2,128
Views
19
CrossRef citations to date
0
Altmetric
Review

A marriage of two “Methusalem” drugs for the treatment of psoriasis?

Arguments for a pilot trial with metformin as add-on for methotrexate

&
Pages 252-263 | Received 12 Dec 2012, Accepted 04 Feb 2013, Published online: 01 Apr 2013

Abstract

In this article we present arguments that the “antidiabetic” drug metformin could be useful as an add-on therapy to methotrexate for the treatment of psoriasis and, perhaps, for rheumatoid arthritis as well. Biochemical data suggest that both drugs may share a common cellular target, the AMP-activated protein kinase (AMPK). This enzyme is a master regulator of metabolism and controls a number of downstream targets, e.g., important for cellular growth or function in many tissues including T-lymphocytes. Clinical observations as well as experimental results argue for anti-inflammatory, antineoplastic and antiproliferative activities of metformin and a case-control study suggests that the drug reduces the risk for psoriasis.

Patients with psoriasis have higher risk of metabolic syndrome, type 2 diabetes and cardiovascular mortality. Metformin has proven efficacy in the treatment of prediabetes and leads to a pronounced and sustained weight loss in overweight individuals. We expect that addition of metformin to methotrexate can lead to positive effects with respect to the PASI score, reduction of the weekly methotrexate dose and of elevated cardiovascular risk factors in patients with metabolic syndrome and psoriasis. For reasons explained later we suggest that only male, overweight patients are to be included in a pilot trial. On the other side of the coin are concerns that the gastrointestinal side effects of metformin are intolerable for patients under low dose, intermittent methotrexate therapy. Metformin has another side effect, namely interference with vitamin B12 and folate metabolism, leading to elevated homocysteine serum levels. As patients must receive folate supplementation and will be controlled with respect to their B12 status increased hematological toxicity is unlikely to result.

Introduction

It is an undeniable fact that there is a paradigmatic change occurring for one of the most common “skin diseases,” psoriasis. It is now more and more recognized as a systemic disorder.Citation1 Being asked to prepare state-of-the-art review talks (pharmacological and dermatological, respectively) for a conference held in Vienna on behalf of the 50th anniversary of methotrexate we discovered a case-control study indicating that risk for psoriasis was lowered by using either glitazones or metformin.Citation2 Risk reduction for metformin was only observed in males. The authors suggested that, perhaps a common mechanism, namely activation of AMP-activated protein kinase (AMPK) was a possible explanation. Although the glitazones can stimulate AMPK in cellular systems after binding to a mitochondrial receptor (mitoNEET)Citation3 AMPK activation was never (in contrast to metformin) demonstrated ex vivo in humans. Association studies do not prove causality but can stimulate further research. Disregarding the glitazones for reasons not detailed here, we concentrate on metformin to collect arguments for a pilot trial adding this “antidiabetic” as an “anti-inflammatory” agent to the low-dose, weekly methotrexate regimen.

Dermatologists are familiar with methotrexate but less so with metformin, therefore the focus is more on the latter. Endocrinologists on the other hand may be less informed about the biochemical mysteries of the most favored “anti-diabetic” drug. Metformin was introduced in Europe during the same decade when Gubner et al.Citation4 published their results with aminopterin for the treatment of psoriasis and rheumatoid arthritis. First we discuss pharmacodynamics and molecular and cellular activities. A brief overview on the cellular and organismal distribution and disposition-kinetics of the two drugs is also given. We subsequently present biochemical data, animal experiments and clinical observations for our proposal that an “add-on” pilot trial with male psoriasis patients having metabolic syndrome is a rational approach and ethically justified. It must be emphasized that this is not a review article on anti-inflammatory activities of metformin (although we heavily rely on existing evidence) or of methotrexate. Discussions of complex intracellular pathways (including signaling of cytokines, protein kinase cascades, etc.) are beyond our scope. For readers interested in more molecular details we cite key review articles and publications. It is hoped—although both drugs are old (but by no means outdated)—that what was selected from the most recent and long forgotten literature can be fascinating. Perhaps we can stimulate interest of researchers and clinicians in dermatology for metformin and related compounds as systemic (as proposed here for psoriasis) or even as topical drugs.

Methods

We searched the literature via pubmed, Google Scholar, Web of Science with key phrases alone and (more) often in combination: biguanide(s), phenformin, metformin, AICAR, AICAr, methotrexate, T-lymphocytes, monocytes, macrophages, anti-inflammatory, metabolic syndrome, psoriasis, AMP-activated protein kinase, AMPK, reactive oxygen species, complex I (inhibitors), inflammation, inflammatory biomarkers, clinical trials, organic cation transporters. We also approached some of the leading researchers in the field to supply us with further information including publications in press.

Results

Methotrexate and metformin are not only very old drugs but have a remarkable history of an expanding pharmacodynamic profile

Metformin: From an anti-diabetic to anti-inflammatory, anti-aging, rejuvenating, antidote, anti-neoplastic and “anti-androgen” drug

The history of the biguanidines as drugs can be traced back to the Middle Ages. The herb Galega officinalis (French lilac, Goat’s rue) was used among other for symptoms of diabetes mellitus in humans or to increase milk production in farmed animals. The active principles of the toxic (or poisonous) plant are guanidine and isoamylene guanidine (galegine) recognized as hypoglycemic principles early in the 20th century.Citation5 Galegine was used in doses of 150 mg/day with success in more than 3 dozen patients, lacking the abdominal discomfort of synthetic synthalin A (decamethylene-guanidine) and B (dodecamethylene diguanidine).Citation6 The synthalins remained in the pharmaceutical armamentarium of Germany for many years despite their known toxicity. Under the suggestive name of flumamine, injected (most likely i.m. in a dose of 32.5 mg!) metformin enjoyed a brief episode in 1949 as an extremely potent, fast-acting antipyretic and analgesic for viral influenza.Citation7,Citation8 Although this miracle cure was never reproduced, the author suggested that flumamine acts on malaria parasites and promised to report on his malaria cases later. His idea that metformin (and related biguanidines) may be active on malaria parasites is supported by the structural similarity of RJF 00719 [2-(4-cyclohexylphenyl)-1-diaminomethylidene-guanidine)] and RJF 01059 (phenylbiguanide) to phenformin. Both compounds (especially RJF 00719 in the 20 micromolar range) are inhibitors of the Plasmodium falciparum bi-functional thymidilate synthase-dihydrofolate reductase.Citation9

In any event, the publicationCitation8 from the Philippines, mentioning glucose lowering effects of substituted biguanidines in the discussion (but not for his treated patients), stimulated a French researcher (Jean Sterne) 1957 to try “flumamine” for diabetics. Jean Sterne was (luckily) unaware of data published a quarter of a century earlier in the at that time most important pharmacological journal of the world.Citation10 The latter German authors, who investigated the antipyretic (sic) activity of biguanide and its chemically modified analogs, actually strongly warned against a human trial for diabetes: 1, 1-dimethylbiguanide (now metformin) was extremely toxic in their animal experiments. In any event, the above mentioned “miracle” cure was the first hint that metformin and structurally related antidiabetics such as phenformin can exhibit “anti-inflammatory” activity. Indeed, in the pre-methotrexate era phenformin was tried with some success for rheumatoid arthritis because it demonstrated “fibrinolytic” activity.Citation11-Citation13 After the discontinuation of phenformin and buformin in many (but not all) countries, only metformin (introduced in Europe in 1957) survived and is—with respect to tons of drug consumed—the world leader. For example, the influent concentration in 5 German wastewater treatment plants varied from 18 to 105 μg/L,Citation14 which is equivalent to an estimated metformin consumption in Germany of 370–897 tons per year. Metformin is an environmental contaminant, ends up in some plants, can be even highly enriched in seedsCitation15 and may be present in concentration of 1 to 3 μg/L in our drinking water. As a caloric restriction mimetic metformin increases life span in some laboratory animals and the worm C.elegans.Citation16 It is gaining experimental popularity as rejuvenating drug or gerosuppressant, being able to stimulate neurogenesis.Citation17 It acts as an antidote since it completely prevents gentamycin kidney toxicity in rats.Citation18 Highly significant in the context of this article, metformin ameliorates hepatic methotrexate toxicity in experimental animals.Citation19 Of equal interest for dermatology, metformin as a topical (or systemic) treatment blocks UV-B induced tumor formation in hairless miceCitation20 and reduces growth of squamous cell carcinoma.Citation21 As an adjunct to conventional cancer therapy or to prevent relapse more than 40 human trials with metformin are listed in ClinicalTrials.gov (www.clinicaltrials.gov accessed November 7, 2012). There is, however, an ongoing debate, if metformin acts via a systemic effect solely by decreased insulin levels (and/or less obesity) or has direct actions on tumor cells as is observed experimentally.Citation22 A detailed discussion of metformin and related biguanidines for cancer prevention and treatment is beyond the scope of this article and the reader is referred to excellent reviews (see refs. Citation148 and Citation149).

Metformin was first used for women with (and later without) diabetes to increase insulin sensitivity, promote ovulation and decrease androgen production from the ovaries, by lowering the “hyperinsulinemic” status (see ref. Citation23). However, adrenal cells in vitro are highly sensitive to metformin,Citation24 suggesting that subtle alterations of adrenal steroid metabolism may occur by direct action. Adrenal glands and especially the cortex due to the presence of transporters accumulate the drug similar to small intestine, liver, skeletal muscle and kidney.Citation25 As will be outlined later metformin appears to have some beneficial effects only in males, which may be related to differential effects on steroid hormones. Metformin’s status to improve fertility in PCOS has not been convincingly demonstrated,Citation23 but it certainly ameliorates symptoms of this highly prevalent condition. In the context of co-morbidities for psoriasis it is remarkable that PCOS women also have an increased risk of cardiovascular events and diabetes.Citation26

Methotrexate: From an anti-neoplastic (anti-metabolic) to anti-inflammatory, vasculoprotective and “antidiabetic” drug

Methotrexate and its more toxic precursor aminopterin were first introduced as anti-neoplastic agents. More by chance the anti-folates, especially low-dose, once weekly methotrexate, found entry into dermatology, not only for treatment of psoriasis but for numerous other skin conditions.Citation27 Methotrexate is still the most effective disease modifying antirheumatic drug for rheumatoid arthritis and related disorders. More recently, based on data from observational studies with low-dose methotrexate for rheumatoid arthritis or psoriasis indicating lower cardiovascular risk,Citation28 it was suggested to employ methotrexate as “vasculoprotective drug” for patients with stable cardiovascular disease and elevated hsCRP.Citation29 Our concept, namely that part of the action of methotrexate is based on activation of AMPK via AICAR (ZMP), is supported by experiments with genetically diabetic (db+/db+) mice, where glucose and insulin levels are significantly lowered after only 4 weekly i.p. methotrexate doses (0.5 mg/kg).Citation30 Unfortunately, the authors failed to demonstrate that either AMPK or own of its downstream effectors such as acetyl-Coenzyme A -carboxylase were phosphorylated.

Metformin and methotrexate have some common features

Uptake via carriers or transporters, elimination via the kidney and “deep compartement”

Both drugs are taken up into cells (including enterocytes of the intestine) by specific carrier/transporter systems. Methotrexate enters as a “blind passenger” via folate uptake mechanisms and is trapped intracellularly by polyglutamination.Citation27 Metformin is taken up via organic cation- (OCTs)Citation31 and plasma membrane monamine transporters.Citation32 Secretion of metformin by kidney tubules and liver luminal membranes is driven by multidrug and toxin extrusion (MATE) proteins.Citation33 Much is known about the pharmacogenetics of metformin especially for naturally occurring variants of OCT1 (import) and OCT2 (export by the kidney) but less so for methotrexate. For both drugs, systemic bioavailability after oral administration is dose-dependent and limited in capacity: the higher the dose the less percentage of it is adsorbed. Both are (mainly) excreted via the kidney albeit via different transporters. Therefore plasma clearance is depending on kidney function. Under conditions of weekly (MTX) or daily dosing (metformin) a “deep” compartement is building up. Higher polyglutamated (MTX-PGs) species slowly increase, if red blood cell MTX-PG content is taken as proxy for tissue (or target cell, “site of action”) residence. A steady-state with respect to the composition of red blood cell MTX-PGs is only reached after several months,Citation34 often paralleling clinical improvement.Citation35 Although on a different time scale (hours and days), the distribution volume for metformin increases from about 1.6 L/kg after a single doseCitation36 to about 4 L/kgCitation25 under continuous dosing. This indicates a large tissue reservoir, mainly but not only dominated by expression and activity of the various transporters. Similar to MTX (and the MTX-PGs), metformin accumulates in red blood cellsCitation37 reflecting “deep compartements” elsewhere. Existence of such compartements is also apparent by late phases in plasma or urine elimination kinetics after long-term daily dosing.Citation38 Metformin has a pK of 11.5 and is highly water soluble. According to first principles in pharmacology, as a singly positively charged molecule at almost all pH values in the body, it prefers to reside in lower pH in preference to higher pH. Therefore metformin can also distribute via simple diffusion but on a much prolonged time-scale in contrast to rapid transport via OCTs. The average metformin concentrations at steady-state and daily doses of 2–3 g in plasma of humans are around 1 mg/L (8 micromolar) with an upper limit of 2.4 mg/L.Citation25 After oral intake peak levels in portal blood exposing circulating blood cells and, later, hepatocytes are likely to be much higher.Citation39 Within cells 1000-fold higher concentrations compared with the cytosol are calculated to exist in the intermembrane space of respiring, phosphorylating mitochondria with a high membrane potential.Citation40 Similar arguments may hold for lysosomes, endosomes, caveolae, Golgi and other acidified organelles. Surprisingly little attention has been paid to the fact that the transporters involved in uptake and extrusion of metformin were not designed by nature for drugs but mainly for endogenous substrates. The Km values for metformin are ~1 mM for OCT1Citation41 and 0.24–2 mM for MATE1 and MATE2, respectively.Citation33 Considering that experiments with cells are often performed with 1 or 10 mM metformin for many (24–48–72) h (see refs. Citation42Citation44), one wonders if metformin as a competing substrate did not change import and export of endogenous metabolites or nutrients. One metabolite which is substrate for both importers and exporters is the essential amino acid tryptophan. The Km values for this amino acid for OCT1 or OCT2 are in the 5–10 mM range ~ten times higher than for metformin. Plasma concentrations of tryptophan are ~60 μMCitation45 and those for metformin are up to 20–30 μM, making competition very likely. Indeed, upon chronic dosing with metformin plasma or serum levels of tryptophan in patients increase.Citation46,Citation47 Tryptophan concentrations in urine have recently proposed as markers to predict patients pharmacokinetic parameters for metformin determined by the naturally occurring human variants of OCT2.Citation48 Thus, it cannot be excluded that part of the metabolic and gene expression changes observed after organ perfusion, cellular experiments and in patients are a consequence of substrate competition with transporters. It is interesting to note that competition of the hypoglycemic biguanidines with mitochondrial (not plasma membrane) cation transporters was suggested as early as 1979.Citation42

The mechanism by which methotrexate and metformin act is still not fully understood

Methotrexate: Are adenosine, AICAR or reactive oxygen species (ROS) mediators of the anti-inflammatory activity?

In we show a highly simplified version of the mechanism by which methotrexate is proposed to act as an “anti-inflammatory agent,”Citation27,Citation43 namely by increasing adenosine efflux. Enhanced urinary secretion of AICA (5- aminoimidazole-4-carboxamide), a degradation product of AICAR and of adenosine, was first observed in patients with psoriasis treated with methotrexate,Citation44 pointing to an inhibitory action of MTX (or MTX-PGs) at the level of AICAR transformylase (AICART). The importance of adenosine as an anti-inflammatory mediator was questionedCitation53,Citation54 and could not been convincingly demonstrated in later experiments.Citation55 It is, however, generally accepted that AICAR (ZMP) increases after methotrexateCitation56—either by direct inhibition of AICART or alternatively, favoring the reverse reaction by increasing dihydrofolate via blockade of dihydrofolate reductase.Citation52 (For more details, see the legend of ) Another inhibitor of thymidylate synthase and AICART, pemetrexed, increases ZMP and activates AMPK.Citation57-Citation59 In summary, experimental evidence in cell culture, animal experimentsCitation30 and observations in patients support the view that methotrexate and related compounds can increase AICAR production and activate AMPK in addition to their inhibitory effects on nucleotide synthesis, which is still the claimed basis for immune-suppressive activity. Nevertheless, the molecular mechanisms underlying the “anti-inflammatory” activity and the postulated targets cells (T-lymphocytes in psoriasis, macrophages and monocytes in rheumatoid arthritis) are still not completely understood. The increased AICAR production and appearance of its degradation product AICA in the general circulation after methotrexate are intriguing. Does the release indicate rapid apoptosis of target cells or a continuous excretory mechanism? In any event, AICA and perhaps the corresponding riboside could act as a kind of “second messenger” to neighboring cells or elsewhere. The riboside may be taken up by the ubiquitous adenosine transporter and phosphorylated by adenosine kinase to ZMP. The purine base AICA may be taken up by nucleobase transportersCitation60 and converted via adenine-phosphoribosyl-transferase to ZMP. That AICA (and the riboside) could be the mediators or signals for clinical efficacy was already suggested by.Citation55 Reactive oxygen species (ROS) as mediators for methotrexate were postulated,Citation54,Citation61 but this was not supported by subsequent research.

Figure 1. Metformin and methotrexate may interact at the cellular level via AMP-activated protein kinase. Simplified schemeCitation43,Citation49 of the proposed “anti-inflammatory” activity of methotrexate. Increase of AICAR (structure shown in the inset) by inhibition of the bifunctional enzyme AICAR transformylase (AICART or ATIC)Citation50 is suggested to inhibit AMP and adenosine deaminase, leading to increased release of adenosine. Although it is generally postulated that the inhibition of AICART is by direct action of methotrexate polyglutamates, there are alternative explanations. These are based on the reversibility of the reaction (indicated by the double-headed arrow) and that AICART has rather low affinity for the polyglutamates. As an exception to all other reactions in mammalian one-carbon metabolismCitation51 10-formyl-7, 8-dihydrofolate and dihydrofolate can be substrate and product, respectively.Citation52 Increase of the product by blockade of the dihydrofolate reductase will slow the reaction and lead to increase in AICAR. FPGS, Folyl-Poly-Glutamyl- Synthase. Green arrows indicate activation; red lines with masthead indicate inhibition, broken line: disputed activity.

Figure 1. Metformin and methotrexate may interact at the cellular level via AMP-activated protein kinase. Simplified schemeCitation43,Citation49 of the proposed “anti-inflammatory” activity of methotrexate. Increase of AICAR (structure shown in the inset) by inhibition of the bifunctional enzyme AICAR transformylase (AICART or ATIC)Citation50 is suggested to inhibit AMP and adenosine deaminase, leading to increased release of adenosine. Although it is generally postulated that the inhibition of AICART is by direct action of methotrexate polyglutamates, there are alternative explanations. These are based on the reversibility of the reaction (indicated by the double-headed arrow) and that AICART has rather low affinity for the polyglutamates. As an exception to all other reactions in mammalian one-carbon metabolismCitation51 10-formyl-7, 8-dihydrofolate and dihydrofolate can be substrate and product, respectively.Citation52 Increase of the product by blockade of the dihydrofolate reductase will slow the reaction and lead to increase in AICAR. FPGS, Folyl-Poly-Glutamyl- Synthase. Green arrows indicate activation; red lines with masthead indicate inhibition, broken line: disputed activity.

AICAr is often employed as an experimental tool in animal models for disease (see ) or given i.v. in human studies (as acadesine)Citation63,Citation64 to activate AMPK as an “AMP mimetic.” Despite the expectation that the increased glucose uptake of human skeletal muscle after AICAr infusion was mediated by AMPK activation no increased phosphorylation of the master kinase was observed. Instead the extracellular signal-regulated kinase 1/2 (ERK1/2) was phosphorylated.Citation65 A word of caution is appropriate here, as some of the experiments claiming AMPK activation as a result of AICAr treatment fail to rigorously prove the claim as direct activators (A-769662) have differential effects on pyrimidine metabolism and the purine salvage pathway compared with AICAr.Citation66 Notably, however, is that AICAr can influence human glucose metabolism in vivo, underlining the hypothesis that it can serve as a signalCitation55 not necessarily via AMPK.

Table 1. Metformin and AICar exhibit anti-inflammatory and immunosuppressive activity in experimental systems

Metformin: A drug in search of a primary target

Mitochondrial complex I inhibition, the mitochondrial permeability pore, the metformin-copper hypothesis and phosphatase inhibition

Ever since guanidines were reported to inhibit energy transfer reactions of mitochondria in 1963,Citation67 interest has focused on their effects mitochondrial function, including transport of divalent cations,Citation68 the still unexplained reversal of respiratory inhibition by uncouplers or fatty acidsCitation69 and much more. Accepted today as a prime (not primary!) target is complex I. Complex I (NADH-ubichinone-reductase) is part of the respiratory chain located in the inner mitochondrial membrane. This huge multi-subunit complex (around 950kDa) pumps four protons out of the matrix into the inter-membrane space tightly coupled to the reduction of ubichinone and oxidation of NADH +H+. Metformin has inhibitory action (only) on complex I of mitochondriaCitation40,Citation70 and can thereby change the cellular ATP production via oxidative phosphorylation. The mystery is that the inhibitory effect (which is never higher than about 40% of the maximal activity) can only be observed when metformin acts on “intact” isolated cells (permeabilized cells are unresponsive) or after organ perfusion with the drug. Mitochondria isolated from such pretreated cells appear irreversibly impregnated, as if metformin somehow, passing plasma cell membranes, acquired a novel property which cannot be mimicked by mixing cytosol and/or plasma membrane particles in the presence of metformin with naive mitochondria. This has led one group of the original discoverers to the suggestion that metformin somehow changed the mitochondrial machinery via phosphorylation/ dephosphorylation or proteolysis.Citation70 One argument was that the degree of inhibition of complex I in isolated intact hepatocytes but not its kinetics was a clear function of the temperature, being optimal at 37°C and almost absent at 15°C. However, it was not known at the time that transport of metformin via OCT1 is highly temperature dependent.Citation41 In any event, inhibition of complex I and the resulting changes of the AMP/ATP ratios are believed to be responsible for the activation of AMPKCitation71 (see below). It should be noted that in contrast to AMPK activation, demonstrated in humans after chronic dosing with metformin, complex I inhibition of isolated mitochondria from skeletal muscle biopsies was not observed.Citation72

Inhibition of complex I by metformin (similar and additive to blockade of cyclophylin D via cyclosporine A) sensitizes the permeability transition pore (PTP) in the inner mitochondrial membrane to attain a state of lower open probability.Citation73,Citation74 The pore plays an important role in cell death induced by free radicals, UV radiation, oxidative stress or ischemia. Results with metformin as a “protecting agent”Citation18,Citation20 suggest that there may be benefits of the drug beyond AMPK activation as explained in the following paragraph.

Inhibition of complex I of mitochondria isolated from cells pretreated with metformin leads to a block in “reverse electron flux”-induced superoxide anion formation.Citation75 Superoxide anions may be formed by escaping electrons at complex I and III. When there is, in simple words, some kind of block downstream of complex I, electrons can flow back, reduce NAD to NADH, increase the ratio of ubiquinol to ubiquinone and contribute to superoxide formation. Superoxide anion cannot pass the inner mitochondrial membrane and is converted to hydrogen peroxide by manganese superoxide dismutase. In the cytosol H2O2 can activate redox-dependent transcription factors and is an indispensable signal (as is calcium) for T-cell activation.Citation76,Citation77 Thus, metformin by reducing ROS, can profoundly change T-cell responses in experimental systems.

Metformin and related biguanidines are excellent chelators (e.g., zinc and copper).Citation78,Citation79 Metformin can form stable complexesCitation80 with copper (association constant around 1016 M−1). A more apolar (hydrophobic) dimeric metformin Cu 2+ complex, which must pass the plasma membrane, slowly forms inside cells and accumulates in mitochondria is now suggested to be the key mediator.Citation81,Citation82 If so, metformin could be a “prodrug,” but the final activity may be not via simple copper chelation.

Other biguanido metal complexes are very potent inhibitors of protein phosphatasesCitation83 suggesting that copper or other metal proteins could be among the long sought “primary” targets. Metformin can inhibit tyrosine phosphatases (such as recombinant human protein phosphatase 1 B) in vitro at very low concentrations (8 μM) but only if plasma membrane fragments (from oocytes) are added.Citation84 One wonders about this truly magic potentiation, but perhaps, formation of an active “intermediate” with a phospholipid, a metal ion or a metallo protein occurred.

AMPK activation

Metformin activates a master regulator in metabolism, the AMPK,Citation77,Citation85 via increased phosphorylation of its catalytic α-subunit in position threonine-172, mediated by LKB1. This drug-induced activation is proven by animal experiments,Citation86 in experimental cell systemsCitation71 as well as in humans after chronic dosage.Citation87,Citation88 AMPK consists of three subunits where the regulatory gamma subunit is functioning as a cellular AMP or ADP sensor, which changes its conformation upon binding AMP or ZMP or other activators such as salicylateCitation89 and increases the availability of the α subunit to be substrate of upstream protein kinases. Metformin binds to the gamma subunit of AMPK in vitroCitation90 but in contrast to direct activators such as A-769662,Citation91 salicylate, ZMP, AMP or OSU-53Citation92 activation of the enzyme in cell free systems has never been reported. Metformin cannot activate AMPK in cells where the gamma subunit of the enzyme is manipulated to be unresponsive to AMP. Consequently it is generally acceptedCitation71 that by lowering the cellular energy status and changing the ATP /AMP ratio via mitochondrial inhibition AMPK is activated. An alternative view, namely that metformin acts by inhibiting AMP deaminaseCitation93 is questioned.Citation85 It is not necessary here to discuss the many downstream targets of AMPK, its connections to the mTOR signaling pathwayCitation94 or its role in the regulation of sirtuins.Citation95 We also omitted that increased intracellular calcium (via calmodulin-dependent protein kinase β) can phosphorylate and activate AMPK, as no evidence so far was found that metformin acts through this pathway.

The physiological role of the AMPK is mainly to control metabolism, depending on nutrient supply. Thus it is natural to assume its main action in diabetics is on liver, adipose tissue and skeletal muscle. However, there is good evidence that this master protein kinase is also involved in the regulation of inflammatory responses and has controlling activity in dendritic cells, T-lymphocytes, macrophages, endothelial cells and monocytes. This is no surprise, as often these cells must either rapidly proliferate (and switch to more glycolysisCitation96,Citation97) or synthesize, e.g., cytokines for secretion and other mediators, which require major changes in metabolism and import of nutrients.

AMPK independent activities of metformin

Inhibition of mTORC1 (MammalianTarget of Rapamycin Complex I) or activation of protein phosphatase 2A by metformin (and phenformin) were shown to be independent from AMPK,Citation98-Citation100 as well as most of the stimulation of glucose uptake in myotubes.Citation101

Is metformin a methotrexate mimetic?

Methotrexate increases plasma homocysteine levels in patients with psoriasis.Citation102,Citation103 Metformin (upon long-term treatment) can lower plasma vitamin B12Citation104 and increases plasma homocysteine levels.Citation105,Citation106 The underlying assumption is that metformin impairs the intestinal reabsorption of vitamin B12. However, short-term studies in diabetic patients,Citation107 non-diabetic malesCitation108 and patients with PCOSCitation109 indicate that the increase in homocysteine levels is not correlated with vitamin B12 deficiency. Plasma levels of the most sensitive indicator of vitamin B12 deficiency, methymalonic acid, are not changed. Furthermore, the increase in homocysteine can be completely prevented or corrected by administration of folate and conventional doses of oral vitamin B12 are unable to attain “normal levels” in metformin treated patients.Citation104 The reversal by folate (and not by vitamin B12) indicates that metformin inhibits target(s) in tetrahydrofolate-dependent metabolic pathways. These observations stimulated a study in which metformin was employed at rather high concentrations in several breast cancer cell lines. Here it was found, by analysis of metabolic changes, that metformin dramatically increased the intracellular levels of 5-formiminotetrahydrofolate.Citation110 This metabolite is the product of histidine catabolism to glutamate and can be converted by two subsequent enzymatic steps to N,Citation5 NCitation10-methylene-tetrahydrofolate. This is the essential substrate (not cofactor) for thymidilate synthase. The authors did not speculate how metformin acts as an “antifolate” but as expected from the ability of metformin to block transport of endogenous substrates (see the section “Uptake via carriers or transporters, elimination via the kidney and “deep compartement”) they observed a marked decrease in intracellular tryptophan, which delivers formate to tetrahydrofolate in the pathway to kynurenines and is precursor for NAD+. Metformin also decreased glutathione which was confirmed by others.Citation111

Consult for a brief summary of the accepted targets for metformin and methotrexate.

Table 2. Molecular mechanisms and targets for metformin and methotrexate

Metformin is an anti-inflammatory drug

Evidence that metformin or/and AMPK activation via AICAr exhibit anti-inflammatory and immunosuppressive activity in cellular and animal experiments

In , we present key publications that support the view that metformin (and AICAr as an AMP-mimetic) is anti-inflammatory and immunosuppressive in various experimental mouse models of autoimmune diseases. The key role of AMPK in macrophages was demonstrated by Sag et al.Citation115 They observed that anti-inflammatory stimuli such as IL-10 or TGFβ increase levels of phosphorylated AMPK within minutes, whereas stimulation of Toll-like receptors (TLR4) by lipopolysacccaride (LPS) decreases AMPK activation. TNFα, IL-6 secretion after LPS was increased by stably transfecting with a dominant negative AMPK but IL-10 decreased. Exactly opposite changes were observed by transfection with a constitutively active AMPK. In dendritic cells the anti-inflammatory cytokine IL-10 partially blocked the hypophosphorylation of AMPK by LPS and AICAr blocked the LPS induced maturation.Citation116 The role of AMPK in T-lymphocytes is reviewed by ref. Citation117. All human white blood cells (neutrophils > monocytes > lymphocytes) express OCT-1 and can transport metformin.Citation118 Human monocytes in vitro are highly sensitive to metformin. Concentrations observed in patients (10 mM) blocked tissue factor activity increase and TNF release after exposing cells to oxidized LDL or LPS.Citation119 T-lymphocytes play a key role in psoriasis and other inflammatory processes therefore inhibition of complex I by metforminCitation77 and reduced ROS formation are possible beneficial actions independent from AMPK.

In , we show a highly simplified scheme, how metformin and AICAr may exert anti-inflammatory and immunosuppressive activity. (For more molecular details, see ref. Citation121.)

Figure 2. Simplified scheme how metformin and AICAr act as anti-inflammatory or immunosuppressive agents. Green arrows indicate activation or stimulation; Red lines with masthead indicate inhibition. Metformin may not only stimulate AMPK but block excessive reactive oxygen species (ROS) formation by mitochondria, which can activate the Nuclear Factor (NF)–ΚB pathway (not shown). For details on the role of AMPK in macrophages, see ref. Citation115. Molecular detailed pathways are found in ref. Citation120.

Figure 2. Simplified scheme how metformin and AICAr act as anti-inflammatory or immunosuppressive agents. Green arrows indicate activation or stimulation; Red lines with masthead indicate inhibition. Metformin may not only stimulate AMPK but block excessive reactive oxygen species (ROS) formation by mitochondria, which can activate the Nuclear Factor (NF)–ΚB pathway (not shown). For details on the role of AMPK in macrophages, see ref. Citation115. Molecular detailed pathways are found in ref. Citation120.

Evidence that metformin decreases inflammatory markers in patients

In the first section of our Results, we mentioned briefly that patients with rheumatoid arthritis were successfully treated with phenformin. The intervention was based on the hypothesis that phenformin acts as a fibrinolytic drug. Indeed clinical symptoms of inflammation improved in some patients and the elevated erythrocyte sedimentation rate (ESR), substituted now by CRP, fell significantly. The same authors treated 27 patients with coronary-artery disease for four months with 1.500 mg metformin/day, followed by two months placebo. Twenty-one patients (there were seven drop-outs) demonstrated an increase of plasminogen activator and lower cholesterol. They also noted that the effect was greatest in patients with lowest fibrinolytic activity and that upon cessation of the drug it took a month or more until the starting level was reached.Citation122 Significant decreases of tissue-type plasminogen activator inhibitor and of von Willebrand factor were later confirmed in a randomized double-blind controlled trial with obese subjects but no overt diabetes.Citation121 Other symptoms of the metabolic syndrome, such as higher systolic blood pressure and elevated LDL-cholesterol also improved.Citation123 In non-obese patients with type 2 diabetes, comparing metformin with repaglinide, markers of endothelial dysfunction (plasma plasminogen activator inhibitor-1antigen, von Willebrand factor) and TNFα were significantly decreased by metformin.Citation124 A recent placebo-controlled study in which subjects with impaired fasting glucose (and other symptoms of metabolic syndrome) were treated for 3 mo with simvastatin received metformin as “add-on” experimental treatment. Here it was observed that even in the background of statin-induced “anti-inflammatory” activity, metformin inhibited lymphocyte cytokine release (TNF α, Interferon γ, Interleukin 2).Citation117 There are a few studies with metformin in patients with PCOS. A randomized comparative study proves a significant decrease of IL-6,Citation125 which is confirmed by observations in Taiwan.Citation126 Others found significant decreases of hsCRPCitation127,Citation128 after metformin. With the possible exception of the early discovery of the anti-inflammatory and anti-fibrinolytic activities of the biguanides, all other studies with metformin were performed in patients or subjects with elevated basal insulin levels or increased insulin resistance. The generally accepted interpretation is that the decrease of inflammatory markers is mediated by increased insulin sensitivity or lowering of insulin levels in PCOS and weight loss. However, the animal models listed in argue that anti-inflammatory effects of metformin are observable under experimental conditions where insulin and/or obesity are not drivers of inflammation.

Psoriasis and metabolic syndrome

Inflammatory signaling mechanisms play an important role in psoriasisCitation129 and may contribute for developing so-called co-morbidities, such as higher cardiovascular risk or type 2 diabetes.Citation130 Indeed—depending on severity—psoriasis is associated with metabolic syndromeCitation131-Citation133 and diabetes.Citation134,Citation135 Cohort studies prove an association of psoriasis with increased, mainly cardiovascular mortality.Citation136 Interrogation of the psoriasis transcriptome in comparison to healthy controls revealed links to pathways involved in atherosclerosis, fatty acid metabolism, diabetes and hypertension.Citation137 A recent meta-analysis including all results from this and other studies, employing ingenuity pathway analysis, found overrepresented canonical pathways for atherosclerosis, cancer and cardiovascular disease. Interestingly, pathways identified in the pathogenesis of multiple sclerosis (see ) were also among the top five.Citation138 Needless to say, that the same pathways are in the focus of current research with metformin.

Expanded intra-abdominal adipose tissue is suggested to play an important pro-inflammatory role in obese patients with metabolic syndrome and psoriasis. Correcting extreme obesity with bariatric surgery indeed improves psoriasis,Citation139 and one study suggests that weight loss per se had an additive effect to improve PASI scores.Citation140 As will be discussed below metformin is able to prevent progression of patients with metabolic syndrome to overt diabetes and can induce significant and permanent weight loss solely due to reduction of adipose tissue.

Metformin is the only drug with a favorable benefit/risk profile to prevent progression to type 2 diabetes for patients with metabolic syndrome and/or reduced glucose tolerance

Although glitazones and α-glucosidase inhibitors may prevent progression of prediabetesCitation141 or favor regressionCitation142 to normoglycemia we will not discuss these drugs. Instead a few results of the metformin and lifestyle intervention study, DPP (diabetes prevention program)Citation143 and its open label follow-up studyCitation144 are interesting in the context of our proposal. Metformin reduced incidence of metabolic syndrome and favored its regression but only in males.Citation145 The authors speculated that hormonal alterations are responsible for this highly unexpected result. Weight loss is highly correlated with adherence to metformin and solely due to loss of adipose tissue.Citation146

Conclusion

Metformin is a logical add-on therapy for overweight male patients with psoriasis and metabolic syndrome treated with methotrexate

Several of the newly introduced, highly expensive drugs for psoriasis (e.g., anti-cytokines) are “targeted therapeutics.” In contrast, although approved as an antidiabetic or employed to lower hyperinsulinemic conditions (PCOS) a primary target of the long off-patent and cheap metformin is not (yet) known. This probably explains complete lack of clinical observations not to mention human trials, based on its ability to act, among other, as a powerful anti-inflammatory and immunosuppressive drug. Without a doubt, metformin is—next to lifestyle intervention and for example, statins—a most logical choice for male overweight psoriatric patients with metabolic syndrome. This is ethically justified because evidence exists in the literature that it can prevent further detoriation of prediabetes or regress symptoms of metabolic syndrome.

We collected arguments to support the hypothesis that there may be additive anti-inflammatory effects of metformin for methotrexate-treated patients, which can be first tested in a prospective, open-label clinical pilot trial. Perhaps of equal importance, metformin decreases hepatotoxicity of methotrexate in animal experiments. This observation alone, taken up as a lead, could stimulate clinicians first to check retrospectively if metformin, once started, improved signs for hepatic injury after methotrexate.

Fifty years ago, researchers who treated patients with rheumatoid arthritis (successfully) and coronary artery disease with phenformin or metformin, respectively, could measure only a few laboratory parameters.

Today’s diagnostic armentarium (including the OMICS world) is much greater and in addition to for example, the PASI score, biomarkers systemically and locally can be followed. Both drugs have the same limiting factor, namely kidney function. A major concern is abdominal discomfort after metformin,Citation20,† which may be reduced by starting with low doses (e.g., 0.5 g/day) and retarded formulations, slowly escalating up to the tolerated limit of 2.5 g/day. As a further safety measure, a metformin drug pause is suggested for the day methotrexate is given. Of less importance is the mechanistically unexplained increase of serum or plasma homocysteine after metformin because patients routinely take folate.

If our hypothesis is supported by favorable results in a pilot trial, perhaps one can expect less use of “targeted” drugs in the future. Even if our hypothesis of an additive anti-inflammatory effect is disproven, male, overweight patients with psoriasis and metabolic syndrome, given that metformin is tolerated, will profit from the intervention.

Abbreviations:
AICA=

5-aminoimidazole-4-carboxamide

AICART=

AICAR transformylase

AICAr=

5-aminoimidazole-4-carboxamide ribonucleoside

AICAR=

5-aminoimidazole-4-carboxamide ribonucleotide, synonymous with ZMP

AICART=

AICAR transformylase

AMPK=

AMP-activated protein kinase (Thr/Ser)

DPP=

Diabetes Prevention Program

IL=

interleukin

LDL=

low density lipoprotein

LPS=

lipopolysaccaride

MATE=

multidrug and toxin extrusion proteins

mTORC1=

Mammalian Target of Rapamycin Complex I

MTX=

methotrexate

MTXPG=

methotrexate polyglutamate

OCT=

organic cation transporter

PASI=

Psoriasis Area and Severity index

PCOS=

polycystic ovary syndrome

PTP=

permeability transition pore

ROS=

reactive oxygen species

TLR=

toll like receptor

TNF=

tumor necrosis factor

PTP=

permeability transition pore

Acknowledgments

The authors would like to thank numerous colleagues who supplied us with additional information including papers in press or reprints. Prof. Alan Dronsfield informed us on the real facts behind the “Magic Cure” with flumamime. Librarians of the University of Innsbruck were helpful providing us with long forgotten, not easily obtainable literature. Paula Olbrich and Johannes Werner supported us in organizing the literature, design of figures and more. We are very grateful for their help.

Disclosure of Potential Conflicts of Interest

Both authors have received an honorarium by Pfizer for their contributions at a Symposium in Vienna on occasion of the 50th anniversary of methotrexate. No other conflicts of interest exist with respect to the above work.

Notes

† If abdominal discomfort is a problem and since dermatology is interested in transdermal delivery or local therapy, the reader is referred to www.freepatentsonline.com/y2012/0283332.html.

References

  • Li W, Han J, Hu FB, Curhan GC, Qureshi AA. Psoriasis and risk of type 2 diabetes among women and men in the United States: a population-based cohort study. J Invest Dermatol 2012; 132:291 - 8; http://dx.doi.org/10.1038/jid.2011.319; PMID: 21993559
  • Brauchli YB, Jick SS, Curtin F, Meier CR. Association between use of thiazolidinediones or other oral antidiabetics and psoriasis: A population based case-control study. J Am Acad Dermatol 2008; 58:421 - 9; http://dx.doi.org/10.1016/j.jaad.2007.11.023; PMID: 18194825
  • Wiley SE, Murphy AN, Ross SA, van der Geer P, Dixon JE. MitoNEET is an iron-containing outer mitochondrial membrane protein that regulates oxidative capacity. Proc Natl Acad Sci U S A 2007; 104:5318 - 23; http://dx.doi.org/10.1073/pnas.0701078104; PMID: 17376863
  • Gubner R, August S, Ginsberg V. Therapeutic suppression of tissue reactivity. II. Effect of aminopterin in rheumatoid arthritis and psoriasis. Am J Med Sci 1951; 221:176 - 82; http://dx.doi.org/10.1097/00000441-195102000-00009; PMID: 14799481
  • Bailey CJ, Day C. Metformin:its botanical background. Pract Diabetes Int 2004; 21:115 - 7; http://dx.doi.org/10.1002/pdi.606
  • Reinwein H. Ueber die therapeutische Verwendbarkeit des Galegins bei Diabetikern. Munch Med Wochenschr 1927; 74:1794 - 5
  • Dronsfield A, Ellis P. Drug discovery: metformin and the control of diabetes. Education in Chemistry. 2011;185–7.
  • Garcia EY. Flumamine, a new synthetic analgesic and anti-flu drug. J Philipp Med Assoc 1950; 26:287 - 93; PMID: 14779282
  • Dasgupta T, Chitnumsub P, Kamchonwongpaisan S, Maneeruttanarungroj C, Nichols SE, Lyons TM, et al. Exploiting structural analysis, in silico screening, and serendipity to identify novel inhibitors of drug-resistant falciparum malaria. ACS Chem Biol 2009; 4:29 - 40; http://dx.doi.org/10.1021/cb8002804; PMID: 19146480
  • Hesse G, Taubmann G.. Die Wirkung des Biguanids und seiner Dervate auf den Zuckerstoffwechsel. Naunyn-Schmiedebergs Arch Exp Path Pharmacol 1929; (142):290 - 308; http://dx.doi.org/10.1007/BF02000097
  • Fearnley GR, Chakrabarti R. Fibrinolytic treatment of rheumatoid arthritis with phenformin plus ethyloestrenol. Lancet 1966; 2:757 - 61; http://dx.doi.org/10.1016/S0140-6736(66)90360-6; PMID: 4162320
  • Fearnley GR, Chakrabarti R, Hocking E. PHENFORMIN IN RHEUMATOID ARTHRITIS A FIBRINOLYTIC APPROACH. Lancet 1965; 285:9 - 13; http://dx.doi.org/10.1016/S0140-6736(65)90921-9
  • Kersley GD. Phenformin (Dibotin) in polyarthritis. Ann Rheum Dis 1968; 27:374 - 6; http://dx.doi.org/10.1136/ard.27.4.374; PMID: 4875263
  • Scheurer M, Michel A, Brauch H-J, Ruck W, Sacher F. Occurrence and fate of the antidiabetic drug metformin and its metabolite guanylurea in the environment and during drinking water treatment. Water Res 2012; 46:4790 - 802; http://dx.doi.org/10.1016/j.watres.2012.06.019; PMID: 22770965
  • Eggen T, Lillo C. Antidiabetic II drug metformin in plants: uptake and translocation to edible parts of cereals, oily seeds, beans, tomato, squash, carrots, and potatoes. J Agric Food Chem 2012; 60:6929 - 35; http://dx.doi.org/10.1021/jf301267c; PMID: 22712757
  • Onken B, Driscoll M. Metformin induces a dietary restriction-like state and the oxidative stress response to extend C. elegans Healthspan via AMPK, LKB1, and SKN-1. PLoS One 2010; 5:e8758; http://dx.doi.org/10.1371/journal.pone.0008758; PMID: 20090912
  • Menendez JA, Vazquez-Martin A. Rejuvenating regeneration: metformin activates endogenous adult stem cells. Cell Cycle 2012; 11:3521 - 2; http://dx.doi.org/10.4161/cc.21878; PMID: 22935702
  • Morales AI, Detaille D, Prieto M, Puente A, Briones E, Arévalo M, et al. Metformin prevents experimental gentamicin-induced nephropathy by a mitochondria-dependent pathway. Kidney Int 2010; 77:861 - 9; http://dx.doi.org/10.1038/ki.2010.11; PMID: 20164825
  • Hadi NR, Al-Amran FG, Swadi A. Metformin ameliorates methotrexate-induced hepatotoxicity. J Pharmacol Pharmacother 2012; 3:248 - 53; http://dx.doi.org/10.4103/0976-500X.99426; PMID: 23129960
  • Wu CL, Qiang L, Han W, Ming M, Viollet B, He YY. Role of AMPK in UVB-induced DNA damage repair and growth control. Oncogene 2012; •••:1 - 8; http://dx.doi.org/10.1038/onc.2012.279; PMID: 22751115
  • Chaudhary SC, Kurundkar D, Elmets CA, Kopelovich L, Athar M. Metformin, an antidiabetic agent reduces growth of cutaneous squamous cell carcinoma by targeting mTOR signaling pathway. Photochem Photobiol 2012; 88:1149 - 56; http://dx.doi.org/10.1111/j.1751-1097.2012.01165.x; PMID: 22540890
  • Martin M, Marais R. Metformin: a diabetes drug for cancer, or a cancer drug for diabetics?. J Clin Oncol 2012; 30:2698 - 700; http://dx.doi.org/10.1200/JCO.2012.42.1677; PMID: 22565000
  • Misso ML, Teede HJ, Hart R, Wong J, Rombauts L, Melder AM, et al. Status of clomiphene citrate and metformin for infertility in PCOS. Trends Endocrinol Metab 2012; 23:533 - 43; http://dx.doi.org/10.1016/j.tem.2012.07.001; PMID: 22939889
  • Hirsch A, Hahn D, Kempná P, Hofer G, Nuoffer J-M, Mullis PE, et al. Metformin inhibits human androgen production by regulating steroidogenic enzymes HSD3B2 and CYP17A1 and complex I activity of the respiratory chain. Endocrinology 2012; 153:4354 - 66; http://dx.doi.org/10.1210/en.2012-1145; PMID: 22778212
  • Graham GG, Punt J, Arora M, Day RO, Doogue MP, Duong JK, et al. Clinical pharmacokinetics of metformin. Clin Pharmacokinet 2011; 50:81 - 98; http://dx.doi.org/10.2165/11534750-000000000-00000; PMID: 21241070
  • Mani H, Levy MJ, Davies MJ, Morris DH, Gray LJ, Bankart J, et al. Diabetes and cardiovascular events in women with polycystic ovary syndrome; a 20 years retrospective cohort study. Clin Endocrinol (Oxf) 2012; In press http://dx.doi.org/10.1111/cen.12068; PMID: 23046078
  • Fiehn C. Methotrexate transport mechanisms: the basis for targeted drug delivery and ß-folate-receptor-specific treatment. Clin Exp Rheumatol 2010; 28:Suppl 61 S40 - 5; PMID: 21044432
  • Ahlehoff O, Skov L, Gislason G, Lindhardsen J, Kristensen SL, Iversen L, et al. Cardiovascular disease event rates in patients with severe psoriasis treated with systemic anti-inflammatory drugs: a Danish real-world cohort study. J Intern Med 2013; 273:197 - 204; PMID: 22963528
  • Ridker PM. Testing the inflammatory hypothesis of atherothrombosis: scientific rationale for the cardiovascular inflammation reduction trial (CIRT). J Thromb Haemost 2009; 7:Suppl 1 332 - 9; http://dx.doi.org/10.1111/j.1538-7836.2009.03404.x; PMID: 19630828
  • Russo GT, Minutoli L, Bitto A, Altavilla D, Alessi E, Giandalia A, et al. Methotrexate Increases Skeletal Muscle GLUT4 Expression and Improves Metabolic Control in Experimental Diabetes. J Nutr Metab 2012; 2012:132056; http://dx.doi.org/10.1155/2012/132056; PMID: 22778921
  • Zolk O. Disposition of metformin: variability due to polymorphisms of organic cation transporters. Ann Med 2012; 44:119 - 29; http://dx.doi.org/10.3109/07853890.2010.549144; PMID: 21366511
  • Zhou M, Xia L, Wang J. Metformin transport by a newly cloned proton-stimulated organic cation transporter (plasma membrane monoamine transporter) expressed in human intestine. Drug Metab Dispos 2007; 35:1956 - 62; http://dx.doi.org/10.1124/dmd.107.015495; PMID: 17600084
  • Damme K, Nies AT, Schaeffeler E, Schwab M. Mammalian MATE (SLC47A) transport proteins: impact on efflux of endogenous substrates and xenobiotics. Drug Metab Rev 2011; 43:499 - 523; http://dx.doi.org/10.3109/03602532.2011.602687; PMID: 21923552
  • Dervieux T, Zablocki R, Kremer J. Red blood cell methotrexate polyglutamates emerge as a function of dosage intensity and route of administration during pulse methotrexate therapy in rheumatoid arthritis. Rheumatology (Oxford) 2010; 49:2337 - 45; http://dx.doi.org/10.1093/rheumatology/keq216; PMID: 20713496
  • van Roon EN, van de Laar MA. Methotrexate bioavailability. Clin Exp Rheumatol 2010; 28:Suppl 61 S27 - 32; PMID: 21044430
  • Chae JW, Baek IH, Lee BY, Cho SK, Kwon KI. Population PK/PD analysis of metformin using the signal transduction model. Br J Clin Pharmacol 2012; 74:815 - 23; http://dx.doi.org/10.1111/j.1365-2125.2012.04260.x; PMID: 22380769
  • Robert F, Fendri S, Hary L, Lacroix C, Andréjak M, Lalau JD. Kinetics of plasma and erythrocyte metformin after acute administration in healthy subjects. Diabetes Metab 2003; 29:279 - 83; http://dx.doi.org/10.1016/S1262-3636(07)70037-X; PMID: 12909816
  • Sambol NC, Chiang J, O’Conner M, Liu CY, Lin ET, Goodman AM, et al. Pharmacokinetics and pharmacodynamics of metformin in healthy subjects and patients with noninsulin-dependent diabetes mellitus. J Clin Pharmacol 1996; 36:1012 - 21; http://dx.doi.org/10.1177/009127009603601105; PMID: 8973990
  • Stepensky D, Friedman M, Raz I, Hoffman A. Pharmacokinetic-pharmacodynamic analysis of the glucose-lowering effect of metformin in diabetic rats reveals first-pass pharmacodynamic effect. Drug Metab Dispos 2002; 30:861 - 8; http://dx.doi.org/10.1124/dmd.30.8.861; PMID: 12124302
  • Owen MR, Doran E, Halestrap AP. Evidence that metformin exerts its anti-diabetic effects through inhibition of complex 1 of the mitochondrial respiratory chain. Biochem J 2000; 348:607 - 14; http://dx.doi.org/10.1042/0264-6021:3480607; PMID: 10839993
  • Sogame Y, Kitamura A, Yabuki M, Komuro S. A comparison of uptake of metformin and phenformin mediated by hOCT1 in human hepatocytes. Biopharm Drug Dispos 2009; 30:476 - 84; http://dx.doi.org/10.1002/bdd.684; PMID: 19768675
  • Davidoff F, Haas D. Phenethylbiguanide effect on mitochondrial Ca+ and Mg2+ content. Biochem Pharmacol 1979; 28:3457 - 63; http://dx.doi.org/10.1016/0006-2952(79)90086-8; PMID: 394746
  • Chan ESL, Cronstein BN. Methotrexate--how does it really work?. Nat Rev Rheumatol 2010; 6:175 - 8; http://dx.doi.org/10.1038/nrrheum.2010.5; PMID: 20197777
  • Baggott JE, Morgan SL, Sams WM, Linden J. Urinary adenosine and aminoimidazolecarboxamide excretion in methotrexate-treated patients with psoriasis. Arch Dermatol 1999; 135:813 - 7; http://dx.doi.org/10.1001/archderm.135.7.813; PMID: 10411156
  • Hiratsuka C, Fukuwatari T, Shibata K. Fate of dietary tryptophan in young Japanese women. Int J Tryptophan Res 2012; 5:33 - 47; PMID: 23150724
  • Huo T, Cai S, Lu X, Sha Y, Yu M, Li F. Metabonomic study of biochemical changes in the serum of type 2 diabetes mellitus patients after the treatment of metformin hydrochloride. J Pharm Biomed Anal 2009; 49:976 - 82; http://dx.doi.org/10.1016/j.jpba.2009.01.008; PMID: 19249171
  • Marchetti P, Masiello P, Benzi L, Cecchetti P, Fierabracci V, Giannarelli R, et al. Effects of metformin therapy on plasma amino acid pattern in patients with maturity-onset diabetes. Drugs Exp Clin Res 1989; 15:565 - 70; PMID: 2637797
  • Song IS, Lee Y, Shin M-H, Kim H, Ahn YG, Park I, et al. Pharmacogenetics meets metabolomics: discovery of tryptophan as a new endogenous OCT2 substrate related to metformin disposition. PLoS One 2012; 7:e36637; http://dx.doi.org/10.1371/journal.pone.0036637; PMID: 22590580
  • Shen S, O’Brien T, Yap LM, Prince HM, McCormack CJ. The use of methotrexate in dermatology: a review. Australas J Dermatol 2012; 53:1 - 18; http://dx.doi.org/10.1111/j.1440-0960.2011.00839.x; PMID: 22309324
  • Greasley SE, Horton P, Ramcharan J, Beardsley GP, Benkovic SJ, Wilson IA. Crystal structure of a bifunctional transformylase and cyclohydrolase enzyme in purine biosynthesis. Nat Struct Biol 2001; 8:402 - 6; http://dx.doi.org/10.1038/87555; PMID: 11323713
  • Tibbetts AS, Appling DR. Compartmentalization of Mammalian folate-mediated one-carbon metabolism. Annu Rev Nutr 2010; 30:57 - 81; http://dx.doi.org/10.1146/annurev.nutr.012809.104810; PMID: 20645850
  • Baggott JE, Tamura T. Evidence for the hypothesis that 10-formyldihydrofolate is the in vivo substrate for aminoimidazolecarboxamide ribotide transformylase. Exp Biol Med (Maywood) 2010; 235:271 - 7; http://dx.doi.org/10.1258/ebm.2009.009151; PMID: 20404044
  • Andersson SE, Johansson LH, Lexmüller K, Ekström GM. Anti-arthritic effect of methotrexate: is it really mediated by adenosine?. Eur J Pharm Sci 2000; 9:333 - 43; http://dx.doi.org/10.1016/S0928-0987(99)00073-1; PMID: 10664473
  • Phillips DC, Woollard KJ, Griffiths HR. The anti-inflammatory actions of methotrexate are critically dependent upon the production of reactive oxygen species. Br J Pharmacol 2003; 138:501 - 11; http://dx.doi.org/10.1038/sj.bjp.0705054; PMID: 12569075
  • Morgan SL, Oster RA, Lee JY, Alarcón GS, Baggott JE. The effect of folic acid and folinic acid supplements on purine metabolism in methotrexate-treated rheumatoid arthritis. Arthritis Rheum 2004; 50:3104 - 11; http://dx.doi.org/10.1002/art.20516; PMID: 15476202
  • Beckers A, Organe S, Timmermans L, Vanderhoydonc F, Deboel L, Derua R, et al. Methotrexate enhances the antianabolic and antiproliferative effects of 5-aminoimidazole-4-carboxamide riboside. Mol Cancer Ther 2006; 5:2211 - 7; http://dx.doi.org/10.1158/1535-7163.MCT-06-0001; PMID: 16985054
  • Racanelli AC, Rothbart SB, Heyer CL, Moran RG. Therapeutics by cytotoxic metabolite accumulation: pemetrexed causes ZMP accumulation, AMPK activation, and mammalian target of rapamycin inhibition. Cancer Res 2009; 69:5467 - 74; http://dx.doi.org/10.1158/0008-5472.CAN-08-4979; PMID: 19549896
  • Rothbart SB, Racanelli AC, Moran RG. Pemetrexed indirectly activates the metabolic kinase AMPK in human carcinomas. Cancer Res 2010; 70:10299 - 309; http://dx.doi.org/10.1158/0008-5472.CAN-10-1873; PMID: 21159649
  • Kuznetsov JN, Leclerc GJ, Leclerc GM, Barredo JC. AMPK and Akt determine apoptotic cell death following perturbations of one-carbon metabolism by regulating ER stress in acute lymphoblastic leukemia. Mol Cancer Ther 2011; 10:437 - 47; http://dx.doi.org/10.1158/1535-7163.MCT-10-0777; PMID: 21262957
  • Yao SYM, Ng AML, Cass CE, Baldwin SA, Young JD. Nucleobase transport by human equilibrative nucleoside transporter 1 (hENT1). J Biol Chem 2011; 286:32552 - 62; http://dx.doi.org/10.1074/jbc.M111.236117; PMID: 21795683
  • Griffiths HR, Dunston CR, Bennett SJ, Grant MM, Phillips DC, Kitas GD. Free radicals and redox signalling in T-cells during chronic inflammation and ageing. Biochem Soc Trans 2011; 39:1273 - 8; http://dx.doi.org/10.1042/BST0391273; PMID: 21936801
  • Gregori-Puigjané E, Setola V, Hert J, Crews BA, Irwin JJ, Lounkine E, et al. Identifying mechanism-of-action targets for drugs and probes. Proc Natl Acad Sci U S A 2012; 109:11178 - 83; http://dx.doi.org/10.1073/pnas.1204524109; PMID: 22711801
  • Cuthbertson DJ, Babraj JA, Mustard KJW, Towler MC, Green KA, Wackerhage H, et al. 5-aminoimidazole-4-carboxamide 1-beta-D-ribofuranoside acutely stimulates skeletal muscle 2-deoxyglucose uptake in healthy men. Diabetes 2007; 56:2078 - 84; http://dx.doi.org/10.2337/db06-1716; PMID: 17513706
  • Boon H, Bosselaar M, Praet SFE, Blaak EE, Saris WHM, Wagenmakers AJM, et al. Intravenous AICAR administration reduces hepatic glucose output and inhibits whole body lipolysis in type 2 diabetic patients. Diabetologia 2008; 51:1893 - 900; http://dx.doi.org/10.1007/s00125-008-1108-7; PMID: 18709353
  • Babraj JA, Mustard K, Sutherland C, Towler MC, Chen S, Smith K, et al. Blunting of AICAR-induced human skeletal muscle glucose uptake in type 2 diabetes is dependent on age rather than diabetic status. Am J Physiol Endocrinol Metab 2009; 296:E1042 - 8; http://dx.doi.org/10.1152/ajpendo.90811.2008; PMID: 19190259
  • Paglia G, Hrafnsdóttir S, Magnúsdóttir M, Fleming RMT, Thorlacius S, Palsson BØ, et al. Monitoring metabolites consumption and secretion in cultured cells using ultra-performance liquid chromatography quadrupole-time of flight mass spectrometry (UPLC-Q-ToF-MS). Anal Bioanal Chem 2012; 402:1183 - 98; http://dx.doi.org/10.1007/s00216-011-5556-4; PMID: 22159369
  • Pressman BC. The Effects Energy of Guanidine and Alkylguanidines Transfer Reactions of Mitochondria. J Biol Chem 1963; 238:401 - 9
  • Davidoff F. Effects of guanidine derivatives on mitochondrial function. Ca2+ uptake and release. J Biol Chem 1974; 249:6406 - 15; PMID: 4418697
  • Davidoff F. Effects of guanidine derivatives on mitochondrial function. II. Reversal of guanidine-derivative inhibiton by free fatty acids. J Clin Invest 1968; 47:2344 - 58; http://dx.doi.org/10.1172/JCI105919; PMID: 5676528
  • El-Mir MY, Nogueira V, Fontaine E, Avéret N, Rigoulet M, Leverve X. Dimethylbiguanide inhibits cell respiration via an indirect effect targeted on the respiratory chain complex I. J Biol Chem 2000; 275:223 - 8; http://dx.doi.org/10.1074/jbc.275.1.223; PMID: 10617608
  • Viollet B, Guigas B, Sanz Garcia N, Leclerc J, Foretz M, Andreelli F. Cellular and molecular mechanisms of metformin: an overview. Clin Sci (Lond) 2012; 122:253 - 70; http://dx.doi.org/10.1042/CS20110386; PMID: 22117616
  • Larsen S, Rabøl R, Hansen CN, Madsbad S, Helge JW, Dela F. Metformin-treated patients with type 2 diabetes have normal mitochondrial complex I respiration. Diabetologia 2012; 55:443 - 9; http://dx.doi.org/10.1007/s00125-011-2340-0; PMID: 22009334
  • Guigas B, Detaille D, Chauvin C, Batandier C, De Oliveira F, Fontaine E, et al. Metformin inhibits mitochondrial permeability transition and cell death: a pharmacological in vitro study. Biochem J 2004; 382:877 - 84; http://dx.doi.org/10.1042/BJ20040885; PMID: 15175014
  • Li B, Chauvin C, De Paulis D, De Oliveira F, Gharib A, Vial G, et al. Inhibition of complex I regulates the mitochondrial permeability transition through a phosphate-sensitive inhibitory site masked by cyclophilin D. Biochim Biophys Acta 2012; 1817:1628 - 34; http://dx.doi.org/10.1016/j.bbabio.2012.05.011; PMID: 22659400
  • Batandier C, Guigas B, Detaille D, El-Mir MY, Fontaine E, Rigoulet M, et al. The ROS production induced by a reverse-electron flux at respiratory-chain complex 1 is hampered by metformin. J Bioenerg Biomembr 2006; 38:33 - 42; http://dx.doi.org/10.1007/s10863-006-9003-8; PMID: 16732470
  • Kaminski MM, Sauer SW, Klemke CD, Süss D, Okun JG, Krammer PH, et al. Mitochondrial reactive oxygen species control T cell activation by regulating IL-2 and IL-4 expression: mechanism of ciprofloxacin-mediated immunosuppression. J Immunol 2010; 184:4827 - 41; http://dx.doi.org/10.4049/jimmunol.0901662; PMID: 20335530
  • Mihaylova MM, Shaw RJ. The AMPK signalling pathway coordinates cell growth, autophagy and metabolism. Nat Cell Biol 2011; 13:1016 - 23; http://dx.doi.org/10.1038/ncb2329; PMID: 21892142
  • Sweeney D, Raymer ML, Lockwood TD. Antidiabetic and antimalarial biguanide drugs are metal-interactive antiproteolytic agents. Biochem Pharmacol 2003; 66:663 - 77; http://dx.doi.org/10.1016/S0006-2952(03)00338-1; PMID: 12906931
  • Lockwood TD. The lysosome among targets of metformin: new anti-inflammatory uses for an old drug?. Expert Opin Ther Targets 2010; 14:467 - 78; http://dx.doi.org/10.1517/14728221003774135; PMID: 20392164
  • Lu LP, Yang P, Qin SD, Zhu ML. Bis[1,1-dimethylbiguanide(1-)-kappa2N2,N5]copper(II) monohydrate. Acta Crystallogr C 2004; 60:m219 - 20; http://dx.doi.org/10.1107/S0108270104006729; PMID: 15131370
  • Logie L, Harthill J, Patel K, Bacon S, Hamilton DL, Macrae K, et al. Cellular responses to the metal-binding properties of metformin. Diabetes 2012; 61:1423 - 33; http://dx.doi.org/10.2337/db11-0961; PMID: 22492524
  • Rena G, Pearson ER, Sakamoto K. Molecular action and pharmacogenetics of metformin: current understanding of an old drug. Diabetes Management 2012; 2:439 - 52; http://dx.doi.org/10.2217/dmt.12.42
  • Lu L, Gao X, Zhu M, Wang S, Wu Q, Xing S, et al. Exploration of biguanido-oxovanadium complexes as potent and selective inhibitors of protein tyrosine phosphatases. Biometals 2012; 25:599 - 610; http://dx.doi.org/10.1007/s10534-012-9548-4; PMID: 22547055
  • Holland W, Morrison T, Chang Y, Wiernsperger N, Stith BJ. Metformin (Glucophage) inhibits tyrosine phosphatase activity to stimulate the insulin receptor tyrosine kinase. Biochem Pharmacol 2004; 67:2081 - 91; http://dx.doi.org/10.1016/j.bcp.2004.02.016; PMID: 15135305
  • Hardie DG, Ross FA, Hawley SA. AMPK: a nutrient and energy sensor that maintains energy homeostasis. Nat Rev Mol Cell Biol 2012; 13:251 - 62; http://dx.doi.org/10.1038/nrm3311; PMID: 22436748
  • Zhou G, Myers R, Li Y, Chen Y, Shen X, Fenyk-Melody J, et al. Role of AMP-activated protein kinase in mechanism of metformin action. J Clin Invest 2001; 108:1167 - 74; PMID: 11602624
  • Musi N, Hirshman MF, Nygren J, Svanfeldt M, Bavenholm P, Rooyackers O, et al. Metformin increases AMP-activated protein kinase activity in skeletal muscle of subjects with type 2 diabetes. Diabetes 2002; 51:2074 - 81; http://dx.doi.org/10.2337/diabetes.51.7.2074; PMID: 12086935
  • Boyle JG, Logan PJ, Jones GC, Small M, Sattar N, Connell JMC, et al. AMP-activated protein kinase is activated in adipose tissue of individuals with type 2 diabetes treated with metformin: a randomised glycaemia-controlled crossover study. Diabetologia 2011; 54:1799 - 809; http://dx.doi.org/10.1007/s00125-011-2126-4; PMID: 21455728
  • Hawley SA, Fullerton MD, Ross FA, Schertzer JD, Chevtzoff C, Walker KJ, et al. The ancient drug salicylate directly activates AMP-activated protein kinase. Science 2012; 336:918 - 22; http://dx.doi.org/10.1126/science.1215327; PMID: 22517326
  • Zhang Y, Wang Y, Bao C, Xu Y, Shen H, Chen J, et al. Metformin interacts with AMPK through binding to γ subunit. Mol Cell Biochem 2012; 368:69 - 76; http://dx.doi.org/10.1007/s11010-012-1344-5; PMID: 22644486
  • Cool B, Zinker B, Chiou W, Kifle L, Cao N, Perham M, et al. Identification and characterization of a small molecule AMPK activator that treats key components of type 2 diabetes and the metabolic syndrome. Cell Metab 2006; 3:403 - 16; http://dx.doi.org/10.1016/j.cmet.2006.05.005; PMID: 16753576
  • Lee KH, Hsu EC, Guh JH, Yang HC, Wang D, Kulp SK, et al. Targeting energy metabolic and oncogenic signaling pathways in triple-negative breast cancer by a novel adenosine monophosphate-activated protein kinase (AMPK) activator. J Biol Chem 2011; 286:39247 - 58; http://dx.doi.org/10.1074/jbc.M111.264598; PMID: 21917926
  • Ouyang J, Parakhia RA, Ochs RS. Metformin activates AMP kinase through inhibition of AMP deaminase. J Biol Chem 2011; 286:1 - 11; http://dx.doi.org/10.1074/jbc.M110.121806; PMID: 21059655
  • Magnuson B, Ekim B, Fingar DC. Regulation and function of ribosomal protein S6 kinase (S6K) within mTOR signalling networks. Biochem J 2012; 441:1 - 21; http://dx.doi.org/10.1042/BJ20110892; PMID: 22168436
  • Cantó C, Auwerx J. Targeting sirtuin 1 to improve metabolism: all you need is NAD(+)?. Pharmacol Rev 2012; 64:166 - 87; http://dx.doi.org/10.1124/pr.110.003905; PMID: 22106091
  • Eigenbrodt E, Glossmann H. Glycolysis–one of the keys to cancer?. Trends Pharmacol Sci 1980; 1:240 - 5; http://dx.doi.org/10.1016/0165-6147(80)90009-7
  • Kamiński MM, Sauer SW, Kamiński M, Opp S, Ruppert T, Grigaravičius P, et al. T cell activation is driven by an ADP-dependent glucokinase linking enhanced glycolysis with mitochondrial reactive oxygen species generation. Cell Rep 2012; 2:1300 - 15; http://dx.doi.org/10.1016/j.celrep.2012.10.009; PMID: 23168256
  • Kalender A, Selvaraj A, Kim SY, Gulati P, Brûlé S, Viollet B, et al. Metformin, independent of AMPK, inhibits mTORC1 in a rag GTPase-dependent manner. Cell Metab 2010; 11:390 - 401; http://dx.doi.org/10.1016/j.cmet.2010.03.014; PMID: 20444419
  • Kickstein E, Krauss S, Thornhill P, Rutschow D, Zeller R, Sharkey J, et al. Biguanide metformin acts on tau phosphorylation via mTOR/protein phosphatase 2A (PP2A) signaling. Proc Natl Acad Sci U S A 2010; 107:21830 - 5; http://dx.doi.org/10.1073/pnas.0912793107; PMID: 21098287
  • Ben Sahra I, Regazzetti C, Robert G, Laurent K, Le Marchand-Brustel Y, Auberger P, et al. Metformin, independent of AMPK, induces mTOR inhibition and cell-cycle arrest through REDD1. Cancer Res 2011; 71:4366 - 72; http://dx.doi.org/10.1158/0008-5472.CAN-10-1769; PMID: 21540236
  • Turban S, Stretton C, Drouin O, Green CJ, Watson ML, Gray A, et al. Defining the contribution of AMP-activated protein kinase (AMPK) and protein kinase C (PKC) in regulation of glucose uptake by metformin in skeletal muscle cells. J Biol Chem 2012; 287:20088 - 99; http://dx.doi.org/10.1074/jbc.M111.330746; PMID: 22511782
  • Refsum H, Helland S, Ueland PM. Fasting plasma homocysteine as a sensitive parameter of antifolate effect: a study of psoriasis patients receiving low-dose methotrexate treatment. Clin Pharmacol Ther 1989; 46:510 - 20; http://dx.doi.org/10.1038/clpt.1989.179; PMID: 2582708
  • Hroch M, Chladek J, Simkova M, Vaneckova J, Grim J, Martinkova J. A pilot study of pharmacokinetically guided dosing of oral methotrexate in the initial phase of psoriasis treatment. J Eur Acad Dermatol Venereol 2008; 22:19 - 24; PMID: 18031504
  • Reinstatler L, Qi YP, Williamson RS, Garn JV, Oakley GP Jr.. Association of biochemical B₁₂ deficiency with metformin therapy and vitamin B₁₂ supplements: the National Health and Nutrition Examination Survey, 1999-2006. Diabetes Care 2012; 35:327 - 33; http://dx.doi.org/10.2337/dc11-1582; PMID: 22179958
  • Wulffelé MG, Kooy A, Lehert P, Bets D, Ogterop JC, Borger van der Burg B, et al. Effects of short-term treatment with metformin on serum concentrations of homocysteine, folate and vitamin B12 in type 2 diabetes mellitus: a randomized, placebo-controlled trial. J Intern Med 2003; 254:455 - 63; http://dx.doi.org/10.1046/j.1365-2796.2003.01213.x; PMID: 14535967
  • de Jager J, Kooy A, Lehert P, Wulffelé MG, van der Kolk J, Bets D, et al. Long term treatment with metformin in patients with type 2 diabetes and risk of vitamin B-12 deficiency: randomised placebo controlled trial. BMJ 2010; 340:c2181; http://dx.doi.org/10.1136/bmj.c2181; PMID: 20488910
  • Aarsand AK, Carlsen SM. Folate administration reduces circulating homocysteine levels in NIDDM patients on long-term metformin treatment. J Intern Med 1998; 244:169 - 74; http://dx.doi.org/10.1046/j.1365-2796.1998.00361.x; PMID: 10095804
  • Carlsen SM, Følling I, Grill V, Bjerve KS, Schneede J, Refsum H. Metformin increases total serum homocysteine levels in non-diabetic male patients with coronary heart disease. Scand J Clin Lab Invest 1997; 57:521 - 7; http://dx.doi.org/10.3109/00365519709084603; PMID: 9350072
  • Palomba S, Falbo A, Giallauria F, Russo T, Tolino A, Zullo F, et al. Effects of metformin with or without supplementation with folate on homocysteine levels and vascular endothelium of women with polycystic ovary syndrome. Diabetes Care 2010; 33:246 - 51; http://dx.doi.org/10.2337/dc09-1516; PMID: 19933994
  • Corominas-Faja B, Quirantes-Piné R, Oliveras-Ferraros C, Vazquez-Martin A, Cufí S, Martin-Castillo B, et al. Metabolomic fingerprint reveals that metformin impairs one-carbon metabolism in a manner similar to the antifolate class of chemotherapy drugs. Aging (Albany NY) 2012; 4:480 - 98; PMID: 22837425
  • Zakikhani M, Bazile M, Hashemi S, Javeshghani S, Avizonis D, St Pierre J, et al. Alterations in cellular energy metabolism associated with the antiproliferative effects of the ATM inhibitor KU-55933 and with metformin. PLoS One 2012; 7:e49513; http://dx.doi.org/10.1371/journal.pone.0049513; PMID: 23185347
  • Nath N, Khan M, Paintlia MK, Singh I, Hoda MN, Giri S. Metformin attenuated the autoimmune disease of the central nervous system in animal models of multiple sclerosis. J Immunol 2009; 182:8005 - 14; http://dx.doi.org/10.4049/jimmunol.0803563; PMID: 19494326
  • Bai A, Ma AG, Yong M, Weiss CR, Ma Y, Guan Q, et al. AMPK agonist downregulates innate and adaptive immune responses in TNBS-induced murine acute and relapsing colitis. Biochem Pharmacol 2010; 80:1708 - 17; http://dx.doi.org/10.1016/j.bcp.2010.08.009; PMID: 20797389
  • Park CS, Bang BR, Kwon HS, Moon KA, Kim TB, Lee KY, et al. Metformin reduces airway inflammation and remodeling via activation of AMP-activated protein kinase. Biochem Pharmacol 2012; 84:1660 - 70; http://dx.doi.org/10.1016/j.bcp.2012.09.025; PMID: 23041647
  • Sag D, Carling D, Stout RD, Suttles J. Adenosine 5′-monophosphate-activated protein kinase promotes macrophage polarization to an anti-inflammatory functional phenotype. J Immunol 2008; 181:8633 - 41; PMID: 19050283
  • Krawczyk CM, Holowka T, Sun J, Blagih J, Amiel E, DeBerardinis RJ, et al. Toll-like receptor-induced changes in glycolytic metabolism regulate dendritic cell activation. Blood 2010; 115:4742 - 9; http://dx.doi.org/10.1182/blood-2009-10-249540; PMID: 20351312
  • Krysiak R, Okopien B. Lymphocyte-suppressing and systemic anti-inflammatory effects of high-dose metformin in simvastatin-treated patients with impaired fasting glucose. Atherosclerosis 2012; 225:403 - 7; http://dx.doi.org/10.1016/j.atherosclerosis.2012.09.034; PMID: 23107042
  • Engler JR, Zannettino ACW, Bailey CG, Rasko JEJ, Hughes TP, White DL. OCT-1 function varies with cell lineage but is not influenced by BCR-ABL. Haematologica 2011; 96:213 - 20; http://dx.doi.org/10.3324/haematol.2010.033290; PMID: 20971815
  • Arai M, Uchiba M, Komura H, Mizuochi Y, Harada N, Okajima K. Metformin, an antidiabetic agent, suppresses the production of tumor necrosis factor and tissue factor by inhibiting early growth response factor-1 expression in human monocytes in vitro. J Pharmacol Exp Ther 2010; 334:206 - 13; http://dx.doi.org/10.1124/jpet.109.164970; PMID: 20371705
  • Salt IP, Palmer TM. Exploiting the anti-inflammatory effects of AMP-activated protein kinase activation. Expert Opin Investig Drugs 2012; 21:1155 - 67; PMID: 22694351
  • Charles MA, Morange P, Eschwège E, André P, Vague P, Juhan-Vague I. Effect of weight change and metformin on fibrinolysis and the von Willebrand factor in obese nondiabetic subjects: the BIGPRO1 Study. Biguanides and the Prevention of the Risk of Obesity. Diabetes Care 1998; 21:1967 - 72; http://dx.doi.org/10.2337/diacare.21.11.1967; PMID: 9802752
  • Chakrabarti R. Hockin, Elisabeth D, Fearnley GR. Fibrinolytic effect of metformin in coronary-artery disease. Lancet 1965; 286:256 - 9; http://dx.doi.org/10.1016/S0140-6736(65)92383-4
  • Fontbonne A, Diouf I, Baccara-Dinet M, Eschwege E, Charles MA. Effects of 1-year treatment with metformin on metabolic and cardiovascular risk factors in non-diabetic upper-body obese subjects with mild glucose anomalies: a post-hoc analysis of the BIGPRO1 trial. Diabetes Metab 2009; 35:385 - 91; http://dx.doi.org/10.1016/j.diabet.2009.03.005; PMID: 19665415
  • Lund SS, Tarnow L, Stehouwer CDA, Schalkwijk CG, Teerlink T, Gram J, et al. Impact of metformin versus repaglinide on non-glycaemic cardiovascular risk markers related to inflammation and endothelial dysfunction in non-obese patients with type 2 diabetes. Eur J Endocrinol 2008; 158:631 - 41; http://dx.doi.org/10.1530/EJE-07-0815; PMID: 18426821
  • Luque-Ramírez M, Escobar-Morreale HF. Treatment of polycystic ovary syndrome (PCOS) with metformin ameliorates insulin resistance in parallel with the decrease of serum interleukin-6 concentrations. Horm Metab Res 2010; 42:815 - 20; http://dx.doi.org/10.1055/s-0030-1262855; PMID: 20730705
  • Lin YS, Tsai SJ, Lin MW, Yang CT, Huang MF, Wu MH. Interleukin-6 as an early chronic inflammatory marker in polycystic ovary syndrome with insulin receptor substrate-2 polymorphism. Am J Reprod Immunol 2011; 66:527 - 33; http://dx.doi.org/10.1111/j.1600-0897.2011.01059.x; PMID: 21801267
  • Diamanti-Kandarakis E, Paterakis T, Alexandraki K, Piperi C, Aessopos A, Katsikis I, et al. Indices of low-grade chronic inflammation in polycystic ovary syndrome and the beneficial effect of metformin. Hum Reprod 2006; 21:1426 - 31; http://dx.doi.org/10.1093/humrep/del003; PMID: 16497699
  • Banaszewska B, Pawelczyk L, Spaczynski RZ, Duleba AJ. Comparison of simvastatin and metformin in treatment of polycystic ovary syndrome: prospective randomized trial. J Clin Endocrinol Metab 2009; 94:4938 - 45; http://dx.doi.org/10.1210/jc.2009-1674; PMID: 19890022
  • Perera GK, Di Meglio P, Nestle FO. Psoriasis. Ann Rev Pathol 2012; 7:385 - 422; http://dx.doi.org/10.1146/annurev-pathol-011811-132448
  • Davidovici BB, Sattar N, Prinz JC, Puig L, Emery P, Barker JN, et al. Psoriasis and systemic inflammatory diseases: potential mechanistic links between skin disease and co-morbid conditions. J Invest Dermatol 2010; 130:1785 - 96; http://dx.doi.org/10.1038/jid.2010.103; PMID: 20445552
  • Langan SM, Seminara NM, Shin DB, Troxel AB, Kimmel SE, Mehta NN, et al. Prevalence of metabolic syndrome in patients with psoriasis: a population-based study in the United Kingdom. J Invest Dermatol 2012; 132:556 - 62; http://dx.doi.org/10.1038/jid.2011.365; PMID: 22113483
  • Gisondi P, Tessari G, Conti A, Piaserico S, Schianchi S, Peserico A, et al. Prevalence of metabolic syndrome in patients with psoriasis: a hospital-based case-control study. Br J Dermatol 2007; 157:68 - 73; http://dx.doi.org/10.1111/j.1365-2133.2007.07986.x; PMID: 17553036
  • Love TJ, Qureshi AA, Karlson EW, Gelfand JM, Choi HK. Prevalence of the metabolic syndrome in psoriasis: results from the National Health and Nutrition Examination Survey, 2003-2006. Arch Dermatol 2011; 147:419 - 24; http://dx.doi.org/10.1001/archdermatol.2010.370; PMID: 21173301
  • Ford ES, Zhao G, Tsai J, Li C. Vitamin D and all-cause mortality among adults in USA: findings from the National Health and Nutrition Examination Survey Linked Mortality Study. Int J Epidemiol 2011; 40:998 - 1005; http://dx.doi.org/10.1093/ije/dyq264; PMID: 21266455
  • Azfar RS, Seminara NM, Shin DB, Troxel AB, Margolis DJ, Gelfand JM. Increased risk of diabetes mellitus and likelihood of receiving diabetes mellitus treatment in patients with psoriasis. Arch Dermatol 2012; 148:995 - 1000; http://dx.doi.org/10.1001/archdermatol.2012.1401; PMID: 22710320
  • Ahlehoff O. Psoriasis and Cardiovascular Disease: epidemiological studies. Dan Med Bull 2011; 58:B4347; PMID: 22047936
  • Suárez-Fariñas M, Li K, Fuentes-Duculan J, Hayden K, Brodmerkel C, Krueger JG. Expanding the psoriasis disease profile: interrogation of the skin and serum of patients with moderate-to-severe psoriasis. J Invest Dermatol 2012; 132:2552 - 64; http://dx.doi.org/10.1038/jid.2012.184; PMID: 22763790
  • Tian S, Krueger JG, Li K, Jabbari A, Brodmerkel C, Lowes MA, et al. Meta-analysis derived (MAD) transcriptome of psoriasis defines the “core” pathogenesis of disease. PLoS One 2012; 7:e44274; http://dx.doi.org/10.1371/journal.pone.0044274; PMID: 22957057
  • Farias MM, Achurra P, Boza C, Vega A, de la Cruz C. Psoriasis following bariatric surgery: clinical evolution and impact on quality of life on 10 patients. Obes Surg 2012; 22:877 - 80; http://dx.doi.org/10.1007/s11695-012-0646-8; PMID: 22488682
  • Gisondi P, Del Giglio M, Di Francesco V, Zamboni M, Girolomoni G. Weight loss improves the response of obese patients with moderate-to-severe chronic plaque psoriasis to low-dose cyclosporine therapy: a randomized, controlled, investigator-blinded clinical trial. Am J Clin Nutr 2008; 88:1242 - 7; PMID: 18996858
  • Phung OJ, Sood NA, Sill BE, Coleman CI. Oral anti-diabetic drugs for the prevention of Type 2 diabetes. Diabet Med 2011; 28:948 - 64; http://dx.doi.org/10.1111/j.1464-5491.2011.03303.x; PMID: 21429006
  • Phung OJ, Baker WL, Tongbram V, Bhardwaj A, Coleman CI. Oral antidiabetic drugs and regression from prediabetes to normoglycemia: a meta-analysis. Ann Pharmacother 2012; 46:469 - 76; http://dx.doi.org/10.1345/aph.1Q554; PMID: 22474136
  • Knowler WC, Barrett-Connor E, Fowler SE, Hamman RF, Lachin JM, Walker EA, et al, Diabetes Prevention Program Research Group. Reduction in the incidence of type 2 diabetes with lifestyle intervention or metformin. N Engl J Med 2002; 346:393 - 403; http://dx.doi.org/10.1056/NEJMoa012512; PMID: 11832527
  • Diabetes Prevention Program Research Group. The 10-year cost-effectiveness of lifestyle intervention or metformin for diabetes prevention: an intent-to-treat analysis of the DPP/DPPOS. Diabetes Care 2012; 35:723 - 30; http://dx.doi.org/10.2337/dc11-1468; PMID: 22442395
  • Orchard TJ, Temprosa M, Goldberg R, Haffner S, Ratner R, Marcovina S, et al, Diabetes Prevention Program Research Group. The effect of metformin and intensive lifestyle intervention on the metabolic syndrome: the Diabetes Prevention Program randomized trial. Ann Intern Med 2005; 142:611 - 9; PMID: 15838067
  • Diabetes Prevention Program Research Group. Long-term safety, tolerability, and weight loss associated with metformin in the Diabetes Prevention Program Outcomes Study. Diabetes Care 2012; 35:731 - 7; http://dx.doi.org/10.2337/dc11-1299; PMID: 22442396
  • Suzuki J, Yoshimura T, Simeonova M, Takeuchi K, Murakami Y, Morizane Y, et al. Aminoimidazole carboxamide ribonucleotide ameliorates experimental autoimmune uveitis. Invest Ophthalmol Vis Sci 2012; 53:4158 - 69; http://dx.doi.org/10.1167/iovs.11-9323; PMID: 22562518
  • Pollak M. Investigating Metformin for Cancer Prevention and Treatment:The End of the Beginning. Cancer Discovery 2012; 2:778 - 90; PMID: 22926251
  • Emami Riedmaier A, Fisel P, Niles AT, Schaeffeler E, Schwab M. Metformin and cancer:from the old medicine cabinet to pharmacological pitfalls and prospects. Trends Pharmacol Sci 2013; 34:126 - 35; PMID: 23277337