597.
Aslanger, E., Uslu, B., Akdeniz, C., Polat, N., Cizgici, Y., & Oflaz, H. (2012). Intrarenal appli-
It is vital to think of NAC as a pro-drug, the actions of which are al- cation of N-acetylcysteine for the prevention of contrast medium-induced nephrop-
most exclusively driven by, and dependent on, successful conversion athy in primary angioplasty. Coron Artery Dis 23, 265–270.
to the powerful detoxifying agent and antioxidant, GSH. In those thera- Baker, C. S. R., Wragg, A., Kumar, S., De Palma, R., Baker, L. R. I., & Knight, C. J. (2003). A
rapid protocol for the prevention of contrast-induced renal dysfunction: the
peutic targets where an antioxidant activity is the principal mode of
RAPPID study. J Am Coll Cardiol 41, 2114–2118.
activity, it is apparent that intracellular incorporation into GSH is vital Banner, W. J., Koch, M., Capin, D.M., Hopf, S. B., Chang, S., & Tong, T. G. (1986). Experimental
for efficacy. The success of NAC in acetaminophen overdose is testament chelation therapy in chromium, lead, and boron intoxication with N-acetylcysteine
and other compounds. Toxicol Appl Pharmacol 83, 142–147.
to this concept: there is an absolute requirement for de novo synthesis
Bartoli, G. M., & Sies, H. (1978). Reduced and oxidized glutathione efflux from liver. FEBS
of GSH to replace that used to detoxify the NAPQI metabolite in these Lett 86, 89–91.
patients. In this setting, there is no question that NAC is a life-saving Becker, A., & Soliman, K. F. (2009). The role of intracellular glutathione in inorganic
drug, with the only apparent limitation in its efficacy related to the mercury-induced toxicity in neuroblastoma cells. Neurochem Res 34, 1677–1684.
Behr, J., Maier, K., Degenkolb, B., Krombach, F., & Volgelmeier, C. (1997). Antioxidative and
time delay between overdose and therapy. clinical effects of high-dose N-acetylcysteine in fibrosing alveolitis: adjunctive thera-
NAC is often referred to as a “powerful antioxidant”, a misconception py to maintenance immunosuppression. Am J Respir Crit Care Med 156, 1897–1901.
that has potentially fuelled many of the studies associated with indica- Berk, M., Copolov, D. L., Dean, O., Lu, K., Jeavons, S., Schapkaitz, I., et al. (2008a). N-acetyl
cysteine for depressive symptoms in bipolar disorder—a double-blind randomized
tions other than acetaminophen overdose. NAC in itself has some placebo-controlled trial. Biol Psychiatry 64, 468–475.
antioxidant potential, but it is a much weaker antioxidant than GSH Berk, M., Copolov, D., Dean, O., Lu, K., Jeavons, S., Schapkaitz, I., et al. (2008b). N-acetyl
and most, if not all, of the other endogenous antioxidant agents and cysteine as a glutathione precursor for schizophrenia—a double-blind, randomized,
placebo-controlled trial. Biol Psychiatry 64, 361–368.
enzymes. This knowledge, coupled with the relatively low plasma con- Berk, M., Dean, O., Cotton, S. M., Gama, C. S., Kapczinski, F., Fernandes, B.S., et al. (2011).
centrations that are achievable with intravenous (100–500 μM) and The efficacy of N-acetylcysteine as an adjunctive treatment in bipolar depression:
certainly oral (b5 μM) dosing points to the fact that NAC itself is unlikely an open label trial. J Affect Disord 135, 389–394.
Biswas, S. K., Newby, D. E., Rahman, I., & Megson, I. L. (2005). Depressed glutathione syn-
to contribute significantly to antioxidant defence. Similarly, Cys gener-
thesis precedes oxidative stress and atherogenesis in Apo-E-/- mice. Biochem Biophys
ated through hydrolysis of NAC is rapidly oxidised to cystine with the Res Commun 338, 1368–1373.
loss of any antioxidant potential. Instead, the antioxidant potential of Blanusa, M., Varnai, V. M., Piasek, M., & Kostial, K. (2005). Chelators as antidotes of metal
toxicity: therapeutic and experimental aspects. Curr Med Chem 12, 2771–2794.
NAC is in the provision of substrate for synthesis of intracellular GSH,
Borgstrom, L., Kagedal, B., & Paulsen, O. (1986). Pharmacokinetics of N-acetylcysteine in
present in healthy cells at millimolar concentrations. Rapid conversion man. Eur J Clin Pharmacol 31, 217–222.
of NAC in GSH-deficient cells maintains a steep concentration gradient Brandao, R., Santos, F. W., Zeni, G., Rocha, J. B., & Nogueira, C. W. (2006). DMPS and
across the membrane, and drives sufficient accumulation of GSH to re- N-acetylcysteine induce renal toxicity in mice exposed to mercury. Biometals 19,
389–398.
store the concentration to healthy levels. Crucially, there is no further Buyukhatipoglu, H., Sezen, Y., Yildiz, A., Bas, M., Kirhan, I., Ulas, T., et al. (2010). N-
accumulation of GSH in those cells already replete in GSH, rendering acetylcysteine fails to prevent renal dysfunction and oxidative stress after noniodine
NAC ineffectual once GSH has been replenished. contrast media administration during percutaneous coronary interventions. Pol Arch
Med Wewn 120, 383–389.
In light of these issues, NAC should not be considered a powerful an- Cantin, A.M., Hubbard, R. C., & Crystal, R. G. (1989). Glutathione deficiency in the epithe-
tioxidant, nor should it be seen as a panacea in all conditions driven by lial lining fluid of the lower respiratory tract in idiopathic pulmonary fibrosis. Am Rev
oxidative stress. However, there is a significant opportunity for a strati- Respir Dis 139, 370–372.
Cantin, A.M., North, S. L., Fells, G. A., Hubbard, R. C., & Crystal, R. G. (1987). Oxidant-mediated
fied approach to NAC therapy in a wide range of conditions, using intra- epithelial cell injury in idiopathic pulmonary fibrosis. J Clin Invest 79, 1665–1673.
cellular GSH as the predictor of efficacy. This approach is likely to be Carbonell, N., Sanjuan, R., Blasco, M., Jorda, A., & Miguel, A. (2010). N-acetylcysteine:
most successful in conditions where target cells are easily accessible short-term clinical benefits after coronary angiography in high-risk renal patients.
Rev Esp Cardiol 63, 12–19.
for GSH measurement (e.g. platelets for anti-thrombotic therapy,
Chyka, P. A., Butler, A. Y., Holliman, B. J., & Herman, M. I. (2000). Utility of acetylcysteine in
T-cells for HIV treatment, monocytes and neutrophils for certain anti- treating poisonings and adverse drug reactions. Drug Saf 22, 123–148.
inflammatory therapy), but there is the potential for GSH in these cells Cotgreave, I. A., Berggren, M., Jones, T. W., Dawson, J., & Moldeus, P. (1987a). Gastrointes-
to also act as surrogate markers for systemic GSH depletion that might tinal metabolism of N-acetylcysteine in the rat, including an assay for sulfite in bio-
logical systems. Biopharm Drug Dispos 8, 337–386.
be relevant in some other conditions. Overall, researchers should Cotgreave, I. A., Eklund, A., Larsson, K., & Moldeus, P. W. (1987b). No penetration of orally
not be discouraged by some of the recent negative findings for NAC in administered N-aceylcysteine into bronchoalveolar lavage fluid. Eur J Respir Dis 70,
novel indications. Instead, we should take heart from the monumental 73–77.
Cotgreave, I., Moldéus, P., & Schuppe, I. (1991). The metabolism of N-acetylcysteine by
success of NAC in acetaminophen overdose therapy and seek similar op- human endothelial cells. Biochem Pharmacol 42, 13–16.
portunities where GSH depletion is a key component in the disease Dart, R. C., Erdman, A.R., Olson, K. R., Christianson, G., Manoguerra, A. S., Chyka, P. A., et al.
aetiology. Only in these situations is NAC therapy likely to represent a (2006). Acetaminophen poisoning: an evidence-based consensus guideline for
out-of-hospital management. Clin Toxicol 44, 1–18.
credible therapeutic solution. Dauletbaev, N., Fisher, P., Aulbach, B., Gross, J., Kusche, W., Thyroff-Friesinger, U., et al.
(2009). A phase II study on safety and efficacy of high-dose N-acetylcysteine in
patients with cystic fibrosis. Eur J Med Res 14, 352–358.
Dauletbaev, N., Viel, K., Buhl, R., Wagner, T. O. F., & Bargon, J. (2004). Glutathione and glu-
Conflict of interest statement tathione peroxidase in sputum samples of adult patients with cystic fibrosis. J Cyst
Fibros 3, 119–124.
The authors declare that there are no conflicts of interest. De Flora, S., Benicelli, C., Camoirano, A., Serra, D., Romano, M., Rossi, G. A., et al. (1985). In
vivo effects of N-acetylcysteine on glutathione metabolism and on the biotransforma-
tion of carcinogenic and/or mutagenic compounds. Carcinogenesis 6, 1735–1745.
De Flora, S., Bennicelli, C., Zanacchi, P., Camoirano, A., Morelli, A., & De Flora, A. (1984). In
Author declaration vitro effects of N-acetylcysteine on the mutagenicity of direct-acting compounds and
procarcinogens. Carcinogenesis 5, 505–510.
De Flora, S., Izzotti, A., D'Agostini, F., & Balansky, R. M. (2001). Mechanisms of N-
The authors declare that this manuscript has not been published or acetylcysteine in the prevention of DNA damage and cancer, with special reference
submitted for publication to any other journals. to smoking-related end-points. Carcinogenesis 22, 999–1013.
158 G.F. Rushworth, I.L. Megson / Pharmacology & Therapeutics 141 (2014) 150–159

De Rosa, S.C., Zaretsky, M.D., Dubs, J. G., Roederer, M., Anderson, M., Green, A., et al. Kaur, P., Aschner, M., & Syversen, T. (2006). Glutathione modulation influences methyl
(2000). N-acetylcysteine replenishes glutathione in HIV infection. Eur J Clin Invest mercury induced neurotoxicity in primary cell cultures of neurones and astrocytes.
30, 915–929. Neurotoxicology 27, 492–500.
Dean, O., Giorlando, F., & Berk, M. (2011). N-acetylcysteine in psychiatry: current thera- Kirkham, P., & Rahman, I. (2006). Oxidative stress in asthma and COPD: antioxidants as a
peutic evidence and potential mechanisms of action. J Psychiatry Neurosci 36, 78–86. therapeutic strategy. Pharmacol Ther 111, 476–494.
Demedts, M., Behr, J., Buhl, R., Costabel, U., Dekhuijzen, R., & Jansen, H. M. (2005). High- Knackstedt, L. A., LaRowe, S., Mardikian, P., Malcolm, R., Upadhyaya, H., Hedden, S., et al.
dose acetylcysteine in idiopathic pulmonary fibrosis. N Engl J Med 353, 2229–2242. (2009). The role of cystine-glutamate exchange in nicotine dependence in rats and
Dickinson, D. A., Levonen, A., Moellering, D. R., Arnold, E. K., Zhang, H., Darley-Usmar, V. humans. Biol Psychiatry 65, 841–845.
M., et al. (2004). Human glutamate cysteine ligase gene regulation through the elec- Koc, F., Ozdemir, K., Kaya, M. G., Dogdu, O., Vatankulu, M.A., Ayhan, S., et al. (2012). Intra-
trophile response element. Free Radical Biol Med 37, 1152–1159. venous N-acetylcysteine plus high-dose hydration versus high-dose hydration and
Drager, L. F., Andrade, L., de Toledo, J. F. B., Laurindo, F. R. M., Cesar, L. A.M., & Seguro, A.C. standard hydration for the prevention of contrast-induced nephropathy: CASIS—A
(2004). Renal effects of N-acetylcysteine in patients at risk for contrast nephropathy: multicenter prospective controlled trial. Int J Cardiol 155, 418–423.
decrease in oxidant stress-mediated renal tubular injury. Nephrol Dial Transplant 19, Lafleur, D. L., Pittenger, C., Kelmendi, B., Gardner, T., Wasylink, S., Malison, R. T., et al.
1803–1807. (2006). N-acetylcysteine augmentation in serotonin reuptake inhibitor refractory ob-
Dworski, R. (2000). Oxidant stress in asthma. Thorax 55, S51–S53. sessive–compulsive disorder. Psychopharmacology 184, 254–256.
Fioret, D., Perez, R. L., & Roman, J. (2011). Management of idiopathic pulmonary fibrosis. Lapenna, D., de Gioia, S., Ciofani, G., Mezzetti, A., Ucchino, S., Calafiore, A.M., et al. (1998).
Am J Med Sci 341, 450–453. Glutathione-related antioxidant defenses in human atherosclerotic plaques. Circulation
Flanagan, R. J., & Meredith, T. J. (1991). Use of N-acetylcysteine in clinical toxicology. Am J 97, 1930–1934.
Med 91, S131–S139. Lavoie, S., Murray, M. M., Deppen, P., Knyazeva, M. G., Berk, M., Boulat, O., et al. (2008).
Gao, L., Kim, K. J., Yankaskas, J. R., & Forman, H. J. (1999). Abnormal glutathione trans- Glutathione precursor, N-acetyl-cysteine, improves mismatch negativity in schizo-
port in cystic fibrosis airway epithelia. Am J Physiol Lung Cell Mol Physiol 277, phrenia patients. Neuropsychopharmacology 33, 2187–2199.
L113–L118. Le Brocq, M., Leslie, S. J., Milliken, P., & Megson, I. L. (2008). Endothelial dysfunction: From
Gibson, K. R., Neilson, I. L., Barrett, F., Winterburn, T. J., Sharma, S., Macrury, S. M., et al. molecular mechanisms to measurement, clinical implications, and therapeutic op-
(2009). Evaluation of the antioxidant properties of N-acetylcysteine in human portunities. Antioxid Redox Signal 10, 1631–1674.
platelets: prerequisite for bioconversion to glutathione for antioxidant and Leelarungrayub, D., Khansuwan, R., Pothongsunun, P., & Klaphajone, J. (2011). N-
antiplatelet activity. J Cardiovasc Pharmacol 54, 319–326. acetylcysteine supplementation controls total antioxidant capacity, creatinine kinase,
Gibson, K. R., Winterburn, T. J., Barrett, F., Sharma, S., MacRury, S., & Megson, I. L. (2011). lactate, and tumour necrotic factor-alpha against oxidative stress induced by graded
Therapeutic potential of N-acetylcysteine as an antiplatelet agent in patients with exercise in sedentary men. Oxid Med Cell Longev 2011, 329643.
type-2 diabetes. Cardiovasc Diabetol 10, 1–8. Li, X., Qu, Z. C., & May, J. M. (2001). GSH is required to recycle ascorbic acid in cultured
Giustarini, D., Milzani, A., Dalle-Donne, I., Tsikas, D., & Rossi, R. (2012). N-acetylcysteine liver cell lines. Antioxid Redox Signal 3, 1089–1097.
derivative with an unusual pharmacokinetic feature and remarkable antioxidant Lodge, J. K., Traber, M. G., & Packer, L. (1998). Thiol chelation of Cu2+ by dihydrolipoic acid
potential. Biochem Pharmacol 84, 1522–1533. prevents human low density lipoprotein peroxidation. Free Radic Biol Med 25, 287–297.
Goldenberg, I., Shechter, M., Matetzky, S., Jonas, M., Adam, M., Pres, H., et al. (2004). Oral Lota, H. K., & Wells, A. U. (2013). The evolving pharmacotherapy of pulmonary fibrosis.
acetylcysteine as an adjunct to saline hydration for the prevention of contrast- Expert Opin Pharmacother 14, 79–89.
induced nephropathy following coronary angiography: a randomized controlled Louwerse, E. S., Weverling, G. J., Bossuyt, P.M., Meyje, F. E., & de Jong, J. M. (1995).
trial and review of the current literature. Eur Heart J 25, 212–218. Randomized double-blind, controlled trial of acetylcysteine in amyotrophic lateral
Gosselin, S., Hoffman, R. S., Juurlink, D. N., Whyte, I., Yarema, M., & Caro, J. (2013). Treating sclerosis. Arch Neurol 52, 559–564.
acetaminophen overdose: thresholds, costs and uncertainties. Clin Toxicol 51, 252–378. Lu, S.C. (2009). Regulation of glutathione synthesis. Mol Aspects Med 30, 42–59.
Gray, T., Hoffman, R. S., & Bateman, D. N. (2011). Intravenous paracetamol—an interna- Lund, M. E., Banner, W. J., Clarkson, T. W., & Berlin, M. (1984). Treatment of acute meth-
tional perspective of toxicity. Clin Toxicol 49, 150–152. ylmercury ingestion by hemodialysis with N-acetylcysteine (Mucomyst) infusion and
Gray, K. M., Watson, N. L., Carpenter, M. J., & LaRowe, S. D. (2010). N-acetylcysteine (NAC) 2,3-dimercaptopropane sulfonate. J Toxicol Clin Toxicol 22, 31–49.
in young marijuana users: An open-label pilot study. Am J Addict 19, 187–189. Luo, J., Tsuji, T., Yasuda, H., Sun, Y., Fujigaki, Y., & Hishida, A. (2008). The molecular mech-
Green, J. L., Heard, K. J., Reynolds, K. M., & Albert, D. (2013). Oral and intravenous anisms of the attenuation of cisplatin-induced acute renal failure by N-acetylcysteine
acetylcysteine for treatment of acetaminophen toxicity: a systematic review and in rats. Nephrol Dial Transplant 23, 2198–2205.
meta-analysis. West J Emerg Med 14, 218–226. Maher, T. M. (2011). Current and future therapies for idiopathic pulmonary fibrosis. Clin
Griendling, K. K., & Alexander, R. W. (1997). Oxidative stress and cardiovascular disease. Pulm Med 18, 257–264.
Circulation 96, 3264–3275. Mardikian, P. N., LaRowe, S. D., Hedden, S., Kalivas, P. W., & Malcolm, R. J. (2007). An open-
Griffith, O. W. (1999). Biologic and pharmacologic regulation of mammalian glutathione label trial of N-acetylcysteine for the treatment of cocaine dependence: a pilot study.
synthesis. Free Radic Biol Med 27, 922–935. Prog Neuropsychopharmacol Biol Psychiatry 31, 389–394.
Gurm, H. S., Smith, D. E., Berwanger, O., Share, D., Schreiber, T., Moscucci, M., et al. (2012). Marmor, M., Alcabes, P., Titus, S., Frenkel, K., Krasinski, K., Penn, A., et al. (1997). Low serum
Contemporary use and effectiveness of N-acetylcysteine in preventing contrast- thiol levels predict shorter times-to-death among HIV-infected injecting drug users.
induced nephropathy among patients undergoing percutaneous coronary interven- AIDS 11, 1389–1393.
tion. JACC Cardiovasc Interv 5, 98–104. Martina, V., Masha, A., Gigliardi, V. R., Brocato, L., Manzato, E., Berchio, A., et al. (2008).
Halliwell, B. (1996). Vitamin C: Antioxidant or pro-oxidant in vivo. Free Radic Res 25, Long-term N-acetylcysteine and L-arginine administration reduces endothelial acti-
439–454. vation and systolic blood pressure in hypertensive patients with type 2 diabetes.
Halpner, A.D., Handelman, G. J., Harris, J. M., Belmont, C. A., & Blumberg, J. B. (1998). Pro- Diabetes Care 31, 940–944.
tection by vitamin C of loss of vitamin E in cultured rat hepatocytes. Arch Biochem May, J. M., Qu, Z. C., & Mendiratta, S. (1998). Protection and recycling of alpha-tocopherol
Biophys 359, 305–309. in human erythrocytes by intracellular ascorbic acid. Arch Biochem Biophys 349,
Hardan, A. Y., Fung, L. K., Libove, R. A., Obukhanych, T. V., Nair, S., Herzenberg, L. A., et al. 281–289.
(2012). A randomised controlled pilot trial of oral N-acetylcysteine in children with Mazzanti, L., & Mutus, B. (1997). Diabetes-induced alterations in platelet metabolism.
autism. Biol Psychiatry 71, 956–961. Clin Biochem 30, 509–515.
Harrison, P.M., Keays, R., Bray, G. P., Alexander, G. J. M., & Williams, R. (1990). Improved McDonald, R. J., McDonald, J. S., Bida, J. P., Carter, R. E., Fleming, C. J., Misra, S., et al. (2013).
outcome of paracetamol-induced fulminant hepatic failure by late administration of Intravenous contrast material-induced nephropathy: causal or coincident phenome-
acetylcysteine. Lancet 335, 1572–1573. non? Radiology 267, 106–118.
Hayakawa, M., Miyashita, H., Sakamoto, I., Kitagawa, M., Tanaka, H., Y., Hideyo, et al. McPherson, R. A., & Hardy, G. (2012). Cysteine: the Fun-Ke nutraceutical. Nutrition 28,
(2003). Evidence that reactive oxygen species do not mediate NF-kB activation. 336–337.
EMBO J 22, 3356–3366. Meister, A., & Anderson, M. E. (1983). Glutathione. Annu Rev Biochem 52, 711–760.
Herzenberg, L. A., De Rosa, S.C., Dubs, J. G., Roederer, M., Anderson, M. T., Ela, S. W., et al. Meyer, A., Buhl, R., Kampf, S., & Magnussen, H. (1995). Intravenous N-acetylcysteine and
(1997). Glutathione deficiency is associated with impaired survival in HIV disease. lung glutathione of patients with pulmonary fibrosis and normals. Am J Respir Crit
Proc Natl Acad Sci U S A 94, 1967–1972. Care Med 152, 1055–1060.
Holdiness, M. R. (1991). Clinical pharmacokinetics of N-acetylcysteine. Clin Pharmacokinet Meyer, A., Buhl, R., & Magnussen, H. (1994). The effect of oral N-acetylcysteine on lung
20, 123–134. glutathione levels in idiopathic pulmonary fibrosis. Eur Respir J 7, 431–436.
Hsu, T., Huang, M., Yu, S., Yen, D. H., Kao, W., Chen, Y., et al. (2012). N-acetylcysteine for Montecinos, V., Guzmán, P., Barra, V., Villagrán, M., Muñoz-Montesino, C., Sotomayor, K.,
the prevention of contrast-induced nephropathy in the emergency department. et al. (2007). Vitamin C is an essential antioxidant that enhances survival of oxida-
Intern Med 51, 2709–2714. tively stressed human vascular endothelial cells in the presence of a vast molar excess
Hurst, G. A., Shaw, P. B., & LeMaistre, C. A. (1967). Laboratory and clinical evaluation of the of glutathione. J Biol Chem 282, 15506–15515.
mucolytic properties of acetylcysteine. Am Rev Respir Dis 96, 962–970. Multicentre Study Group (1980). Long-term oral acetylcysteine in chronic bronchitis. A
Jo, S. (2011). N-acetylcysteine for prevention of contrast-induced nephropathy: a narra- double-blind controlled study. Eur J Respir Dis 61, 93–108.
tive review. Korean Circ J 41, 695–702. Nambiar, N. J. (2012). Management of paracetamol poisoning: The old and the new. J Clin
Johnson, W. M., Wilson-Delfosse, A. L., & Mieyal, J. J. (2012). Dysregulation of glutathione Diagn Res 6, 1101–1104.
homeostasis in neurodegenerative diseases. Nutrients 4, 1399–1440. Nicholson, A. G., Colby, T. V., du Bois, R. M., Hansell, D.M., & Wells, A. U. (2000). The
Joshi, D., Mittal, D. K., Shrivastava, S., & Shukla, S. (2010). Protective role of thiol chelators prognostic significance of the histologic pattern of interstitial pneumonia in patients
against dimethylmercury induced toxicity in male rats. Bull Environ Contam Toxicol presenting with the clinical entity of cryptogenic fibrosing alveolitis. Am J Respir Crit
84, 613–617. Care Med 162, 2213–2217.
Kanter, M. Z. (2006). Comparison of oral and iv acetylcysteine in the treatment of acet- Olsson, B., Johansson, M., Gabrielsson, J., & Bolme, P. (1988). Pharmacokinetics and bio-
aminophen poisoning. Am J Health Syst Pharm 63, 1821–1827. availability of reduced and oxidized N-acetylcysteine. Eur J Clin Pharmacol 34, 77–82.
G.F. Rushworth, I.L. Megson / Pharmacology & Therapeutics 141 (2014) 150–159 159

Pakravan, N., Waring, W. S., Sharma, S., Ludlam, C., Megson, I. L., & Bateman, N. (2008). Shalansky, S. J., Pate, G. E., Levin, A., & Webb, J. G. (2005). N-acetylcysteine for prevention
Risk factors and mechanisms of anaphylactoid reactions to acetylcysteine in acet- of radiocontrast induced nephrotoxicity: The importance of dose and route of admin-
aminophen overdose. Clin Toxicol 46, 697–702. istration. Heart 91, 997–999.
Parfrey, P. (2005). The clinical epidemiology of contrast-induced nephropathy. Cardiovasc Sheffner, A. L., Medler, E. M., Jacobs, L. W., & Sarett, H. P. (1964). The in vitro reduction in
Intervent Radiol 28(Suppl. 2), S3–S11. viscosity of human tracheobronchial secretions by acetylcysteine. Am Rev Respir Dis
Pendyala, L., & Creaven, P. J. (1995). Pharmacokinetic and pharmacodynamic studies of 90, 721–729.
N-acetylcysteine, a potential chemopreventive agent during a phase I trial. Cancer Sosa, V., Moliné, T., Somoza, R., Paciucci, R., Kondoh, H., & Leonart, M. E. (2013). Oxidative
Epidemiol Biomarkers Prev 4, 245–251. stress and cancer: an overview. Ageing Res Rev 12, 376–390.
Pieper, G. M., & Siebeneich, W. (1998). Oral administration of the antioxidant, N-acetyl- Stadler, K. (2012). Oxidative stress in diabetes. Adv Exp Med Biol 771, 272–287.
cysteine, abrogates diabetes-induced endothelial dysfunction. J Cardiovasc Pharmacol Stey, C., Steurer, J., Bachmann, S., Medici, T. C., & Tramer, M. R. (2000). The effect of oral
32, 101–105. N-acetylcysteine in chronic bronchitis: a quantitative systematic review. Eur Respir J
Pompella, A., Visvikis, A., Paolicchi, A., De Tata, V., & Casini, A. F. (2003). The changing 16, 253–262.
faces of glutathione, a cellular protagonist. Biochem Pharmacol 66, 1499–1503. Suk, J. S., Boylan, N. J., Trehan, K., Tang, B. C., Schneider, C. S., Lin, J. G., et al. (2011).
Poole, P. J., & Black, P. N. (2001). Oral mucolytic drugs for exacerbations of chronic N-acetylcysteine enhances cystic fibrosis sputum penetration an airway gene transfer
obstructive pulmonary disease: Systematic review. Br Med J 332, 1–6. by highly compacted DNA nanoparticles. Mol Ther 19, 1981–1989.
Prescott, L. F., Illingworth, R. N., Critchley, J. A., Stewart, M. J., Adam, R. D., & Proudfoot, A. T. Sun, Z., Fu, Q., Cao, L., Jin, W., Cheng, L., & Li, Z. (2013). Intravenous N-acetylcysteine for
(1979). Intravenous N-acetylcysteine: The treatment of choice for paracetamol prevention of contrast-induced nephropathy: a meta-analysis of randomised, con-
poisoning. Br Med J 2, 1097–1100. trolled trials. PLoS One 8, 1–8.
Prescott, L. F., Park, J., Ballantyne, A., Adriaenssens, P., & Proudfoot, A. T. (1977). Treatment of Suzuki, K. (2009). Anti-oxidants for therapeutic use: Why are only a few drugs in clinical
paracetamol (acetaminophen) poisoning with N-acetylcysteine. Br Med J 2, 432–434. use? Adv Drug Deliv Rev 61, 287–289.
Prescott, L. F., Sutherland, G. R., Park, J., Smith, I. J., & Proudfoot, A. T. (1976). Cysteamine, Tattersall, A.B., Bridgman, K. M., & Huitson, A. (1983). Acetylcysteine (Fabrol) in chronic
methionine, and penicillamine in the treatment of paracetamol poisoning. Lancet 308, bronchitis — a study on general practice. J Int Med Res 11, 279–284.
109–113. Tirouvanziam, R., Conrad, C. K., Bottiglieri, T., Herzenberg, L. A., Moss, R. B., & Herzenberg,
Raghu, G., Anstrom, K. J., King, T. E., Lasky, J. A., Martinez, F. J., & Arbor, A. (2012). Prednisolone, L. A. (2006). High-dose oral N-acetylcysteine, a glutathione prodrug, modulates
azathioprine, and N-acetylcysteine for pulmonary fibrosis. N Engl J Med 366, 1968–1977. inflammation in cystic fibrosis. Proc Natl Acad Sci U S A 103, 4628–4633.
Rahman, I., & MacNee, W. (2000). Regulation of redox glutathione levels and gene transcrip- Treweeke, A. T., Winterburn, T. J., MacKenzie, I., Barrett, F., Barr, C., Rushworth, G. F., et al.
tion in lung inflammation: therapeutic approaches. Free Radic Biol Med 28, 1405–1420. (2012). N-acetylcysteine inhibits platelet–monocyte conjugation in patients with
Ratjen, F., & Grasemann, H. (2012). New therapies in cystic fibrosis. Curr Pharm Des 18, type 2 diabetes with depleted intraplatelet glutathione: a randomised controlled
614–627. trial. Diabetologia 55, 2920–2928.
Riga, M. G., Chelis, L., Kakolyris, S., Papadopoulos, S., Stathakidou, S., Chamalidou, E., et al. Turell, L., Radi, R., & Alvarez, B. (2013). The thiol pool in human plasma: the central
(2013). Transtympanic injections of N-acetylcysteine for the prevention of cisplatin- contribition of albumin to redox processes. Free Radic Biol Med 65, 244–253.
induced ototoxicity: A feasible method with promising efficacy. Am J Clin Oncol 36, 1–6. Ullian, M. E., Gelasco, A. K., Fitzgibbon, W. R., Beck, C. N., & Morinelli, T. A. (2005).
Rosenblat, M., & Aviram, M. (1998). Macrophage glutathione content and glutathione N-acetylcysteine decreases angiotensin II receptor binding in vascular smooth muscle
peroxidase activity are inversely related to cell-mediated oxidation of LDL: in vitro cells. J Am Soc Nephrol 16, 2346–2353.
and in vivo studies. Free Radic Biol Med 24, 305–317. van Zandwijk, N., Dalesio, O., Pastorino, U., de Vries, N., & van Tinteren, H. (2000).
Rowe, S. M., & Clancy, J. P. (2006). Advances in cystic fibrosis therapies. Curr Opin Pediatr EUROSCAN, a randomised trial of vitamin A and N-acetylcysteine in patients with
18, 604–613. head and neck cancer or lung cancer. J Natl Cancer Inst 92, 977–986.
Rudd, R. M., Prescott, R. J., Chalmers, J. C., & Johnston, I. D. A. (2007). British Thoracic Waring, W. S. (2012). Criteria for acetylcysteine treatment and clinical outcomes after
Society Study on cryptogenic fibrosing alveolitis: response to treatment and survival. paracetamol poisoning. Expert Rev Clin Pharmacol 5, 311–318.
Thorax 62, 62–66. Waring, W. S., Stephen, A. F., Robinson, O. D., Dow, M.A., & Pettie, J. M. (2008). Lower
Ryan, R. (Ed.). (2012). British National Formulary. BMA, RPS. incidence of anaphylactoid reactions to N-acetylcysteine in patients with high acet-
Saito, C., Zwingmann, C., & Jaechke, H. (2010). Novel mechanisms of protection against aminophen concentrations after overdose. Clin Toxicol 46, 496–500.
hepatotoxicity in mice by glutathione and N-acetylcysteine. Hepatology 51, 246–254. Winterbourne, C. C., & Metodiewa, D. (1999). Reactivity pf biologically important thiol
Samuni, Y., Goldstein, S., Dean, O. M., & Berk, M. (2013). The chemistry and biological compounds with superoxide and hydrogen peroxide. Free Radic Biol Med 27, 322–328.
activities of N-acetylcysteine. Biochim Biophys Acta 1830, 4117–4129. Zalups, R. K., & Ahmad, S. (2005). Transport of N-acetylcysteine s-conjugates of methyl-
Sandilands, E. A., & Bateman, D. N. (2009). Adverse reactions associated with acetylcysteine. mercury in Madin-Darby canine kidney cells stably transfected with human isoform
Clin Toxicol 47, 81–88. of organic anion transporter 1. J Pharmacol Exp Ther 314, 1158–1168.
Sandilands, E. A., Camerson, S., Parterson, F., Donaldson, S., Briody, L., Crowe, J., et al. Zalups, R. K., & Barfuss, D. W. (1998). Participation of mercuric conjugates of cysteine,
(2012). Mechanisms for an effect of acetylcysteine on renal function after exposure homocysteine, and N-acetylcysteine in mechanisms involved in the renal tubular
to radio-graphic contrast material: study protocol. BMC Clin Pharmacol 12, 1–8. uptake of inorganic mercury. J Am Soc Nephrol 9, 551–561.

You might also like