Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

PRIMER

Gout
Nicola Dalbeth1*, Hyon K. Choi2, Leo A. B. Joosten3, Puja P. Khanna4, Hirotaka Matsuo5,
Fernando Perez-​Ruiz6 and Lisa K. Stamp7
Abstract | Gout is a chronic disease caused by monosodium urate (MSU) crystal deposition.
Gout typically presents as an acute, self-​limiting inflammatory monoarthritis that affects the
joints of the lower limb. Elevated serum urate level (hyperuricaemia) is the major risk factor for
MSU crystal deposition and development of gout. Although traditionally considered a disorder
of purine metabolism, altered urate transport, both in the gut and the kidneys, has a key role in
the pathogenesis of hyperuricaemia. Anti-​inflammatory agents, such corticosteroids, NSAIDs
and colchicine, are widely used for the treatment of gout flare; recognition of the importance of
NLRP3 inflammasome activation and bioactive IL-1β release in initiation of the gout flare has
led to the development of anti-​IL-1β biological therapy for gout flares. Sustained reduction in
serum urate levels using urate-​lowering therapy is vital in the long-​term management of gout,
which aims to dissolve MSU crystals, suppress gout flares and resolve tophi. Allopurinol is the
first-line urate-​lowering therapy and should be started at a low dose, with gradual dose
escalation. Low-dose anti-​inflammatory therapies can reduce gout flares during initiation of
urate-lowering therapy. Models of care, such as nurse-​led strategies that focus on patient
engagement and education, substantially improve clinical outcomes and now represent best
practice for gout management.

Gout is the most common form of inflammatory arthri- In some individuals with gout, tophi can also form.
tis that affects adults1. Gout is caused by the chronic These nodules, consisting of both deposits of MSU
1
Department of Medicine, deposition of monosodium urate (MSU) crystals (Fig. 1). crystals and chronic granulomatous inflammatory tis-
University of Auckland, Clinical disease occurs as a result of the inflammatory sue, represent a high burden of MSU crystal deposition
Auckland, New Zealand.
response of host tissue to deposited MSU crystals2. and contribute to chronic joint pain and damage. In
2
Department of Medicine, MSU crystal deposition and the clinical presentation the absence of urate-​lowering therapy, tophaceous gout
Massachusetts General
Hospital, Boston, MA, USA.
of gout occur in some individuals with elevated serum usually develops ~10 years after the initial gout flare5.
3
Department of Internal
urate concentrations (hyperuricaemia) (Fig. 1). Urate is Classification criteria for gout have been endorsed
Medicine, Radboud University the degradation product of purines, the main sources of by the American College of Rheumatology (ACR) and
Medical Centre, Nijmegen, which are cell turnover, dietary intake and de novo syn- European League Against Rheumatism (EULAR)6.
Netherlands. thesis. A generally agreed definition of hyperuricaemia, Of note, the classification criteria were developed for
4
Division of Rheumatology, serum urate >0.41 mmol/l (6.8 mg/dl), was established research purposes and should not replace comprehensive
University of Michigan, using laboratory-​based crystallization assays in which clinical assessment for diagnosis in individual patients.
Ann Arbor, MI, USA.
different concentrations of urate were added at varying These criteria capture the key aspects of the illness, which
5
Department of Integrative
temperature and pH. However, MSU crystals may form are identified by clinical history and examination, labo-
Physiology and Bio-​Nano
Medicine, National Defense
at lower urate concentrations, particularly in tissues with ratory tests and imaging. An online calculator, the ACR–
Medical College, Tokorozawa, a low temperature or acidic pH3. EULAR Gout Classification Criteria calculator (http://
Japan. As MSU crystals preferentially deposit within joints goutclassificationcalculator.auckland.ac.nz), is available
6
Rheumatology Division, and periarticular structures, particularly at the first to enable rapid scoring of individuals for eligibility for
Hospital Universitario Cruces, metatarsophalangeal joint, the midfoot and the knee, clinical trials. A recent consensus statement endorsed
Baracaldo, Vizcaya, Spain.
gout typically presents as an acute, intensely painful, by the Gout, Hyperuricemia and Crystal-​Associated
7
Department of Medicine, inflammatory arthritis in these regions4. Gout flares usu- Disease Network (G-​CAN) has also described agreed
University of Otago,
Christchurch, New Zealand.
ally self-​resolve within 7–10 days and are interspersed labels and definitions for basic elements of gout7 (Box 1).
between asymptomatic ‘intercritical periods’. Over time, In this Primer, we discuss the epidemiology, patho-
*e-​mail: n.dalbeth@
auckland.ac.nz prolonged hyperuricaemia may result in more frequent physiology, diagnosis, screening and prevention, as well
https://doi.org/10.1038/ and severe flares that also affect the upper limbs and as the management and quality of life (QOL) issues that
s41572-019-0115-y multiple joints (polyarticular flares)5. are associated with gout.


NATURE REvIEwS | DISeASe PrIMerS | Article citation ID: (2019) 5:69 1

0123456789();
Primer

Clinical Disease Risk factor a substantial increase in soft drink and fructose con-
manifestation progression or cause
sumption (known risk factors for gout) following the
No disease Normouricaemia • Genetic factors (e.g. SNPs widespread introduction of high-​fructose corn syrup9–12.
in urate transporter genes) Most recent estimates indicate that the prevalence has
• Environmental factors
(e.g. diet and BMI) now stabilized in the USA1. UK studies based on general
• Impaired kidney function practice diagnostic indices found similarly increasing
• Increasing age prevalence estimates of gout between the 1970s13 and the
Hyperuricaemia • Male sex
• Medications 1990s14. The trend of increasing prevalence continued
Asymptomatic
state • High cell turnover states, etc. between 1997 and 2012 (the most recent estimate is a
prevalence of 2.5%)15. A Taiwanese Nutrition and Health
MSU crystal deposition • Reduced solubility of urate Survey in the mid-1990s found a prevalence similar to
• Increased nucleation
of MSU crystals that in the USA (3.4%)16, whereas a 2010 study using the
• Growth of MSU crystals Taiwanese National Health Insurance Research Database
Recurrent gout flares
found a higher prevalence of 6.2%17.
Symptomatic Similarly, incidence estimates have increased over
disease the years 18. Serial investigations of the Rochester
Acute inflammatory response
to deposited MSU crystals Epidemiology Project in Minnesota (USA) showed that
the incidence of gout without diuretic exposure doubled
from 20 per 100,000 individuals in 1977–1978 to 46 per
Chronic gouty arthritis 100,000 individuals in 1995–1996 (ref.19). From the 1990s
and tophaceous gout
Symptomatic Chronic granulomatous until 2012, UK and Canadian population-​based stud-
disease with inflammatory response to ies found trends of rising gout incidence over the past
complications deposited crystals
several decades15,20.
The modern gout epidemic is also reflected in the
growing increase in hospitalizations due to gout over
Fig. 1 | Disease progression in gout. The transition from normouricaemia to clinically
the past several decades. For example, from 1993 to 2011
evident gout occurs in a number of steps. The first step is the development of
hyperuricaemia, which can be caused by various factors, including genetic variants, in the USA, the annual hospitalization rate for patients
chronic kidney disease, high body mass index (BMI), medications and dietary factors. with a principal diagnosis of gout doubled (from 4.4 to
In some individuals with hyperuricaemia, monosodium urate (MSU) crystal deposition 8.8 per 100,000 adults; P < 0.001), whereas that for rheum­
occurs, and in some individuals with MSU crystal deposition, the clinical manifestations atoid arthritis declined by 67% (from 13.9 to 4.6 per
of gout (gout flares, chronic gouty arthritis and tophaceous gout) occur. Factors that 100,000 adults; P < 0.001), which represents an inversion
contribute to the transition from hyperuricaemia to clinically evident gout are less well of the hospitalization rates for the two conditions over
understood. SNP, single-​nucleotide polymorphism. the past two decades21. These trends persisted among
subgroups by age, sex, race and geographic region
Epidemiology (P < 0.001 for all)21, and have also been found in other
The disease burden of gout is substantial and is increas- countries, including Canada, Aotearoa/New Zealand
ing in Western countries and parts of the world that are and Sweden20,22,23. This inversion in hospitalization rates
becoming westernized, suggesting the existence of a is likely due to decreases in joint replacements and sys-
modern gout epidemic that is analogous to the obesity temic complications in patients with rheumatoid arthri-
epidemic. Both conditions are aspects of the metabolic tis, and reflects improvements in care for rheumatoid
syndrome, the incidence of which has largely coincided arthritis and ongoing suboptimal gout management24,25,
with the rising prevalence of hyperuricaemia. Reported including high rates worldwide of recurrent gout flares.
estimates of gout prevalence range from 2.7% to 6.7% in
countries with a Western lifestyle. The most recent esti- Risk factors
mate (in 2015–2016) of the lifetime prevalence of gout Sustained elevation of serum urate levels is the main risk
in adults diagnosed by a health professional in the USA factor for developing gout. Different factors that induce
(3.9%) equates to 9.2 million individuals1. This preva­ hyperuricaemia, such as medical conditions, obesity,
lence increases with age to 9% of adults >60 years of lifestyle factors and medications, are therefore associated
age in the USA. The most recent estimate of gout preva­ with the risk of developing gout8. Hyperuricaemia due to
lence in mainland China is 1.1%8. Gout is particularly urate overproduction can occur in individuals with dis-
common in indigenous Taiwanese peoples and Polynesian orders of high cell turnover, such as psoriasis and myelo-
peoples, including Māori (indigenous New Zealanders), proliferative disorders. High intake of purine-​rich foods
and Pacific peoples from both Western and Eastern (such as alcohol) or those that can lead to increased
Polynesia. In these populations, prevalence of gout is purine levels (such as fructose) can also contribute to
estimated at >8% of adults8. hyperuricaemia. Chronic kidney disease (CKD), meta-
In the USA, the prevalence of hyperuricaemia and bolic syndrome and diuretic use can contribute to renal
mean serum urate levels in 2007–2008 were substan- under-​excretion of urate and thereby result in hyper­
tially higher than corresponding estimates for the period uricaemia. Many of these factors are also risk factors for
1988–1994. Between 1969 and 1996, serial US National the development of incident gout.
Health Interview Surveys showed that the annual In patients with established gout, elevated serum
prevalence of gout doubled, and the steepest increase urate level is also a risk factor for future gout flares26. In
occurred between 1969 and 1976, which coincides with addition, gout flares may be triggered by joint trauma,

2 | Article citation ID: (2019) 5:69 www.nature.com/nrdp

0123456789();
Primer

acute medical or surgical illness, high purine intake, of gout have microscopy-​based and imaging-​based
alcohol intake and dehydration16,27–29. evidence of MSU crystal deposition35,36. The gout flare
Sex and advancing age are also important risk factors represents an acute inflammatory response to deposited
for gout. Men are more likely to experience gout than MSU crystals2, and the tophus is a chronic inflammatory
women: among American adults in 2015–2016, the prev- granulomatous response to deposited crystals37.
alence of gout was 5.2% (5.9 million) in men and 2.7%
(3.3 million) in women1. Gout is uncommon in women Hyperuricaemia
before the menopause and the risk of incident gout is During primate evolution, inactivating mutations
modestly reduced by hormone replacement therapy in in the gene encoding uricase occurred that resulted in
postmenopausal women30. Compared to men with gout, an inability to metabolize urate to soluble allantoin
women with gout have later age of onset and are more likely in humans38. Consequently, urate is the end-​product of
to have comorbidities, particularly CKD, hypertension human purine metabolism, and humans have higher
and conditions requiring diuretic treatment31. serum urate levels than most other mammals. Urate has
a number of putative biological functions, including act-
Mechanisms/pathophysiology ing as an antioxidant39, acting as an endogenous danger
Gout develops in a number of stages (Fig. 1). The first signal in innate immune responses40, and maintaining
is development of hyperuricaemia, which is considered blood pressure in individuals with a low-​salt diet41.
necessary but not sufficient for the development of gout. In population studies, various environmental factors
In individuals without gout, there is a concentration-​ are associated with hyperuricaemia and the development
dependent association between serum urate levels and of incident gout, including intake of purine-​rich foods,
the development of incident gout32,33. However, the such as beer, meats and seafood42,43. In addition, fructose
majority of individuals with hyperuricaemia do not intake increases purine nucleotide degradation due to
develop gout, even over prolonged periods of obser- fructose phosphorylation and ATP breakdown into ADP
vation32,34. For gout to develop, the deposition of MSU and AMP. AMP may enter the purine nucleotide degra-
crystals must occur; however, some asymptomatic dation pathway, leading to increased serum urate con-
individuals with hyperuricaemia and no prior history centrations44–46. A similar mechanism has been proposed
to explain the hyperuricaemic effect of alcohol, which
is partially due to its hepatic metabolism into acetate,
Box 1 | G-​CAN consensus labels and definitions in gout a reaction that consumes ATP and generates AMP47,48.
The Gout, Hyperuricemia and Crystal-​Associated Disease Network (G-​CAN) published Some dietary factors are associated with reduced serum
a consensus statement to standardize the nomenclature for describing the basic urate concentrations, including coffee, low-​fat dairy
disease elements in gout7. produce and vitamin C42,49,50. Furthermore, elevated
Monosodium urate crystals body mass index, kidney disease and diuretic use are
The pathogenic crystals in gout (chemical formula: C5H4N4NaO3). associated with hyperuricaemia and development of
Urate gout51,52. Both genetic and environmental factors con-
The circulating form of the final enzymatic product generated by xanthine oxidase in tribute to elevated serum urate concentrations, and in
purine metabolism in humans (chemical formula: C5H3N4O3−). the modern environment, genetic factors have a larger
effect than specific dietary factors on variation in serum
Hyperuricaemia
Elevated blood urate concentration over the saturation threshold.
urate concentrations53,54.
Gout has traditionally been considered a disorder
Tophus of purine metabolism. However, urate overproduction
An ordered structure of monosodium urate crystals and the associated host tissue
is the major cause of hyperuricaemia in only a fairly
response.
small proportion of individuals with gout55. In the
Subcutaneous tophus past decade, it has become apparent that altered urate
A tophus that is detectable by physical examination. transport, both in the gut and the kidneys, has a central
Gout flare role in the pathogenesis of hyperuricaemia and gout.
A clinically evident episode of acute inflammation induced by monosodium urate Genome-​wide association studies (GWAS) have identi-
crystals. fied single-​nucleotide polymorphisms in numerous loci
Intercritical gout that are associated with variation in serum urate con-
The asymptomatic period after or between gout flares, despite the persistence of centrations56,57, many of which are also associated with
monosodium urate crystals. gout20,56–60. The urate transporter genes SLC2A9 (encod-
Chronic gouty arthritis ing GLUT9), SLC22A12 (encoding URAT1), SLC17A1
Persistent joint inflammation induced by monosodium urate crystals. (encoding NPT1) and ABCG2 are most strongly associ-
ated with variation in serum urate levels. Although many
Imaging evidence of monosodium urate crystal deposition
Findings that are highly suggestive of monosodium urate crystals on an imaging test. other genes have been associated with hyperuricaemia
and gout, serum urate levels are mainly regulated by the
Gouty bone erosion activity of these four transporters in renal transport, and
Evidence of a cortical break in bone suggestive of gout (overhanging edge with
of ABCG2 in intestinal transport20 (Fig. 2).
sclerotic margin).
URAT1 has a major role in reabsorption of urate
Podagra at the apical membrane in the proximal renal tubule61.
A gout flare at the first metatarsophalangeal joint. G-CAN recommends that the label This transporter is also inhibited by uricosuric agents,
‘chronic gout’ should be avoided.
such as probenecid, benzbromarone and lesinurad61–63.


NATURE REvIEwS | DISeASe PrIMerS | Article citation ID: (2019) 5:69 3

0123456789();
Primer

a
Proximal tubular cell

Kidney Nephron
URAT1

Urate
GLUT9 Urate ABCG2
Urate
Urate NPT1
Basolateral Apical
Renal (blood (urine
tubule side) side)

b Enterocyte

Intestine

Urate ABCG2

Basolateral Apical
Intestinal (blood (luminal
lumen side) side)

Fig. 2 | Urate transporters in the kidneys and intestine in humans. The urate transporters GLUT9, URAT1, NPT1 and
ABCG2 regulate serum urate levels in humans. GLUT9 in the basolateral membrane and URAT1 in the apical membrane of
proximal tubule cells mediate renal urate reabsorption. NPT1 is present on the apical membrane of renal proximal tubule
cells and mediates urate excretion, and a gain-​of-function mutation in the gene encoding NPT1 results in increased renal
urate excretion and can decrease the risk of gout. ABCG2 is present on the apical surface of proximal tubule cells and
enterocytes and is involved in urate export in both the intestine and the kidneys. Dysfunctional variants of ABCG2
increase serum urate levels and the risk of gout. A number of other molecules may also regulate urate transport in the
renal tubule and gut, including the organic acid transporters. Adapted from Ann. Rheum. Dis., Nakayama, A. et al. 76,
869–877 (2017) with permission from BMJ Publishing Group Ltd.20, and from ref.210, Springer Nature Limited.

GLUT9 is responsible for renal reabsorption of urate at Although not all functions of these susceptibility
the basolateral membrane in the proximal renal tubule64. genes are known, analysis of genetic function has led to
Mutations in SLC22A12 (ref.61) and SLC2A9 (refs64,65) important new insights into the molecular mechanisms
cause renal hypouricaemia type 1 and type 2, respec- of hyperuricaemia78. Based on the clinical parameters,
tively66, which are disorders that were first described in hyperuricaemia has traditionally been classified as
the Japanese population that cause not only hypouri- ‘overproduction type’ or ‘renal underexcretion type’
caemia, but also exercise-​induced acute kidney injury hyperuricaemia79,80. However, analysis of ABCG2 func-
and kidney stones. Common variants of SLC2A9 and tion based on genetic testing in patients with hyper-
SLC22A12 are associated with hyperuricaemia and the uricaemia has led to a revision of this classification.
risk of gout20,56. Although dysfunction of renal tubular ABCG2 is pre-
Genetic variants in secretory renal transporter genes dicted to cause renal underexcretion of urate, patients
are also associated with hyperuricaemia and gout. For with dysfunctional variants of ABCG2 show paradoxi-
example, SLC17A1 encodes the renal urate secretory cal urate overproduction78, based on clinical parameters
transporter NPT1, which is expressed on the apical such as urinary urate excretion and fractional excretion
membrane67 in the proximal renal tubule, and a com- of uric acid (FEUA; the urate clearance/creatinine clear-
mon gain-​of-function variant (I269T) of this gene ance ratio). Analysis of human genetic data and Abcg2-
decreases the risk of gout67,68. ABCG2 is a urate excre- knockout mice revealed that dysfunctional ABCG2 in the
tion transporter69,70 in the kidneys71 and intestine72, and intestine is another cause of hyperuri­caemia, which can
common dysfunctional variants (Q126X and Q141K) be classified as ‘extra-renal underexcretion type’ hyper-
of ABCG2 strongly increase serum urate levels and the uricaemia78. Both (genuine) overproduction and extra-
risk of gout56,69,70,73,74. Moreover, rare non-​synonymous renal underexcretion can contribute to ‘renal overload
variants of ABCG2 also markedly increase the risk of type’ hyperuricaemia78 (Fig. 3).
gout75,76, supporting the ‘Common Disease, Common
Variant’ and ‘Common Disease, Multiple Rare Variant’ MSU crystal deposition
hypotheses of gout susceptibility75. Some sex differences MSU crystal deposition is the next stage in the devel-
have been described for ABCG2 variants and gout risk, opment of gout. MSU crystals appear microscopically
with a greater effect of Q141K on gout risk in men than as negatively birefringent, needle-​shaped crystals,
in women77. and at the molecular level consist of stacked sheets of

4 | Article citation ID: (2019) 5:69 www.nature.com/nrdp

0123456789();
Primer

closely spaced purine rings81. Advanced imaging stud- The gout flare
ies using ultrasonography and dual-​energy CT (DECT) Some individuals with evidence of MSU crystal depo-
have shown that ~25% of individuals with hyperuri­ sition develop clinical evidence of gout, most typically
caemia have evidence of MSU crystal deposition35,36. In a gout flare, with rapid onset of a painful, swollen, hot
individuals with gout, MSU crystal deposition occurs and red joint4. Pathologically, the joint affected by the
preferentially at certain sites, such as the first meta- gout flare is characterized by a pronounced neutrophilic
tarsophalangeal joint, midfoot and Achilles tendon82. infiltration, both in synovial tissue and synovial fluid84.
These observations suggest that factors other than urate Many soluble mediators have been implicated in
concentration contribute to MSU crystal formation, initiation and amplification of the gout flare, including
although specific factors have not been identified. pro-​inflammatory cytokines, lipid mediators and com-
In general, three steps are required for formation plement. However, the cytokine that has a pivotal role in
of crystals: reduced solubility (leading to supersatura- initiation of the gout flare is IL-1β85,86 (Fig. 4). MSU crystals
tion), nucleation and crystal growth. Increased urate interact with resident macrophages to activate the NOD-,
concentration is an essential factor for all three stages LRR- and pyrin domain-​containing protein 3 (NLRP3)
of MSU crystallization83. In laboratory-​based assays of inflammasome, leading to release of bioactive IL-1β87.
MSU crystal formation, crystallization occurred at urate Clinical trials have shown that biological agents target-
concentrations >0.41 mmol/l (6.8 mg/dl) at 37 °C and pH ing IL-1β have anti-​inflammatory effects in the gout flare;
7.0, whereas at 35 °C, crystallization occurred at a urate these agents can prevent development of the first gout
concentration of 0.36 mmol/l (6 mg/dl)3. In addition to flare88, prevent further gout flares in patients with prior
lower temperature, urate solubility is reduced in the pH flares88,89, and treat flares once they have occurred90,91.
7–8 range and by high concentrations of sodium ions83. After the demonstration that the NLRP3 inflamma­
Many other factors have been reported to contribute to some is involved in the gout flare87, it was believed that
MSU crystallization, including connective tissue factors, only MSU crystals were responsible for the formation
bovine cartilage homogenates, human synovial fluid and and activation of the NLRP3–ASC–caspase 1 protein com-
urate-​binding antibodies83. plex, which processes pro-IL-1β to bioactive IL-1β87,92.
However, MSU crystals are present in some asympto-
Traditional classification matic individuals with hyperuricaemia36 and in individ-
A Overproduction type uals with prior gout flares, but may not induce an acute
• High cell turnover disorders inflammatory response93. The production of bioactive
• Purine-rich diet
• Lesch–Nyhan syndrome, etc.
IL-1β by tissue-​resident macrophages is dependent on
B Renal underexcretion type
two signals, one signal to upregulate IL1B transcription
• CKD A C B and synthesis of pro-​IL-1β and the second signal to activ­
• Metabolic syndrome ate the NLRP3 inflammasome94. Damage-associated
• Diuretic use molecular patterns (such as free fatty acids) and micro-
• Renal urate transporter
gene variants, etc. bial products (such as lipopolysaccharide) have been
C Combined type (overproduction identified as signals that upregulate IL-1β mRNA and
and renal underexcretion) pro-​IL-1β protein levels95,96. During the initiation phase
of a gout flare, the NLRP3 inflammasome-​mediated
production of bioactive IL-1β seems to be important87.
However, once neutrophils enter the site of inflam-
mation, the role of the inflammasome is bypassed by
New classification A1 serine proteinases, such as proteinase 3 and neutrophil
A Renal overload type C
elastase97, suggesting that effective treatment of gout
A1 Overproduction type flares requires targeting of both the inflammasome and
C
A2 Extra-renal underexcretion type B neutrophil serine proteinases.
• ABCG2 dysfunction in the gut
• Intestinal diseases, etc. Myeloid cells (such as monocytes) from patients with
B Renal underexcretion type C gout or hyperuricaemia produce higher levels of pro-​
C Combined type (renal overload inflammatory cytokines than those from age-​matched
and renal underexcretion) A2 and sex-​matched healthy individuals98. This finding
has been linked to the exposure of monocytes to sol-
uble urate97. Urate exposure causes a state of trained
Fig. 3 | Classification of the causes of hyperuricaemia. In the traditional classification, immunity or innate memory that results in epigenetic
hyperuricaemia is caused by urate overproduction, typically from a purine-​rich diet, high modifications, such as DNA methylation and histone
cell-​turnover disorders or Lesch–Nyhan syndrome, or by renal underexcretion of urate, methylation and acetylation, which lead to increased
which can result from pathological conditions (such as chronic kidney disease (CKD) production of pro-​inflammatory cytokines, such as
or metabolic syndrome), diuretic use or variants in renal urate transporter genes. IL-1β and IL-6, by these cells97,99.
However, genome-​wide association studies in humans and studies in animal models
have revealed that extra-​renal underexcretion also exists. Intestinal diseases (such as
acute gastroenteritis) or dysfunctional variants of the urate transporter ABCG2 in the Resolution of the gout flare. The gout flare is a self-​
gut can result in underexcretion of urate from the gut, resulting in renal overload limiting inflammation — the redness, joint swelling and
hyperuricaemia. The areas of overlap (C) denote theoretical patients with the severe pain usually disappear after 7–10 days. Several
pathophysiological characteristics of urate overload and renal underexcretion of the mechanisms of resolution have been proposed. In addi-
corresponding circumferences. Figure adapted from ref.78, Springer Nature Limited. tion to their previously mentioned role in promoting


NATURE REvIEwS | DISeASe PrIMerS | Article citation ID: (2019) 5:69 5

0123456789();
Primer

Metabolic
Macrophage Uptake of
changes
• Food Pro-IL-1β complement-coated
• Alcohol MSU crystals
• Fasting
IL-8
NLRP3
TIR inflammasome Neutrophil
↑ FFAs TLR2 MyD88
TIR

Neutrophil
PYD serine
ASC proteinases NETosis
CARD
Complement-
coated IL-1β
MSU crystals aggNET

TIR Active
↑ LPS TLR4 MyD88 caspase 1 ↑ Anti-inflammatory
TIR
factors Inactivation of pro-inflammatory
• IL-1Ra • IL-37 cytokines
Bacterial Pro-IL-1β • IL-10 • TGFβ1 • IL-1β • IL-6 • IL-8
infections

Initiation Resolution
Inflammation

Time

Fig. 4 | Mechanisms of initiation and resolution of the gout flare. Acute joint inflammation is initiated by the
production of pro-​inflammatory cytokines, in particular IL-1β, the production and release of which is a multistep process.
IL-1β gene transcription and production of pro-​IL-1β is induced by activation of Toll-​like receptor 2 (TLR2) or TLR4. Free
fatty acids (FFAs; such as palmitic acid (C16:0)) that accumulate due to metabolic alterations or lipopolysaccharide (LPS)
released during bacterial infection can trigger TLR2 or TLR4 signalling, respectively , and thereby induce the intracellular
accumulation of pro-​IL-1β. The uptake of monosodium urate (MSU) crystals activates the NLRP3 inflammasome, resulting
in activation of caspase 1, which enzymatically processes pro-​IL-1β to bioactive IL-1β. IL-1β signalling initiates the
production and secretion of pro-​inflammatory mediators, such as IL-6 and IL-8. IL-8 recruits neutrophils to the joint cavity
and thereby amplifies the gout flare. After some time, the gout flare self-​resolves through the production of anti-​inflammatory
factors in the joint, such as interleukin-1 receptor antagonist (IL-1Ra), IL-10, IL-37 and transforming growth factor β1
(TGFβ1), which suppress the pro-​inflammatory factors. Furthermore, the large numbers of neutrophils in the joint
eventually die, either by normal cell turnover or after uptake of MSU crystals. These dying neutrophils release their
chromosomal DNA (termed neutrophil extracellular traps (NETs)), a process called NETosis. Aggregated NETs (aggNETs)
sequester and thereby inactivate several pro-​inflammatory mediators, which leads to resolution of the gout flare.
MSU, monosodium urate; TIR , Toll/interleukin-1 (IL-1) receptor domain.

inflammation during a gout flare (by processing pro-​ cytokines and chemokines, which are degraded by serine
IL-1β97), neutrophils might also contribute to the reso­ proteases105, leading to resolution of the gout flare.
lution of inflammation in the gout flare. Increased
production of anti-​inflammatory mediators, such inter- Tophaceous gout and joint damage
leukin 1 receptor antagonist (IL-1Ra), IL-10 and TGFβ1, Tophi present as subcutaneous nodules, usually without
or pro-​resolving lipids by neutrophils may downregu- acute inflammation. Pathologically, the tophus has an
late inflammation of the joint100–102. Furthermore, release organized architecture, comprising a region of chronic
of the anti-​inflammatory IL-1 family member IL-37 foreign-​b ody granulomatous response surrounding
by immune and non-​immune cells also suppresses collections of MSU crystals106. In addition to resolu-
inflammation in gout103. tion of the gout flare, aggNETs may also have a role in
In the late stages of a gout flare, phagocytosis of tophus formation by binding to MSU crystals in a non-​
dying neutrophils by live neutrophils (neutrophil canni­ inflammatory state105. Pro-​inflammatory cytokines, such
balism) can promote TGFβ1 production and resolu- as IL-1β and TNF, and the anti-​inflammatory cytokine
tion of the inflammatory response104. In addition, after TGFβ1 are expressed in the tophus, suggesting that there
phagocytosis of MSU crystals, dying neutrophils release is a cycle of chronic inflammation, attempted resolution
their chromosomal DNA (these fibres are termed and tissue remodelling37,107.
neutro­phil extracellular traps (NETs)), a process referred The presence of a tophus is closely linked to joint
to as NETosis105. At high neutrophil densities, aggre- damage in patients with gout. Imaging and histologi-
gated NETs (aggNETs) sequester pro-​inflammatory cal studies have revealed that most joints with bone

6 | Article citation ID: (2019) 5:69 www.nature.com/nrdp

0123456789();
Primer

erosion have a tophus in close proximity to the site accuracy of diagnosis in the emergency room falls from
of erosion106,108. Imaging studies have demonstrated that 85% in patients with typical (monoarticular) flares to
MSU crystals deposit within the joint and erode from 45% in those with atypical (non-​monoarticular) flares117.
the synovial surface into bone, rather than depositing In patients with very high and sustained hyperuricaemia,
within bone109. Disordered osteoclastogenesis has been symmetrical arthritis and flares involving small joints of
implicated in the pathogenesis of bone erosion in gout, the hands may be observed, especially in patients with
as numerous osteoclasts are present at the tophus– underlying osteoarthritis118, and tophi may grow in atypi-
erosion interface110. MSU crystals have direct effects on cal locations119. The presence of osteoarthritis may favour
the stromal cells of bone, leading to reduced cell viabil- urate crystal nucleation and therefore gout120, whereas
ity and function of osteoblasts111. In addition to direct involvement of distal interphalangeal joints, which may
effects on cell viability, soluble pro-​inflammatory medi­ develop in patients affected by osteoarthritis, may chal-
ators induced by MSU crystals can have an indirect effect lenge a diagnosis of gout, as it is a typical distribution
on osteocytes, further promoting a resorptive state112. for psoriatic arthritis. Uncommonly, gout may involve
proximal, large joints other than the knee, and the spine117.
Diagnosis, screening and prevention
Presentation Diagnosis
The natural history of untreated gout was reported by Accurate diagnosis requires recognition of both typical
Gutman in the mid-20th century, using probably the and atypical presentations of gout, together with those
only data of this type available, a clinical database that factors that may influence the presentation and the nat-
was established prior to the advent of effective urate-​ ural history of untreated disease. Although microscopy-​
lowering medications5. Briefly, some patients with gout based diagnosis is still currently considered the gold
will develop polyarticular, erosive, tophaceous gout and standard, imaging techniques (such as ultrasono­graphy)
have an increasing number of flares5. This severe disease that are available for general clinical practice and that
usually takes decades to develop and is often associated provide early and specific findings of gout might be
with high serum urate levels over prolonged periods of useful, and could even replace microscopic diagnosis.
time. Prior to urate-​lowering therapy, 72% of patients Hyperuricaemia is the laboratory hallmark of gout
with a disease duration of 20 years developed tophi, and but is not diagnostic and may lead to misdiagnosis.
severe, crippling tophaceous disease was observed in As hyperuricaemia is a frequent finding in adults, some
24% of these patients5. musculoskeletal disorders may be misdiagnosed as gout;
conversely, atypical gout may be misdiagnosed as other
Typical clinical manifestations diseases. Clinical diagnosis is considered to be mostly
Acute episodes of inflammation (gout flares7) involving accurate in patients presenting with typical clinical mani­
joint structures (associated with local erythema when festations of gout116, but in patients with longstanding
small peripheral joints are affected), with typically asym- gout, both clinical diagnosis121 and classification crite-
metrical involvement of only one joint of the lower limbs, ria122 seem to be less accurate methods than the gold
especially the first metatarsophalangeal (podagra) and standard, microscopy-​based diagnosis.
tarsal joints, are the paradigm of gout presentation113.
Gout flares are mostly self-​limiting, especially early in the Differential diagnosis
natural history of gout. Nodules indicating tophaceous Early in the natural history of the disease, differential
deposits are typically located at the first metatarso­ diagnoses include diseases that acutely and asymmet-
phalangeal joint, other joints and tendons of the foot rically affect the joints of the lower limb, such as acute
and ankle, including the Achilles tendon, the prepatella calcium pyrophosphate crystal arthritis, psoriatic arthri-
bursa, the olecranon bursae and helix of the ear. These tis, reactive arthritis and Lyme disease. Gout flares
nodules may be pain-​free, and a careful examination of with fever and malaise may mimic septic arthritis. In
typical sites is recommended. patients with longstanding gout, polyarticular, symmet-
rical arthritis, even involving the upper limbs, should
Atypical clinical manifestations be differentiated from chronic calcium pyrophosphate
Atypical presentation may include the development crystal arthritis, rheumatoid arthritis (especially when
of tophi without accompanying flares, which may be tophi may be confused for rheumatoid nodules) and
observed in patients who are receiving corticosteroids palindromic rheumatism117.
for other conditions114. Furthermore, first presentation Of note, gout may develop in patients with other com-
as a flare involving multiple joints or several joints sym- mon types of inflammatory arthritis in whom longstand-
metrically distributed is also atypical. Atypical manifes- ing hyperuricaemia is present. This should be considered
tations of gout occur more frequently in women31 and in patients with other musculoskeletal conditions and
elderly individuals115, and this should be considered in hyperuricaemia who develop acute flares of arthritis6.
the diagnosis.
Intermittent typical gout flares (monoarticular, asym- Classification and diagnostic criteria
metrically distributed in the joints of the lower limbs) Classification criteria have been developed for research
may persist for a variable period of time, but some purposes, rather than for clinical diagnosis. The 2015
patients with prolonged disease may develop atypical EULAR and ACR classification criteria are more sensitive
manifestations, which may make diagnosis uncertain and specific than the previous criteria6, and have been
unless it is confirmed by microscopy116. Indeed, the validated in primary and secondary care settings6,123.


NATURE REvIEwS | DISeASe PrIMerS | Article citation ID: (2019) 5:69 7

0123456789();
Primer

A diagnostic rule for gout in patients presenting with this feature is not specific for gout, unless tophi or MSU
monoarthritis has been developed in primary care, using deposits are present84. Of note, to obtain samples that are
microscopy to confirm the diagnosis113. This weighted suitable for pathological analysis, biopsy samples should
score includes the following variables: serum urate be fixed in absolute alcohol, as MSU crystals dissolve in
>0.36 mmol/l, first metatarsophalangeal joint involve- formalin84.
ment, male sex, previous patient-​reported arthritis
attack, hypertension or one or more cardiovascular dis- Imaging for diagnosis and assessment
eases, joint redness and onset within one day. A score of MSU crystals can be detected using ultrasonography
≤4 indicates that gout is unlikely, a score of ≥8 indicates and DECT, which can both be useful to evaluate the
that gout is very likely (>80%), and a score in the range symptoms in patients with gout and complaints at other
4–8 leaves uncertainty about the diagnosis (with gout joints129 or atypical clinical manifestations130. DECT
confirmed in about 30% of patients). In this situation, enables colour coding of urate attenuation by analysing
synovial fluid analysis is recommended to provide fur- differences in materials exposed to two different X-​ray
ther diagnostic certainty. This diagnostic rule has been spectra simultaneously131 (Fig. 6). A ‘double contour’
validated in a number of clinical settings123–125. sign on ultrasonography, a finding that correlates with
crystal deposition in the surface of the hyaline cartilage,
Microscopy-​based diagnosis is included in the new classification criteria for gout,
The presence of MSU crystals confirmed microscopi- as it adds discriminating value4, and the presence of
cally in either synovial fluid samples or material aspi- associated tophi makes diagnosis quite certain (Fig. 7).
rated from tophi is definitive for a diagnosis of gout, and Some studies have shown that ultrasonography is
is recommended for diagnosis in patients with atypical more sensitive for diagnosis than DECT132, especially in
clinical manifestations116. Ultrasonography may be use- early disease or when the crystal deposition burden is
ful to guide intra-​articular aspiration36 and puncture of not high. However, although ultrasonography features
non-​superficial nodules126. Arthrocentesis for collection of MSU crystal deposition have high specificity and
of synovial fluid for gout diagnosis is a safe and well-​ high positive predictive value, they show more limited
tolerated procedure in experienced hands127. Currently, sensitivity for early gout, defined as just the initial, inter-
microscopy-​based confirmation of MSU crystals in tis- mittent, acute flares133. Ultrasonography is also useful to
sues or synovial fluid remains the gold standard for diag- evaluate associated tissue inflammation and both ultra-
nosis. Microscopy may also be very helpful in patients sonography and DECT may also be useful to monitor
with acute intermittent arthritis, hyperuricaemia and the response of MSU deposits to urate-​lowering therapy.
chondrocalcinosis128. Although rare, both MSU and cal- Plain radiographs do not show erosions or other
cium pyrophosphate dihydrate crystals can be present in bone or joint changes until late in the natural history of
a sample of synovial fluid (Fig. 5). disease. Bone erosions with overhanging edges envel-
During a gout flare, the synovial membrane has the oping periarticular soft tissue masses are suggestive of
appearance of an acute neutrophilic synovitis, although gout4. With the exception of tophi, MRI and CT do not
show specific findings of gout but may be useful to eval-
uate soft tissue and bone involvement in gout or other
comorbidities.

Laboratory assays
Hyperuricaemia is the hallmark of laboratory findings
λ in gout. Compared with low serum urate levels, hyper-
uricaemia is a variable that adds discriminating value
for classification compared to gold-​standard diagnosis
by microscopy4. Serum urate concentrations may be
within normal limits during flares134 and especially in
postsurgical flares135. If gout is suspected and serum
urate concentration is within the normal range at the
time of presentation with acute inflammatory arthritis,
this test should be repeated once the flare has resolved.
High leukocyte counts in peripheral blood or syn-
ovial fluid and elevated acute-​phase reactants are non-
specific findings in acute arthritis, irrespective of the
underlying cause117.
Evaluation of renal excretion of uric acid using frac-
tional excretion, Simkin’s index or 24-hour urine collec-
Fig. 5 | Identification of MSU crystals using polarizing light microscopy. Light
tions136 may be useful when evaluating renal efficiency
microscopy using polarized light with a first-​order compensating filter is used to detect
monosodium urate (MSU) crystals in synovial fluid from a patient with gout. MSU crystals to excrete uric acid as a cause for hyperuricaemia. FEUA
appear as large, bright (strongly birefringent) needle-​shaped crystals (black arrows) on a spot urine sample is a convenient and reliable indi-
with negative elongation (yellow when parallel to the polarizing axis (λ)). Calcium cator of renal uric acid clearance in patients with normal
pyrophosphate dihydrate crystals appear as brick-​shaped, mildly birefringent crystals renal function136. FEUA values are lower in individuals
(white arrows) with positive elongation (blue when parallel to λ). Magnification ×400. with gout than in those without gout (average 4.8%

8 | Article citation ID: (2019) 5:69 www.nature.com/nrdp

0123456789();
Primer

a b confirmed diagnosis of gout, screening for evidence of


disease complications, such as tophi and joint damage,
is appropriate, as patients with severe disease are likely
to benefit from more intensive serum urate-​lowering
therapy139,140.
In all patients with gout, screening for commonly
associated comorbidities, including hypertension,
diabetes, dyslipidaemia, CKD, overweight or obesity
and overall cardiovascular risk, is recommended139,140.
Management of these comorbidities is recommended,
as gout is associated with a high risk of cardiovascular
disease and mortality144. Importantly, traditional cardio­
vascular risk scores may under-​estimate the true risk of
cardiovascular disease in individuals with gout145.

Prevention of incident gout


Strategies that reduce serum urate concentrations can
Fig. 6 | Typical clinical and imaging features of gout. a | Gout flare affecting the first prevent incident gout. In obese individuals, an inter-
metatarsophalangeal joint (podagra). b | Dual-​energy CT image of the feet in a patient vention study showed that bariatric surgery reduced
with tophaceous gout, showing monosodium urate crystal deposition (green), serum urate levels and lowered the incidence of gout
particularly at the right first metatarsophalangeal joint. Image in part a courtesy of by 34% over 26-year follow-​up compared with usual
J.G. Puig, Hospital Universitario Quironsauld Madrid, Spain. care146. At 15-year follow-​up, bariatric surgery resulted
in a 3% lower absolute risk of gout. The PPARα ago-
versus 6.9%), although overlap in FEUA values for these nist fenofibrate has uricosuric properties that result in
two groups is common136. Renal excretion analysis may modest reductions in serum urate levels. In a post hoc
also be useful to estimate the baseline risk of develop- analysis of the FIELD (Fenofibrate Intervention and
ing kidney stones in patients to be prescribed uricosuric Event Lowering in Diabetes) trial, fenofibrate treat-
medications137. ment reduced the incidence of gout by 50% in indi-
viduals with type 2 diabetes mellitus compared with
Screening placebo147. In patients with serum urate >0.42 mmol/l
In asymptomatic individuals, screening for hyperuri­ at baseline, the incidence of gout was 13.9% in the pla-
caemia is not routinely advocated138. However, in some cebo group and 5.7% in the fenofibrate group over a
individuals with risk factors for gout, such as family his- 6-year follow-​up. In the placebo-​controlled FEATHER
tory, medical conditions (such as kidney disease, cardio­ study of the urate-​lowering drug febuxostat in Japanese
vascular disease, metabolic syndrome or concomitant patients with CKD stage 3 and asymptomatic hyperuri-
medications. such as diuretics), serum urate testing will caemia (serum urate 0.42–0.60 mmol/l, 7.0–10.0 mg/dl),
identify hyperuricaemia. Although hyperuricaemia is the incidence of gout was lower with febuxostat treat-
the major risk factor for developing gout, the major- ment (0.9%) than with placebo treatment (5.9%) over
ity of asymptomatic individuals with very high serum 108 weeks148.
urate concentrations do not develop gout over prolonged Anti-​inflammatory therapy to prevent the initiation
periods of follow-​up32, and urate-​lowering therapy is of an acute inflammatory response to deposited MSU
not approved in most countries for individuals with crystals might be an alternative strategy for prevention
asymptomatic hyperuricaemia139,140. of incident gout. A post hoc analysis of the CANTOS
In some individuals with hyperuricaemia or other study, a cardiovascular outcomes, placebo-​controlled
major risk factors for gout (such as family history141), trial of the IL-1β-​targeted monoclonal antibody canaki-
MSU crystal deposition can be detected using advanced numab, showed that canakinumab reduced incidence of
imaging methods, such as ultrasonography or DECT. gout by ~50% over a 5-year follow-​up88. This effect was
Positive imaging findings are obtained in ~25% of indi- observed at all serum urate levels, and canakinumab
viduals with hyperuricaemia35,142,143, but it is currently treatment did not alter serum urate concentrations.
unknown whether this predicts development of clini- The cumulative gout incidence in the placebo-​treated
cally evident disease (that is, gout). Furthermore, no group was ~0.025 over the 5-year study period. At
studies have examined the cost–benefit ratio of thera- present, the long-​term consequences of suppressing
peutic interventions, such as urate-​lowering therapy, in inflammation without addressing crystal deposition
asymptomatic individuals with imaging-​based evidence are unknown.
of MSU crystal deposition. Collectively, clinical trial data indicate that strategies
In individuals with hyperuricaemia confirmed by exist to prevent the development of gout in individu-
serum urate testing, the most important screening als with hyperuricaemia, either by reduction of serum
approach is a clinical assessment for features of gout, urate levels or by long-​term anti-​inflammatory ther-
which includes history of gout flares and examination apy. However, of note, the vast majority of participants
for features of arthritis or tophi. Individuals with clinical in the non-​intervention groups in these clinical trials
evidence of gout should be treated according to stan­ did not develop gout over the prolonged observation
dard gout management approaches. In patients with a periods.


NATURE REvIEwS | DISeASe PrIMerS | Article citation ID: (2019) 5:69 9

0123456789();
Primer

a b c

Femur

Fig. 7 | Ultrasonography-​based features of gout. a | The double contour sign (white arrows) in a knee joint, a finding that
correlates with crystal deposition in the surface of the hyaline cartilage. b | A tophus in the first metatarsophalangeal joint
(white circle). c | Synovitis with power Doppler signal in the first metatarsophalangeal joint during a gout flare. Images in
parts a–c courtesy of J.G. Puig and L. Beltrán, Hospital Universitario Quironsauld Madrid, Spain.

Management serum urate levels and titration of urate-lowering ther-


The four general principles in gout management are apy to achieve a specific serum urate target. The gene­
treatment of gout flares, urate-​lowering therapy, anti-​ rally agreed indications for urate-​lowering therapy are an
inflammatory prophylaxis when starting urate-​lowering established diagnosis of gout and one of the following:
therapy, and screening and management of comorbidi- tophi, frequent acute gout flares (two or more per year),
ties associated with gout (Box 2). Patient and health-​care CKD or past urolithiasis. Urate-lowering therapy may
provider education about gout and these management also be indicated in patients with less severe disease, even
principles underpin successful treatment. after the first gout flare, particularly if they have very high
serum urate levels, early age of onset (<40 years of age)
Gout flares or other comorbidities140. If urate-​lowering therapy is
Rapid and effective control of the acute inflammatory indicated, a target urate concentration of <0.36 mmol/l
response is the basis of the treatment of gout flares. is recommended for most patients, although a lower tar-
NSAIDs, colchicine and corticosteroids, alone or in get of <0.30 mmol/l may be considered in patients with
combination, are recommended and have similar effi- tophaceous disease139,140. Urate-​lowering therapy can be
cacy140,149. The choice of anti-​inflammatory agent is safely initiated during a gout flare158,159. Gout flares may
based on the individual’s comorbidities and concomi- be triggered by initiation of urate-lowering therapy at full
tant medications. An action plan and supply of medica- dose, and starting with a low dose and gradually titrating
tion should be available so that patients can commence the dose over several months can reduce the risk of gout
therapy as soon as possible after flare onset. flares during the initiation phase160. After commencing
NSAIDs and corticosteroids should be used for the therapy, serum urate concentration should be moni-
shortest time required to effectively treat the gout flare. tored with dose titration of therapy until the target urate
Typical doses of NSAIDs for treatment of a gout concentration is achieved. Once the target urate level is
flare are 500 mg naproxen twice daily, 50 mg indometh- achieved, less frequent monitoring should continue to
acin three times daily, or 120 mg etoricoxib daily150,151. ensure that the target concentration is maintained139,140.
The typical dose of the corticosteroid prednisone is An increasing number of urate-​lowering drugs are
30–35 mg daily151. Low-​dose colchicine (1.2 mg followed available and can be grouped based on their mechanism
by 0.6 mg after 1 h) is preferred over a higher dose due to of action, which includes inhibition of urate formation
substantially fewer adverse effects but similar benefit152. (such as xanthine oxidase inhibitors), inhibitors of renal
The colchicine dose may need to be further reduced urate transporters (uricosuric drugs) and enzymes
in patients with impaired kidney function or in those that metabolize urate (such as the recombinant uricase
receiving inhibitors of cytochrome P450 3A4 (CYP3A4) pegloticase) (Fig. 8).
or P-​glycoprotein (also known as MDR1). Canakinumab
is also an effective therapy for gout flares91,153 but its use is Xanthine oxidase inhibitors. Xanthine oxidase inhi-
limited by cost and it is currently reserved for patients in bition is considered the first-​line therapy for gout;
whom other options are ineffective or contraindicated. currently available inhibitors include allopurinol and
A randomized controlled trial has also shown that the febuxostat. Allopurinol is rapidly metabolized to its
IL-1 receptor antagonist anakinra is non-​inferior to active metabolite oxypurinol, which is excreted primar-
NSAIDs, colchicine and prednisone90. ily by the kidneys. Consequently, the rate of elimination
declines as kidney function deteriorates. In compari-
Urate-​lowering therapy son, febuxostat is metabolized predominantly by the
Sustained reduction of serum urate level is essential for liver and its elimination is less dependent on kidney
the long-term management of gout. Long-term reduction function.
of serum urate to sub-saturation levels (<0.36 mmol/l, Febuxostat had a superior urate-​lowering effect com-
6 mg/dl) with urate-lowering therapy leads to dissolu- pared with fixed or limited allopurinol dose (maximum
tion of MSU crystals154, prevention of progressive joint 200–300 mg/day) in clinical trials161,162. However, an
damage154, suppression of gout flares155 and improved allopurinol dose titration using a treat-​to-target urate
function155. For these reasons, all major rheumatology concentration strategy with a maximum dose of 900 mg
society guidelines139,140,156 (but not the American College daily is effective in most individuals with gout163,164.
of Physicians157) recommend periodic measurement of A number of variables can influence the allopurinol dose

10 | Article citation ID: (2019) 5:69 www.nature.com/nrdp

0123456789();
Primer

that is required to achieve the serum urate target, includ- with kidney impairment, although this study was not
ing baseline serum urate concentration, body weight, powered to detect AHS163,164.
kidney function and ABCG2 genotype165,166. Although based on limited data, there have not been
Potential serious adverse effects are a concern with the same concerns about the use of febuxostat in individ-
both of these agents. For allopurinol, the greatest concern uals with moderate-​to-severe kidney impairment171,172.
is the rare, potentially fatal allopurinol hypersensitivity However, febuxostat use is a concern in patients with
syndrome (AHS). AHS typically occurs in the first few cardio­vascular disease, as clinical trials conducted as
months after commencing allopurinol treatment, and part of the FDA regulatory process showed a higher num-
various risk factors have been identified, including higher ber of cardiovascular events with febuxostat than with
starting dose of allopurinol, presence of the variant allele placebo. The Cardiovascular Safety of Febuxostat and
HLA-​B*5801, kidney function impairment and con- Allopurinol in Patients with Gout and Cardiovascular
comitant use of diuretics167,168. HLA-​B*5801 is strongly Morbidities (CARES) trial was a large randomized, con-
associated with severe cutaneous adverse reactions trolled trial in 6,190 individuals with gout and cardio-
(SCAR) with allopurinol; this allele is more common vascular disease. Participants were randomly assigned
in some Asian populations, particularly Chinese, and to febuxostat therapy, starting at a low daily dose and
testing for HLA-​ B *5801 in high-​r isk populations increasing the dose if urate remained >0.36 mmol/l
and avoidance of allopurinol in HLA-​B*5801 carriers can (6 mg/dl) after 2 weeks, or to allopurinol therapy, start-
substantially reduce the risk of AHS169. Both the EULAR ing at a low daily dose (starting level dependent on
and the ACR recommend a maximum allopurinol start- kidney function) and increasing each month over the
ing dose of 100 mg/day, or 50 mg/day in individuals with first 10 weeks to a high dose (lower in those with creati-
moderate-​to-severe kidney function impairment139,140. nine clearance 30–60 ml/min). Although febuxostat did
However, whereas the ACR suggests that the allopurinol not increase risk of the primary composite end point
dose can be gradually escalated, even in those with kid- of cardiovascular death, non-​fatal myocardial infarc-
ney impairment, the EULAR recommends that the dose tion, non-​fatal stroke or urgent revascularization for
should be limited in individuals with impaired kidney unstable angina (HR 1.03, 95% CI 0.87–1.23) compared
function, based on concerns about AHS in this group. with allopurinol, the pre-​specified secondary analy-
There is evidence that in individuals who develop AHS, ses revealed an increased risk of cardiovascular death
mortality is higher in those with poor kidney function170. (HR 1.34, 95% CI 1.03–1.73) and all-​cause mortality
However, another study found that in individuals who (HR 1.22, 95% CI 1.01–1.47) with febuxostat173. The
tolerate allopurinol, the dose can be gradually increased strengths and limitations of this study have been dis-
without an increase in adverse effects, even in those cussed elsewhere174 but, importantly, the absence of a
placebo arm or a non-​xanthine oxidase inhibitor urate-​
lowering therapy arm leaves uncertainty about whether
Box 2 | General principles of gout management allopurinol has a protective effect or febuxostat has a
detri­mental effect. Despite its limitations, the CARES
General principles of gout management have been established based on the 2012
study provided evidence that febuxostat may not be an
American College of Rheumatology (ACR) guidelines139,149 and the 2016 European
League Against Rheumatism (EULAR) guidelines140. appropriate first-​line urate-​lowering therapy in patients
with cardiovascular disease. The FDA has issued a black
Treatment of gout flare box warning for febuxostat based on these results, and
NSAIDs, corticosteroids or colchicine. If these anti-​inflammatory drugs are ineffective
also limited the approved use of febuxostat to patients
or not tolerated, use IL-1 inhibitors if available.
who are not treated effectively or experience severe
Urate-lowering therapy adverse effects with allopurinol.
Indications
Established diagnosis of gout and one of the following: tophi, frequent gout flares
Uricosuric drugs. Uricosuric agents are considered
(two or more per year), CKD stage 2 or worse, or past urolithiasis.
second-line urate-lowering therapies for gout. Prob­
Target serum urate concentrations:
enecid has a moderate urate-lowering effect, even in
• <0.36 mmol/l (6 mg/dl) those with an estimated glomerular filtration rate of
• <0.30 mmol/l (5 mg/dl) for severe or tophaceous disease <50 ml/min/1.73 m2 (ref.175). Benzbromarone is a more
Serum urate monitoring: potent uricosuric agent but is not widely available due
• Monthly until at target serum urate concentration to concerns about hepatotoxicity. The URAT1 inhibitor
• 6-monthly to ensure maintenance of target lesinurad provides additional urate lowering when used
in combination with allopurinol176,177 or febuxostat178.
Anti-​inflammatory prophylaxis during initiation of urate-​lowering therapy
Renal function must be monitored with lesinurad use,
First-​line: low-​dose colchicine or NSAIDs. If these anti-​inflammatory drugs are
and lesinurad must be co-​prescribed with a xanthine
ineffective or not tolerated, consider low-​dose corticosteroids (no clinical trial data
available). Unapproved: canakinumab. oxidase inhibitor owing to potential adverse effects on
the kidneys. Lesinurad is no longer marketed in the USA
Comorbidity screening but is available in Europe and some other countries.
Type 2 diabetes mellitus, cardiovascular disease, hypertension, hyperlipidaemia,
CKD, obesity and obstructive sleep apnoea.
Recombinant uricase. The pegylated recombinant
Patient and health-​care provider education uricase pegloticase is considered a third-​line urate-​
Rationale for long-term urate-​lowering therapy, risk of flares during initiation of urate- lowering option, which is reserved for patients with
lowering therapy, action plan for flare management and advice on healthy lifestyle.
severe gout in whom the target urate concentration has


NATURE REvIEwS | DISeASe PrIMerS | Article citation ID: (2019) 5:69 11

0123456789();
Primer

Nucleotides (IMP, GMP, AMP) causal role of hyperuricaemia and gout in most of these
comorbidities183. Nevertheless, detection and manage-
Inosine
ment of these conditions have important long-​term
health implications for individuals with gout, notably
the higher risk of death in patients with gout144.
Hypoxanthine
The presence of these comorbidities and the medi-
Inhibit urate production Xanthine cations that are used to manage them affect the choice
• Allopurinol oxidase Normalize renal urate
• Febuxostat of therapeutic options for management of gout flares.
excretion
Xanthine • Probenecid For example, NSAIDs should be avoided in patients with
• Lesinurad CKD and those who are receiving anti-​coagulants, and
Xanthine • Benzbromarone high-​dose prednisone may lead to weight gain, hyper-
oxidase
tension and poor control of blood sugar levels. The risk
Urate of adverse events from colchicine treatment is increased
Metabolize urate in patients with CKD and in those taking inhibitors of
• Pegloticase
CYP3A4 or P-​glycoprotein, and colchicine doses should
Allantoin Renal excretion be reduced in these patients184.
In addition, comorbidities may influence the choice of
Fig. 8 | Mechanism of action of urate-​lowering drugs currently used for gout urate-​lowering therapy. Patients with CKD also require
treatment. Various agents can be used to block urate production (xanthine oxidase
a lower starting dose of allopurinol, and more gradual
inhibitors, such as allopurinol and febuxostat), increase the renal excretion of
urate (uricosuric agents, such as probenecid, lesinurad and benzbromarone) or dose escalation once allopurinol is initiated. Patients
metabolize urate to soluble allantoin (for example, the recombinant pegylated with a prior history of kidney stones should avoid uri-
uricase pegloticase). cosuric agents, and uricosuric agents may be less effec-
tive in patients with CKD. The use of febuxostat in
patients with established cardiovascular disease requires
not been achieved or who are unable to tolerate other careful consideration173 — in this population, allopu­
therapeutic options. Pegloticase results in a rapid, pro- rinol rather than febuxostat is the first-line xanthine
found reduction in serum urate levels, which is asso- oxidase inhibitor.
ciated with an increase in gout flares in the short term
but dissolution of tophi and improvements in pain and Lifestyle modification
QOL in the longer term179. As with other biological Lifestyle modification has been considered an impor-
therapies, pegloti­case can be associated with infusion tant aspect of gout management185. Although the major
reactions179,180 and the development of antibodies that rheumatology society guidelines recommend lifestyle
are associated with loss of response and increased risk modifications (such as weight loss, reduced intake of
of infusion reactions181. purine-​rich foods, avoidance of alcohol and fructose-​
containing beverages and increased consumption of
Anti-​inflammatory prophylaxis dairy products)139,140, there is limited clinical trial evi-
Initiating urate-​lowering therapy or increasing the dence for benefit from lifestyle modification in people
dose can be associated with gout flares179,182. Therefore, with gout. Weight loss for overweight or obese patients
patients should receive anti-​inflammatory prophylaxis with gout may have beneficial effects on serum urate lev-
against gout flares when commencing urate-​lowering els but the quality of evidence is poor186. Dietary modi­
therapy140,149. Colchicine (0.5 mg once or twice daily) or fication is extremely difficult to maintain in the long term
low-​dose NSAIDs are recommended first-​line agents for and has little effect on serum urate levels187. Similarly,
anti-​inflammatory prophylaxis. Low-​dose corticosteroid there is insufficient evidence to recommend com­
treatment is reserved for patients in whom colchicine or plementary therapies, such as tart cherry concentrate
NSAIDs are contraindicated, ineffective or not tolerated. and vitamin C.
IL-1 inhibitors are also effective in gout flare prophy-
laxis89,149; although canakinumab is approved outside Quality of life
the USA for treatment of gout flares, IL-1 inhibitors Gout has a substantial economic impact on patients,
are not approved for anti-​inflammatory prophylaxis. employers and society, leading to reduced productivity
Gradual febuxostat dose escalation from a low dose to at work and home and increased health-​care utiliza-
a high dose or fixed high-​dose febuxostat with colchi- tion188,189. Health-​related QOL (HRQOL) refers to the
cine prophylaxis have both been reported to effectively physical, mental, and social well-​being of the patient and
reduce gout flares compared with fixed high-​dose febux- is shaped by the patient’s perception and expectations of
ostat without prophylaxis, suggesting that the ‘start low, his or her health. A number of measures have been used
go slow’ dose escalation strategy may reduce the need for to report HRQOL in gout, most often generic instru-
anti-​inflammatory prophylaxis160. ments, such as the 36-item Short Form Health Survey
(SF-36), and gout-​specific instruments, such as the Gout
Comorbidities Impact Scale (GIS).
Gout is associated with a number of important comor- In patients with gout, HRQOL can be affected
bidities, including hypertension, hyperlipidaemia, by gout-​specific factors and by comorbidities190,191.
cardiovascular disease, obesity and CKD, although. Physical function domains, such as ‘physical function-
Mendelian randomization studies do not support a ing’, ‘role physical’, and ‘bodily pain’, are generally more

12 | Article citation ID: (2019) 5:69 www.nature.com/nrdp

0123456789();
Primer

affected than other HRQOL domains, such as mental prevalence and increasing incidence of gout necessitate
and emotional domains192. Gout-​related factors that population-​wide strategies to prevent the disease and
affect HRQOL include current gout flare, frequent ensure effective treatment for those with the illness.
gout flares, tophi, oligoarticular or polyarticular flares, Addressing environmental risk factors for gout,
and pain191,193. Gout comorbidities that affect HRQOL including high fructose intake and increasing body mass
include anxiety and depression, diabetes mellitus, stroke, index, may have an important role in prevention of gout
kidney failure, myocardial infarction and obesity. at the population level. However, there is little evidence
On average, the incremental annual cost of gout care that dietary management alone can lead to durable or
is more than US$3,000 per patient, with substantially clinically meaningful reductions in serum urate con-
higher costs for patients who experience frequent gout centrations once MSU crystal deposition has occurred
flares or who have tophi194. Given the high prevalence and clinically evident gout is present. Avoiding specific
of gout (that affects ~4% of adults in the USA), the total foods that trigger gout flares might lead to clinical ben-
annual cost of gout care in the USA is estimated to be efit in patients with gout28, possibly through effects on
tens of billions of US dollars195. pro-​inflammatory pathways rather than alterations in
Analysis of the 2012 and 2013 US National Health and serum urate levels96. Clinical trials of dietary interven-
Wellness Survey showed that patients with uncontrolled tions in gout have not shown major benefits in lower-
gout (defined as serum urate >0.36 mmol/l (6 mg/dl) or ing serum urate187,198,199, and excessive focus on dietary
the occurrence of one or more gout flares in the past year) management can detract from education approaches
had more emergency department visits and had higher to promote understanding of gout as a chronic disease
presenteeism, overall work impairment and activity of MSU crystal deposition and the importance of long-​
impairment than patients with controlled gout (defined term urate-​lowering therapy to prevent recurrent flares
as serum urate ≤0.36 mmol/l (6 mg/dl) and no gout and the sequelae of tophaceous gout200.
flares in the past year) or individuals without gout196. Many rheumatology societies have published gout
Importantly, the (direct and indirect) costs of gout were management guidelines139,140,149,156. These guidelines
primarily related to the costs of uncontrolled gout, have minor variations, but all endorse long-​term urate-​
and after adjusting for confounding variables including lowering therapy for patients with recurrent gout flares
comorbidities, there was no significant difference in and tophi, using a treat-​to-target serum urate concen-
total costs between patients with controlled gout and tration approach. However, studies from around the
individuals without gout, indicating the potential savings world have consistently shown low rates of initiation and
from controlling gout196. continuation of urate-​lowering therapy in patients with
Only a few clinical trials in patients with gout have gout, and when urate-​lowering therapy is prescribed,
investigated changes in HRQOL as an outcome measure. insufficient dose titration to achieve sub-​saturating
In a randomized controlled trial of pegloticase treat- serum urate targets201–204. In addition to the burden of
ment in patients with severe gout, baseline physical com­ severe joint pain and disability, inadequate prescription
ponent summary (PCS) scores on the SF-36 were more of urate-​lowering therapy in patients with poorly con-
than 1.5 standard deviations below US normative values. trolled gout leads to unnecessary exposure to the adverse
At week 25, mean changes from baseline PCS scores effects of NSAIDs and corticosteroids205.
were statistically significant and exceeded minimum Dominant perceptions and beliefs about the dis-
clinically important differences (MCID) in the biweekly ease are a challenge for effective gout management.
pegloticase group, compared with little change in the Historical narratives about gout as a self-​inflicted dis-
placebo group. Statistically significant improvements ease of excessive alcohol consumption and over-​eating
exceeding the MCID were observed in all SF-36 domains persist in contemporary depictions of the illness206, and
in the biweekly pegloticase group, with the exception can contribute to negative views about patients with
of the ‘role emotional’ and ‘mental health’ domains197. gout, patient embarrassment and excessive focus on
A 2-year study showed improvements in both generic unproven dietary solutions200. Gout is experienced as
and gout-​s pecific HRQOL scores with nurse-​led an intermittently flaring condition, and patients and
care according to the EULAR and British Society for health-​care providers may not understand that MSU
Rheumatology (BSR) gout management guidelines, com- crystal deposition is present even during asymptomatic
pared with usual care by a general practitioner155. The intercritical periods between flares, providing a rationale
SF-36 physical component had significantly improved for long-​term urate-​lowering therapy. Flares soon after
(P < 0.0001) in the nurse-​led care group at the end of initiating urate-​lowering therapy may lead to negative
2 years, but did not change in the usual-​care group, with patient experience of this therapy200, especially if therapy
no difference in the mental component scores. The GIS is initiated with high doses without anti-​inflammatory
scores were also better in the nurse-​led group that in the prophylaxis160. The presence of other serious comorbid-
usual-​care group at 1 year and 2 years. ities, such as diabetes, cardiovascular disease and CKD,
may take priority in health-​care interactions, particularly
Outlook given the limited time available for patient interactions
Urate-​lowering agents, such as allopurinol and probene- with a primary care physician200.
cid, have been available for more than half a century, A number of innovative strategies in primary care have
and new agents have been approved in the past dec- been described to improve the quality of gout care, includ-
ade. However, gout continues to have a major effect on ing quality-improvement disease-management pro-
patients, their families and their communities. The high grammes within primary care207 and pharmacist-based


NATURE REvIEwS | DISeASe PrIMerS | Article citation ID: (2019) 5:69 13

0123456789();
Primer

interactions using a virtual gout management clinic208 concentrations <36 mmol/l at 2 years than those receiv-
or automated telephone technology209. Although these ing usual care (95% versus 30%), and improvements in
interventions have led to substantial improvements in the incidence of gout flares, tophi, physical function and
quality of care, the majority of patients in these pro- scores on the patient-reported GIS questionnaire were
grammes did not achieve the target serum urate concen- greater with nurse-led care. The nurse-led intervention
tration. The most impressive results have been found in was cost-effective in the short-term and potentially
a large primary-​care-based randomized controlled trial cost-saving in the long-term.
that compared nurse-​led care (according to EULAR and The study outcomes reported with nurse-led care
BSR gout management guidelines) with usual care by provide a benchmark for gout treatment and demon-
a general practitioner155. The comprehensive nurse-​led strate the clinical and cost benefits that can be achieved
intervention included holistic assessment, discussion of with high-​quality long-​term management. The next
illness perceptions, full information about gout (nature, challenge is implementation of effective and durable
causes, associations, consequences and treatment community-​b ased gout management programmes;
options) and shared decision-​making. Compared with implementation will require development and funding of
usual care, nurse-​led care was associated with high uptake a trained workforce to deliver the effective intervention
of urate-​lowering therapy and adherence to therapy in within different health-​care settings worldwide.
more than 95% of study participants. A greater propor-
tion of patients receiving nurse-led care had serum urate Published online xx xx xxxx

1. Chen-​Xu, M., Yokose, C., Rai, S. K., Pillinger, M. H. population study. Ann. Rheum. Dis. 74, 661–667 in the Normative Aging Study. Am. J. Med. 82,
& Choi, H. K. Contemporary prevalence of gout and (2015). 421–426 (1987).
hyperuricemia in the United States and decadal 16. Zhang, Y. et al. Alcohol consumption as a trigger of 34. Kapetanovic, M. C. et al. The risk of clinically
trends: the National Health and Nutrition Examination recurrent gout attacks. Am. J. Med. 119, 800. diagnosed gout by serum urate levels: results from
Survey 2007–2016. Arthritis Rheumatol. 71, e11–800.e16 (2006). 30 years follow-​up of the Malmo Preventive Project
991–999 (2019). 17. Kuo, C. F. et al. Epidemiology and management of Cohort in Southern Sweden. Arthritis Res. Ther. 20,
This article describes the contemporary gout in Taiwan: a nationwide population study. 190 (2018).
epidemiology of gout and hyperuricaemia in Arthritis Res. Ther. 17, 13 (2015). 35. Dalbeth, N. et al. Urate crystal deposition in
the USA. 18. Roddy, E. & Doherty, M. Epidemiology of gout. asymptomatic hyperuricaemia and symptomatic gout:
2. Faires, J. S. & McCarty, D. J. Acute arthritis in man Arthritis Res. Ther. 12, 223 (2010). a dual energy CT study. Ann. Rheum. Dis. 74,
and dog after intrasynovial injection of sodium urate 19. Arromdee, E., Michet, C. J., Crowson, C. S., 908–911 (2015).
crystals. Lancet 280, 682–685 (1962). O’Fallon, W. M. & Gabriel, S. E. Epidemiology of gout: 36. De Miguel, E. et al. Diagnosis of gout in patients with
This paper proved that monosodium urate crystals is the incidence rising? J. Rheumatol. 29, 2403–2406 asymptomatic hyperuricaemia: a pilot ultrasound
cause the gout flare. (2002). study. Ann. Rheum. Dis. 71, 157–158 (2012).
3. Loeb, J. N. The influence of temperature on the 20. Nakayama, A. et al. GWAS of clinically defined gout 37. Dalbeth, N. et al. Cellular characterization of the gouty
solubility of monosodium urate. Arthritis Rheum. 15, and subtypes identifies multiple susceptibility loci that tophus: a quantitative analysis. Arthritis Rheum. 62,
189–192 (1972). include urate transporter genes. Ann. Rheum. Dis. 76, 1549–1556 (2010).
4. Taylor, W. J. et al. Study for updated gout classification 869–877 (2017). 38. Wu, X. W., Lee, C. C., Muzny, D. M. & Caskey, C. T.
criteria: identification of features to classify gout. 21. Lim, S. Y. et al. Trends in gout and rheumatoid arthritis Urate oxidase: primary structure and evolutionary
Arthritis Care Res. (Hoboken) 67, 1304–1315 (2015). hospitalizations in the United States, 1993-2011. implications. Proc. Natl Acad. Sci. USA 86,
This study describes the major clinical features of JAMA 315, 2345–2347 (2016). 9412–9416 (1989).
gout in a large multicentre study, using monosodium 22. Robinson, P. C., Merriman, T. R., Herbison, P. & 39. Ames, B. N., Cathcart, R., Schwiers, E. & Hochstein, P.
urate crystal identification as the gold standard. Highton, J. Hospital admissions associated with gout Uric acid provides an antioxidant defense in humans
5. Gutman, A. B. The past four decades of progress in the and their comorbidities in New Zealand and England against oxidant- and radical-​caused aging and cancer:
knowledge of gout, with an assessment of the present 1999–2009. Rheumatology 52, 118–126 (2013). a hypothesis. Proc. Natl Acad. Sci. USA 78,
status. Arthritis Rheum. 16, 431–445 (1973). 23. Dehlin, M. & Jacobsson, L. T. H. Trends in gout 6858–6862 (1981).
6. Neogi, T. et al. 2015 gout classification criteria: an hospitalization in Sweden. J. Rheumatol. 45, 40. Shi, Y., Evans, J. E. & Rock, K. L. Molecular
American College of Rheumatology/European League 145–146 (2018). identification of a danger signal that alerts the immune
Against Rheumatism collaborative initiative. 24. Edwards, N. L. Quality of care in patients with gout: system to dying cells. Nature 425, 516–521 (2003).
Ann. Rheum. Dis. 74, 1789–1798 (2015). why is management suboptimal and what can be done 41. Watanabe, S. et al. Uric acid, hominoid evolution, and
This paper describes the 2015 ACR/EULAR gout about it? Curr. Rheumatol. Rep. 13, 154–159 (2011). the pathogenesis of salt-​sensitivity. Hypertension 40,
classification criteria. 25. Sarawate, C. A. et al. Gout medication treatment 355–360 (2002).
7. Bursill, D. et al. Gout, Hyperuricemia, and Crystal-​ patterns and adherence to standards of care from 42. Choi, H. K., Liu, S. & Curhan, G. Intake of purine-​rich
Associated Disease Network consensus statement a managed care perspective. Mayo Clin. Proc. 81, foods, protein, and dairy products and relationship to
regarding labels and definitions for disease elements 925–934 (2006). serum levels of uric acid: the Third National Health
in gout. Arthritis Care Res. (Hoboken) 71, 427–434 26. Shiozawa, A., Szabo, S. M., Bolzani, A., Cheung, A. & and Nutrition Examination Survey. Arthritis Rheum.
(2019). Choi, H. K. Serum uric acid and the risk of incident and 52, 283–289 (2005).
8. Kuo, C. F., Grainge, M. J., Zhang, W. & Doherty, M. recurrent gout: a systematic review. J. Rheumatol. 44, 43. Choi, H. K., Atkinson, K., Karlson, E. W., Willett, W. &
Global epidemiology of gout: prevalence, incidence 388–396 (2017). Curhan, G. Purine-​rich foods, dairy and protein intake,
and risk factors. Nat. Rev. Rheumatol. 11, 649–662 27. Rothenbacher, D., Primatesta, P., Ferreira, A., and the risk of gout in men. N. Engl. J. Med. 350,
(2015). Cea-​Soriano, L. & Rodriguez, L. A. Frequency and risk 1093–1103 (2004).
9. Zhu, Y., Pandya, B. J. & Choi, H. K. Prevalence of gout factors of gout flares in a large population-​based 44. Stirpe, F. et al. Fructose-​induced hyperuricaemia.
and hyperuricemia in the US general population: the cohort of incident gout. Rheumatology 50, 973–981 Lancet 2, 1310–1311 (1970).
National Health and Nutrition Examination Survey (2011). 45. Bode, C., Schumacher, H., Goebell, H., Zelder, O. &
2007–2008. Arthritis Rheum. 63, 3136–3141 (2011). 28. Zhang, Y. et al. Purine-​rich foods intake and recurrent Pelzel, H. Fructose induced depletion of liver adenine
10. Lawrence, R. C. et al. Estimates of the prevalence of gout attacks. Ann. Rheum. Dis. 71, 1448–1453 nucleotides in man. Hormone Metab. Res. 3,
selected arthritic and musculoskeletal diseases in the (2012). 289–290 (1971).
United States. J. Rheumatol. 16, 427–441 (1989). 29. Dubreuil, M. et al. Increased risk of recurrent gout 46. Choi, J. W., Ford, E. S., Gao, X. & Choi, H. K. Sugar-​
11. Choi, H. K. & Curhan, G. Soft drinks, fructose attacks with hospitalization. Am. J. Med. 126, sweetened soft drinks, diet soft drinks, and serum uric
consumption, and the risk of gout in men: prospective 1138–1141.e1 (2013). acid level: the Third National Health and Nutrition
cohort study. BMJ 336, 309–312 (2008). 30. Hak, A. E., Curhan, G. C., Grodstein, F. & Choi, H. K. Examination Survey. Arthritis Rheum. 59, 109–116
12. Choi, H. K., Willett, W. & Curhan, G. Fructose-​rich Menopause, postmenopausal hormone use and risk of (2008).
beverages and risk of gout in women. JAMA 304, incident gout. Ann. Rheum. Dis. 69, 1305–1309 47. Faller, J. & Fox, I. H. Ethanol-​induced hyperuricemia:
2270–2278 (2010). (2010). evidence for increased urate production by activation
13. Currie, W. J. Prevalence and incidence of the diagnosis 31. Puig, J. G. et al. Female gout. Clinical spectrum and of adenine nucleotide turnover. N. Engl. J. Med. 307,
of gout in Great Britain. Ann. Rheum. Dis. 38, uric acid metabolism. Arch. Intern. Med. 151, 1598–1602 (1982).
101–106 (1979). 726–732 (1991). 48. Puig, J. G. & Fox, I. H. Ethanol-​induced activation
14. Harris, C. M., Lloyd, D. C. & Lewis, J. The prevalence 32. Dalbeth, N. et al. Relationship between serum urate of adenine nucleotide turnover. Evidence for a role of
and prophylaxis of gout in England. J. Clin. Epidemiol. concentration and clinically evident incident gout: an acetate. J. Clin. Invest. 74, 936–941 (1984).
48, 1153–1158 (1995). individual participant data analysis. Ann. Rheum. Dis. 49. Choi, H. K. & Curhan, G. Coffee, tea, and caffeine
15. Kuo, C. F., Grainge, M. J., Mallen, C., Zhang, W. & 77, 1048–1052 (2018). consumption and serum uric acid level: the Third
Doherty, M. Rising burden of gout in the UK but 33. Campion, E. W., Glynn, R. J. & DeLabry, L. O. National Health and Nutrition Examination Survey.
continuing suboptimal management: a nationwide Asymptomatic hyperuricemia. Risks and consequences Arthritis Rheum. 57, 816–821 (2007).

14 | Article citation ID: (2019) 5:69 www.nature.com/nrdp

0123456789();
Primer

50. Gao, X., Curhan, G., Forman, J. P., Ascherio, A. 74. Phipps-​Green, A. J. et al. A strong role for the ABCG2 96. Joosten, L. A. et al. Engagement of fatty acids with
& Choi, H. K. Vitamin C intake and serum uric acid gene in susceptibility to gout in New Zealand Pacific Toll-​like receptor 2 drives interleukin-1β production via
concentration in men. J. Rheumatol. 35, 1853–1858 Island and Caucasian, but not Maori, case and control the ASC/caspase 1 pathway in monosodium urate
(2008). sample sets. Hum. Mol. Genet. 19, 4813–4819 monohydrate crystal-​induced gouty arthritis. Arthritis
51. Choi, H. K., Soriano, L. C., Zhang, Y. & Rodriguez, L. A. (2010). Rheum. 62, 3237–3248 (2010).
Antihypertensive drugs and risk of incident gout 75. Higashino, T. et al. Multiple common and rare variants 97. Crisan, T. O. et al. Soluble uric acid primes TLR-​induced
among patients with hypertension: population based of ABCG2 cause gout. RMD Open 3, e000464 proinflammatory cytokine production by human
case-​control study. BMJ 344, d8190 (2012). (2017). primary cells via inhibition of IL-1Ra. Ann. Rheum. Dis.
52. Zhu, Y., Pandya, B. J. & Choi, H. K. Comorbidities of 76. Stiburkova, B. et al. Functional non-​synonymous 75, 755–762 (2016).
gout and hyperuricemia in the US general population: variants of ABCG2 and gout risk. Rheumatology 56, 98. Mylona, E. E. et al. Enhanced interleukin-1β
NHANES 2007-2008. Am. J. Med. 125, 679–687.e1 1982–1992 (2017). production of PBMCs from patients with gout after
(2012). 77. Dehghan, A. et al. Association of three genetic loci stimulation with Toll-​like receptor-2 ligands and urate
53. Nakayama, A. et al. Common dysfunctional variants of with uric acid concentration and risk of gout: crystals. Arthritis Res. Ther. 14, R158 (2012).
ABCG2 have stronger impact on hyperuricemia a genome-​wide association study. Lancet 372, 99. Crisan, T. O. et al. Uric acid priming in human
progression than typical environmental risk factors. 1953–1961 (2008). monocytes is driven by the AKT-​PRAS40 autophagy
Sci. Rep. 4, 5227 (2014). 78. Ichida, K. et al. Decreased extra-​renal urate excretion pathway. Proc. Natl Acad. Sci. USA 114, 5485–5490
54. Major, T. J., Topless, R. K., Dalbeth, N. & Merriman, T. R. is a common cause of hyperuricemia. Nat. Commun. 3, (2017).
Evaluation of the diet wide contribution to serum urate 764 (2012). 100. Chen, Y. H. et al. Spontaneous resolution of acute
levels: meta-​analysis of population based cohorts. This paper describes decreased extra-​renal urate gouty arthritis is associated with rapid induction of the
BMJ 363, k3951 (2018). excretion caused by ABCG2 dysfunction as a new anti-​inflammatory factors TGFβ1, IL-10 and soluble
55. Perez-​Ruiz, F., Calabozo, M., Erauskin, G. G., Ruibal, A. concept in the pathogenesis of hyperuricaemia. TNF receptors and the intracellular cytokine negative
& Herrero-​Beites, A. M. Renal underexcretion of uric 79. Becker, M. A. in The Metabolic and Molecular Bases regulators CIS and SOCS3. Ann. Rheum. Dis. 70,
acid is present in patients with apparent high urinary of Inherited Disease (eds Scriver, C. R., Childs, B., 1655–1663 (2011).
uric acid output. Arthritis Rheum. 47, 610–613 (2002). Kinzler, K. W., & Vogelstein, B.) Ch. 106, 2513–2535 101. Yagnik, D. R. et al. Macrophage release of transforming
56. Kottgen, A. et al. Genome-​wide association analyses (McGraw-​Hill, 2001). growth factor beta1 during resolution of monosodium
identify 18 new loci associated with serum urate 80. Wortmann, R. L. in Harrison’s Principles of Internal urate monohydrate crystal-​induced inflammation.
concentrations. Nat. Genet. 45, 145–154 (2013). Medicine (eds Fauci, A. S. et al.) Ch. 353, 2444–2449 Arthritis Rheum. 50, 2273–2280 (2004).
This landmark paper describes a large GWAS to (McGraw-​Hill, 2008). 102. Barden, A. E. et al. Specialised pro-​resolving
identify loci associated with serum urate levels and 81. Mandel, N. S. & Mandel, G. S. Monosodium urate mediators of inflammation in inflammatory arthritis.
gout, emphasizing the importance of urate monohydrate, the gout culprit. J. Am. Chem. Soc. 98, Prostaglandins Leukot. Essent. Fatty Acids 107,
transporters and glucose metabolism in regulation 2319–2323 (1976). 24–29 (2016).
of serum urate concentrations. 82. Dalbeth, N. et al. Tendon involvement in the feet 103. Liu, L. et al. Interleukin 37 limits monosodium urate
57. Nakatochi, M. et al. Genome-​wide meta-​analysis of patients with gout: a dual-​energy CT study. crystal-​induced innate immune responses in human
identifies multiple novel loci associated with serum Ann. Rheum. Dis. 72, 1545–1548 (2013). and murine models of gout. Arthritis Res. Ther. 18,
uric acid levels in Japanese individuals. Commun. Biol. 83. Chhana, A., Lee, G. & Dalbeth, N. Factors influencing 268 (2016).
2, 115 (2019). the crystallization of monosodium urate: a systematic 104. Steiger, S. & Harper, J. L. Neutrophil cannibalism
58. Li, C. et al. Genome-​wide association analysis literature review. BMC Musculoskelet. Disord. 16, triggers transforming growth factor β1 production and
identifies three new risk loci for gout arthritis in Han 296 (2015). self regulation of neutrophil inflammatory function in
Chinese. Nat. Commun. 6, 7041 (2015). 84. Schumacher, H. R. Pathology of the synovial monosodium urate monohydrate crystal-​induced
59. Phipps-​Green, A. J. et al. Twenty-​eight loci that membrane in gout. Light and electron microscopic inflammation in mice. Arthritis Rheum. 65, 815–823
influence serum urate levels: analysis of association studies. Interpretation of crystals in electron (2013).
with gout. Ann. Rheum. Dis. 75, 124–130 (2016). micrographs. Arthritis Rheum. 18, 771–782 (1975). 105. Schauer, C. et al. Aggregated neutrophil extracellular
60. Matsuo, H. et al. Genome-​wide association study of 85. Wood, D. D., Ihrie, E. J., Dinarello, C. A. & Cohen, P. L. traps limit inflammation by degrading cytokines and
clinically defined gout identifies multiple risk loci and Isolation of an interleukin-1-like factor from human chemokines. Nat. Med. 20, 511–517 (2014).
its association with clinical subtypes. Ann. Rheum. Dis. joint effusions. Arthritis Rheum. 26, 975–983 (1983). This article describes the role of aggNETs in
75, 652–659 (2016). 86. Di Giovine, F. S., Malawista, S. E., Nuki, G. & Duff, G. W. resolution of the gout flare and also implicates
61. Enomoto, A. et al. Molecular identification of a renal Interleukin 1 (IL 1) as a mediator of crystal arthritis. aggNETs in development of the tophus.
urate anion exchanger that regulates blood urate Stimulation of T cell and synovial fibroblast 106. Sokoloff, L. The pathology of gout. Metabolism 6,
levels. Nature 417, 447–452 (2002). mitogenesis by urate crystal-​induced IL 1. J. Immunol. 230–243 (1957).
This paper describes the discovery of URAT1, 138, 3213–3218 (1987). 107. Schweyer, S., Hemmerlein, B., Radzun, H. J. & Fayyazi, A.
the major target of uricosuric agents. 87. Martinon, F., Petrilli, V., Mayor, A., Tardivel, A. & Continuous recruitment, co-​expression of tumour
62. Shin, H. J. et al. Interactions of urate transporter Tschopp, J. Gout-​associated uric acid crystals activate necrosis factor-​alpha and matrix metalloproteinases,
URAT1 in human kidney with uricosuric drugs. the NALP3 inflammasome. Nature 440, 237–241 and apoptosis of macrophages in gout tophi. Virchows
Nephrology 16, 156–162 (2011). (2006). Archiv. 437, 534–539 (2000).
63. Miner, J. N. et al. Lesinurad, a novel, oral compound This paper describes the crucial role of activation 108. Dalbeth, N. et al. Mechanisms of bone erosion in gout:
for gout, acts to decrease serum uric acid through of the NLRP3 inflammasome and release of mature a quantitative analysis using plain radiography and
inhibition of urate transporters in the kidney. Arthritis IL-1β in the initiation of the gout flare. computed tomography. Ann. Rheum. Dis. 68,
Res. Ther. 18, 214 (2016). 88. Solomon, D. H. et al. Relationship of interleukin-1β 1290–1295 (2009).
64. Matsuo, H. et al. Mutations in glucose transporter 9 blockade with incident gout and serum uric acid levels: 109. Towiwat, P. et al. Urate crystal deposition and bone
gene SLC2A9 cause renal hypouricemia. Am. J. Hum. exploratory analysis of a randomized controlled trial. erosion in gout: ‘inside-​out’ or ‘outside-​in’? A dual-​
Genet. 83, 744–751 (2008). Ann. Intern. Med. 169, 535–542 (2018). energy computed tomography study. Arthritis Res.
65. Dinour, D. et al. Homozygous SLC2A9 mutations 89. Schlesinger, N. et al. Canakinumab reduces the risk Ther. 18, 208 (2016).
cause severe renal hypouricemia. J. Am. Soc. Nephrol. of acute gouty arthritis flares during initiation of 110. Dalbeth, N. et al. Enhanced osteoclastogenesis in
21, 64–72 (2010). allopurinol treatment: results of a double-​blind, patients with tophaceous gout: urate crystals promote
66. Kawamura, Y. et al. Pathogenic GLUT9 mutations randomised study. Ann. Rheum. Dis. 70, 1264–1271 osteoclast development through interactions with
causing renal hypouricemia type 2 (RHUC2). (2011). stromal cells. Arthritis Rheum. 58, 1854–1865
Nucleosides Nucleotides Nucleic Acids 30, 90. Janssen, C. A. et al. Anakinra for the treatment of (2008).
1105–1111 (2011). acute gout flares: a randomized, double-​blind, 111. Chhana, A. et al. Monosodium urate monohydrate
67. Chiba, T. et al. NPT1/SLC17A1 is a renal urate placebo-​controlled, active-​comparator, non-​inferiority crystals inhibit osteoblast viability and function:
exporter in humans and its common gain-​of-function trial. Rheumatology 58, 1344–1352 (2019). implications for development of bone erosion in gout.
variant decreases the risk of renal underexcretion 91. Schlesinger, N. et al. Canakinumab for acute gouty Ann. Rheum. Dis. 70, 1684–1691 (2011).
gout. Arthritis Rheumatol. 67, 281–287 (2015). arthritis in patients with limited treatment options: 112. Chhana, A. et al. Monosodium urate crystals reduce
68. Hollis-​Moffatt, J. E. et al. The renal urate transporter results from two randomised, multicentre, active-​ osteocyte viability and indirectly promote a shift
SLC17A1 locus: confirmation of association with gout. controlled, double-​blind trials and their initial in osteocyte function towards a proinflammatory
Arthritis Res. Ther. 14, R92 (2012). extensions. Ann. Rheum. Dis. 71, 1839–1848 (2012). and proresorptive state. Arthritis Res. Ther. 20, 208
69. Woodward, O. M. et al. Identification of a urate 92. Martinon, F., Burns, K. & Tschopp, J. The (2018).
transporter, ABCG2, with a common functional inflammasome: a molecular platform triggering 113. Janssens, H. J. et al. A diagnostic rule for acute gouty
polymorphism causing gout. Proc. Natl Acad. Sci. USA activation of inflammatory caspases and processing of arthritis in primary care without joint fluid analysis.
106, 10338–10342 (2009). proIL-​beta. Mol. Cell 10, 417–426 (2002). Arch. Intern. Med. 170, 1120–1126 (2010).
70. Matsuo, H. et al. Common defects of ABCG2, a high-​ 93. Pascual, E., Batlle-​Gualda, E., Martinez, A., Rosas, J. This article describes a diagnostic rule for gout in
capacity urate exporter, cause gout: a function-​based & Vela, P. Synovial fluid analysis for diagnosis of patients presenting with monoarthritis in clinical
genetic analysis in a Japanese population. Sci. Transl intercritical gout. Ann. Intern. Med. 131, 756–759 practice, which has subsequently been validated
Med. 1, 5ra11 (2009). (1999). in a number of different clinical situations.
71. Matsuo, H. et al. ABCG2 dysfunction causes 94. Mangan, M. S. J. et al. Targeting the NLRP3 114. Vazquez-​Mellado, J. et al. Intradermal tophi in gout:
hyperuricemia due to both renal urate underexcretion inflammasome in inflammatory diseases. Nat. Rev. a case-​control study. J. Rheumatol. 26, 136–140
and renal urate overload. Sci. Rep. 4, 3755 (2014). Drug Discov. 17, 588–606 (2018). (1999).
72. Matsuo, H. et al. Hyperuricemia in acute 95. Giamarellos-​Bourboulis, E. J. et al. Crystals of 115. El-​Zawawy, H. & Mandell, B. F. Crystal-​induced
gastroenteritis is caused by decreased urate excretion monosodium urate monohydrate enhance arthritides in the elderly: an update. Rheum. Dis. Clin.
via ABCG2. Sci. Rep. 6, 31003 (2016). lipopolysaccharide-​induced release of interleukin North Am. 44, 489–499 (2018).
73. Matsuo, H. et al. Common dysfunctional variants 1 beta by mononuclear cells through a caspase 116. Zhang, W. et al. EULAR evidence based
in ABCG2 are a major cause of early-​onset gout. 1-mediated process. Ann. Rheum. Dis. 68, 273–278 recommendations for gout. Part I: Diagnosis. report
Sci. Rep. 3, 2014 (2013). (2009). of a task force of the Standing Committee for


NATURE REvIEwS | DISeASe PrIMerS | Article citation ID: (2019) 5:69 15

0123456789();
Primer

International Clinical Studies Including Therapeutics hyperuricemia. Arthritis Care Res. 64, 1431–1446 160. Yamanaka, H. et al. Stepwise dose increase of
(ESCISIT). Ann. Rheum. Dis. 65, 1301–1311 (2006). (2012). febuxostat is comparable with colchicine prophylaxis
117. Perez-​Ruiz, F., Castillo, E., Chinchilla, S. P. & Herrero-​ 140. Richette, P. et al. 2016 updated EULAR evidence-​ for the prevention of gout flares during the initial
Beites, A. M. Clinical manifestations and diagnosis of based recommendations for the management of gout. phase of urate-​lowering therapy: results from
gout. Rheum. Dis. Clin. North Am. 40, 193–206 Ann. Rheum. Dis. 76, 29–42 (2017). FORTUNE-1, a prospective, multicentre randomised
(2014). 141. Abhishek, A. et al. Monosodium urate monohydrate study. Ann. Rheum. Dis. 77, 270–276 (2018).
118. Neogi, T., Krasnokutsky, S. & Pillinger, M. H. Urate crystal deposits are common in asymptomatic sons of This clinical trial showed that gradual escalation of
and osteoarthritis: evidence for a reciprocal patients with gout: the Sons of Gout Study. Arthritis the febuxostat dose reduces the risk of gout flares
relationship. Joint Bone Spine https://doi.org/ Rheumatol. 70, 1847–1852 (2018). compared with full-​dose febuxostat when starting
10.1016/j.jbspin.2018.11.002 (2018). 142. Howard, R. G. et al. Reproducibility of musculoskeletal urate-​lowering therapy; this study also demonstrated
119. Forbess, L. J. & Fields, T. R. The broad spectrum of ultrasound for determining monosodium urate the efficacy of low-​dose colchicine (0.5 mg daily) in
urate crystal deposition: unusual presentations of deposition: concordance between readers. Arthritis reducing gout flares when initiating urate-​lowering
gouty tophi. Semin. Arthritis Rheum. 42, 146–154 Care Res. 63, 1456–1462 (2011). therapy.
(2012). 143. Pineda, C. et al. Joint and tendon subclinical 161. Becker, M. et al. Febuxostat compared with
120. Roddy, E., Zhang, W. & Doherty, M. Are joints affected involvement suggestive of gouty arthritis in allopurinol in patients with hyperuricaemia and gout.
by gout also affected by osteoarthritis? Ann. Rheum. asymptomatic hyperuricemia: an ultrasound N. Engl. J. Med. 353, 2450–2461 (2005).
Dis. 66, 1374–1377 (2007). controlled study. Arthritis Res. Ther. 13, R4 (2011). 162. Becker, M., Schumacher, H. R., MacDonald, P., Lloyd, E.
121. Perez Ruiz, F., Ruiz Lopez, J. & Herrero Beites, A. M. 144. Fisher, M. C., Rai, S. K., Lu, N., Zhang, Y. & Choi, H. K. & Lademacher, C. Clinical efficacy and safety of
Influence of the natural history of disease on a The unclosing premature mortality gap in gout: successful longterm urate lowering with febuxostat or
previous diagnosis in patients with gout [Spanish]. a general population-​based study. Ann. Rheum. Dis. allopurinol in subjects with gout. J. Rheumatol. 36,
Reumatol. Clin. 5, 248–251 (2009). 76, 1289–1294 (2017). 1273–1282 (2009).
122. Taylor, W. J. et al. Performance of classification criteria 145. Andres, M. et al. Cardiovascular risk of patients with 163. Stamp, L. et al. A randomised controlled trial of the
for gout in early and established disease. Ann. Rheum. gout seen at rheumatology clinics following a efficacy and safety of allopurinol dose escalation to
Dis. 75, 178–182 (2016). structured assessment. Ann. Rheum. Dis. 76, achieve target serum urate in people with gout.
123. Janssens, H. et al. Performance of the 2015 ACR-​ 1263–1268 (2017). Ann, Rheum, Dis. 76, 1522–1528 (2017).
EULAR classification criteria for gout in a primary care 146. Maglio, C. et al. Effects of bariatric surgery on gout This randomized controlled trial showed that
population presenting with monoarthritis. incidence in the Swedish Obese Subjects study: a non-​ allopurinol dose escalation can be used to attain
Rheumatology 56, 1335–1341 (2017). randomised, prospective, controlled intervention trial. a target serum urate concentration in most
124. Kienhorst, L. B., Janssens, H. J., Fransen, J. & Ann. Rheum. Dis. 76, 688–693 (2017). patients with gout.
Janssen, M. The validation of a diagnostic rule for 147. Waldman, B. et al. Effect of fenofibrate on uric acid 164. Stamp, L. et al. Allopurinol dose escalation to achieve
gout without joint fluid analysis: a prospective study. and gout in type 2 diabetes: a post-​hoc analysis of the serum urate below 6 mg/dl: an open label extension
Rheumatology 54, 609–614 (2015). randomised, controlled FIELD study. Lancet Diabetes study. Ann. Rheum. Dis. 76, 2065–2070 (2017).
125. Lee, K. H., Choi, S. T., Lee, S. K., Lee, J. H. & Yoon, B. Y. Endocrinol. 6, 310–318 (2018). 165. Stamp, L. K. et al. How much allopurinol does it take
Application of a novel diagnostic rule in the 148. Kimura, K. et al. Febuxostat therapy for patients with to get to target urate? Comparison of actual dose with
differential diagnosis between acute gouty arthritis stage 3 CKD and asymptomatic hyperuricemia: creatinine clearance-​based dose. Arthritis Res. Ther.
and septic arthritis. J. Korean Med. Sci. 30, 700–704 a randomized trial. Am. J. Kidney Dis. 72, 798–810 20, 255 (2018).
(2015). (2018). 166. Wallace, M. C. et al. Association between ABCG2
126. Perez-​Ruiz, F., Martin, I. & Canteli, B. 149. Khanna, D. et al. 2012 American College of rs2231142 and poor response to allopurinol:
Ultrasonographic measurement of tophi as an Rheumatology guidelines for management of gout. replication and meta-​analysis. Rheumatology 57,
outcome measure for chronic gout. J. Rheumatol. 34, Part 2: therapy and antiinflammatory prophylaxis of 656–660 (2018).
1888–1893 (2007). acute gouty arthritis. Arthritis Care Res. 64, 167. Stamp, L., Day, R. & Yun, J. Allopurinol
127. Taylor, W. J. et al. Diagnostic arthrocentesis for 1447–1461 (2012). hypersensitivity: investigating the cause and
suspicion of gout is safe and well tolerated. 150. Rubin, B. R. et al. Efficacy and safety profile of minimizing the risk. Nat. Rev. Rheumatol. 12,
J. Rheumatol. 43, 150–153 (2016). treatment with etoricoxib 120 mg once daily 235–242 (2016).
128. Ankli, B. et al. Calcium pyrophosphate deposition compared with indomethacin 50 mg three times daily 168. Hung, S. I. et al. HLA-​B*5801 allele as a genetic
disease: a frequent finding in patients with long-​ in acute gout: a randomized controlled trial. Arthritis marker for severe cutaneous adverse reactions caused
standing erosive gout. Scand. J. Rheumatol. 47, Rheum. 50, 598–606 (2004). by allopurinol. Proc. Natl Acad. Sci. USA 102,
127–130 (2018). 151. Janssens, H. J., Janssen, M., van de Lisdonk, E. H., 4134–4139 (2005).
129. Filippucci, E. et al. Ultrasound imaging for the van Riel, P. L. & van Weel, C. Use of oral prednisolone This was the original description of HLA-​B*5801
rheumatologist. XLVII. Ultrasound of the shoulder or naproxen for the treatment of gout arthritis: as a genetic marker for allopurinol-​induced SCAR.
in patients with gout and calcium pyrophosphate a double-​blind, randomised equivalence trial. Lancet 169. Ko, T. M. et al. Use of HLA-​B*58:01 genotyping to
deposition disease. Clin. Exp. Rheumatol. 31, 371, 1854–1860 (2008). prevent allopurinol induced severe cutaneous adverse
659–664 (2013). 152. Terkeltaub, R. et al. High versus low dosing of oral reactions in Taiwan: national prospective cohort study.
130. Reuss-​Borst, M. A., Pape, C. A. & Tausche, A. K. colchicine for early acute gout flare. Arthritis Rheum. BMJ 351, h4848 (2015).
Hidden gout- Ultrasound findings in patients with 62, 1060–1068 (2010). 170. Chung, W.-H. et al. Insights into the poor prognosis of
musculo-​skeletal problems and hyperuricemia. 153. Schlesinger, N. et al. Canakinumab relieves symptoms allopurinol-​induced severe cutaneous adverse
Springerplus 3, 592 (2014). of acute flares and improves health-​related quality of reactions: the impact of renal insufficiency, high
131. Choi, H. K. et al. Dual energy computed tomography life in patients with difficult to treat gouty arthritis by plasma levels of oxypurinol and granulysin.
in tophaceous gout. Ann. Rheum. Dis. 68, suppressing inflammation: results of a randomized, Ann. Rheum. Dis. 74, 2157–2164 (2015).
1609–1612 (2009). dose-​ranging study. Arthritis Res. Ther. 13, R53 (2011). 171. Shibagaki, Y., Ohno, I., Hosoya, T. & Kimura, K. Safety,
132. Wang, Y., Deng, X., Xu, Y., Ji, L. & Zhang, Z. Detection 154. Dalbeth, N. et al. Effects of allopurinol dose escalation efficacy and renal effect of febuxostat in patients with
of uric acid crystal deposition by ultrasonography and on bone erosion and urate volume in gout: a dual moderate-​to-severe kidney dysfunction. Hypertens.
dual-​energy computed tomography: a cross-​sectional energy CT imaging study of a randomized controlled Res. 37, 919–925 (2014).
study in patients with clinically diagnosed gout. trial. Arthritis Rheumatol. https://doi.org/10.1002/ 172. Saag, K. et al. Impact of febuxostat on renal function
Medicine 97, e12834 (2018). art.40929 (2019). in gout subjects with moderate-​to-severe renal
133. Ogdie, A. et al. Performance of ultrasound in the 155. Doherty, M. et al. Efficacy and cost-​effectiveness of impairment. Arthritis Rheum. 68, 2035–2043
diagnosis of gout in a multicenter study: comparison nurse-​led care involving education and engagement of (2016).
with monosodium urate monohydrate crystal analysis patients and a treat-​to-target urate-​lowering strategy 173. White, W. et al. Cardiovascular safety of febuxostat or
as the gold standard. Arthritis Rheumatol. 69, versus usual care for gout: a randomised controlled allopurinol in patients with gout. N. Engl. J. Med.
429–438 (2017). trial. Lancet 392, 1403–1412 (2018). 378, 1200–1210 (2018).
134. Logan, J. A., Morrison, E. & McGill, P. E. Serum uric This large, randomized controlled trial The published results of the large, randomized
acid in acute gout. Ann. Rheum. Dis. 56, 696–697 demonstrated major improvements in clinical controlled CARES trial show increased
(1997). outcomes in patients receiving nurse-​led gout care cardiovascular and all-​cause mortality from
135. Kang, E. H., Lee, E. Y., Lee, Y. J., Song, Y. W. & Lee, E. B. according to rheumatology guidelines for gout febuxostat treatment compared with allopurinol
Clinical features and risk factors of postsurgical gout. management compared with usual care by a treatment in individuals with gout and established
Ann. Rheum. Dis. 67, 1271–1275 (2008). general practitioner. cardiovascular disease.
136. Kannangara, D. R. et al. Fractional clearance of urate: 156. Hui, M. et al. The British Society for Rheumatology 174. Choi, H., Neogi, T., Stamp, L., Dalbeth, N. &
validation of measurement in spot-​urine samples in guideline for the management of gout. Rheumatology Terkeltaub, R. Implications of the cardiovascular safety
healthy subjects and gouty patients. Arthritis Res. 56, 1056–1059 (2017). of febuxostat and allopurinol in patients with gout and
Ther. 14, R189 (2012). 157. Qaseem, A., Harris, R. P. & Forciea, M. A. cardiovascular morbidities (CARES) trial and
137. Perez-​Ruiz, F., Hernandez-​Baldizon, S., Management of acute and recurrent gout: a clinical associated FDA public safety alert. Arthritis
Herrero-​Beites, A. M. & Gonzalez-​Gay, M. A. Risk practice guideline from the American College of Rheumatol. 70, 1702–1709 (2018).
factors associated with renal lithiasis during uricosuric Physicians. Ann. Intern. Med. 166, 58–68 (2017). 175. Pui, K., Gow, P. & Dalbeth, N. Efficacy and tolerability
treatment of hyperuricemia in patients with gout. 158. Taylor, T. H., Mecchella, J. N., Larson, R. J., Kerin, K. D. of probenecid as urate-​lowering therapy in gout;
Arthritis Care Res. 62, 1299–1305 (2010). & Mackenzie, T. A. Initiation of allopurinol at first clinical experience in high-​prevalence population.
138. Stamp, L. & Dalbeth, N. Screening for hyperuricaemia medical contact for acute attacks of gout: a J. Rheumatol. 40, 872–876 (2013).
and gout: a perspective and research agenda. randomized clinical trial. Am. J. Med. 125, 176. Bardin, T. et al. Lesinurad in combination with
Nat. Rev. Rheumatol. 10, 752–756 (2014). 1126–1134 (2012). e1127. allopurinol: a randomised, double-​blind, placebo-​
139. Khanna, D. et al. 2012 American College of 159. Hill, E. M., Sky, K., Sit, M., Collamer, A. & Higgs, J. controlled study in patients with gout with inadequate
Rheumatology guidelines for management of gout. Does starting allopurinol prolong acute treated gout? response to standard of care (the multinational
Part 1: systematic nonpharmacologic and A randomized clinical trial. J. Clin. Rheumatol. 21, CLEAR 2 study). Ann. Rheum. Dis. 76, 811–820
pharmacologic therapeutic approaches to 120–125 (2015). (2016).

16 | Article citation ID: (2019) 5:69 www.nature.com/nrdp

0123456789();
Primer

177. Saag, K. et al. Lesinurad combined with allopurinol: 193. Khanna, P. P. et al. Health-​related quality of life and management clinic for achieving target serum uric acid
randomized, double-​blind, placebo-​controlled study in treatment satisfaction in patients with gout: results levels: a randomized clinical trial. Perm. J. 20, 18–23
gout subjects with inadequate response to standard of from a cross-​sectional study in a managed care setting. (2016).
care allopurinol (a US-​based study). Arthritis Rheum. Patient Prefer. Adherence 9, 971–981 (2015). 209. Mikuls, T. R. et al. Adherence and outcomes with
69, 203–212 (2017). 194. Shields, G. E. & Beard, S. M. A systematic review urate-​lowering therapy: a site-​randomized trial.
178. Dalbeth, N. et al. Lesinurad, a selective uric acid of the economic and humanistic burden of gout. Am. J. Med. 132, 354–361 (2018).
reabsorption inhibitor, in combination with febuxostat Pharmacoeconomics 33, 1029–1047 (2015). 210. Reginato, A. M., Mount, D. B., Yang, I. & Choi, H. K.
in patients with tophaceous gout. Arthritis Rheum. 69, 195. Wertheimer, A., Morlock, R. & Becker, M. A. A revised The genetics of hyperuricaemia and gout. Nat. Rev.
1903–1913 (2017). estimate of the burden of illness of gout. Curr. Ther. Rheumatol. 8, 610–621 (2012).
179. Sundy, J. et al. Efficacy and tolerability of pegloticase Res. Clin. Exp. 75, 1–4 (2013).
for the treatment of chronic gout in patients refractory 196. Flores, N. M., Nuevo, J., Klein, A. B., Baumgartner, S. Acknowledgements
to conventional treatment: two randomized controlled & Morlock, R. The economic burden of uncontrolled The authors are grateful to A. Nakayama for editing assis-
trials. JAMA 306, 711–720 (2011). gout: how controlling gout reduces cost. J. Med. Econ. tance. H.K.C. is supported by the US National Institutes of
180. Becker, M. et al. Long-​term safety of pegloticase in 22, 1–6 (2018). Health (AR060772).
chronic gout refractory to conventional treatment. 197. Strand, V., Khanna, D., Singh, J. A., Forsythe, A. &
Ann. Rheum. Dis. 72, 1469–1474 (2013). Edwards, N. L. Improved health-​related quality of life Author contributions
181. Lipsky, P. et al. Pegloticase immunogenicity: and physical function in patients with refractory Introduction (N.D.); Epidemiology (H.K.C.); Mechanisms/
the relationship between efficacy and antibody chronic gout following treatment with pegloticase: pathophysiology (H.M. and L.A.B.J.); Diagnosis, screening
development in patients treated for refractory chronic evidence from phase III randomized controlled trials. and prevention (F.P.-R. and N.D.); Management (L.K.S.);
gout. Arthritis Res. Ther. 16, R60 (2014). J. Rheumatol. 39, 1450–1457 (2012). Quality of life (P.P.K.); Outlook (N.D.); overview of the Primer
182. Becker, M. et al. The urate-​lowering efficacy and safety 198. Dalbeth, N. et al. Effects of skim milk powder enriched (N.D.).
of febuxostat in the treatment of the hyperuricaemia of with glycomacropeptide and G600 milk fat extract on
gout: the CONFIRMS trial. Arthritis Res. Ther. 12, R63 frequency of gout flares: a proof-​of-concept randomised Competing interests
(2010). controlled trial. Ann. Rheum. Dis. 71, 929–934 (2012). N.D. has received speaking fees from Pfizer, Horizon, Janssen
183. Li, X. et al. Serum uric acid levels and multiple health 199. Stamp, L. et al. Clinically insignificant effect of and AbbVie; consulting fees from Horizon, Hengrui and
outcomes: umbrella review of evidence from supplemental vitamin C on serum urate in patients Kowa; and research funding from Amgen and AstraZeneca;
observational studies, randomised controlled trials, with gout; a pilot randomised controlled trial. Arthritis and is currently principal investigator on a clinical trial of
and Mendelian randomisation studies. BMJ 357, Rheum. 65, 1636–1642 (2013). intensive urate-​lowering therapy (funded by the Health
j2376 (2017). 200. Rai, S. K. et al. Key barriers to gout care: a systematic Research Council of New Zealand). Within the past 5 years,
184. Terkeltaub, R., Furst, D., DiGiacinto, J., Kook, K. review and thematic synthesis of qualitative studies. N.D. has been principal investigator on a clinical trial of
& Davis, M. Novel evidence-​based colchicine dose-​ Rheumatology 57, 1282–1292 (2018). febuxostat in early gout, and received consulting or speaking
reduction algorithm to predict and prevent colchicine 201. Robinson, P. C., Taylor, W. J. & Dalbeth, N. An fees from Takeda, Menarini and Teijin. H.K.C. received
toxicity in the presence of cytochrome P450 3A4/ observational study of gout prevalence and quality research support from AstraZeneca and consulting fees from
P-​glycoprotein inhibitors. Arthritis Rheum. 63, of care in a national Australian general practice Takeda, Selecta, GSK and Horizon. L.A.B.J. has received a
2226–2237 (2011). population. J. Rheumatol. 42, 1702–1707 (2015). speaking fee from Novartis, research funding from Ardea
185. Jeyaruban, A., Soden, M. & Larkins, S. General 202. FitzGerald, J. D. et al. Development of the American Biosciences/AstraZeneca, and is Scientific Advisory Board
practitioners’ perspectives on the management of College of Rheumatology electronic clinical quality member of Olatec Therapeutics LLC. P.P.K. has received
gout: a qualitative study. Postgrad. Med. J. 92, measures for gout. Arthritis Care Res. 70, 659–671 research grants from Ironwood, Sobi and Horizon. F.P.-R. has
603–607 (2016). (2018). been an advisor for Amgen, Horizon, Grünenthal, Menarini,
186. Nielsen, S. et al. Weight loss for overweight and obese 203. Roddy, E., Zhang, W. & Doherty, M. Concordance of Sanofi and Syneos-​Health; a speaker for Amgen, Astellas,
individuals with gout: a systematic review of longitudinal the management of chronic gout in a UK primary-​care Grünenthal, Lilly, Logarithm, Menarini and the Spanish
studies. Ann. Rheum. Dis. 76, 1870–1882 (2017). population with the EULAR gout recommendations. Foundation for Rheumatology; received investigation grants
187. Holland, R. & McGill, N. Comprehensive dietary Ann. Rheum. Dis. 66, 1311–1315 (2007). from Cruces Rheumatology Association and the Spanish
education in treated gout patients does not further 204. Scheepers, L. et al. Medication adherence among Foundation for Rheumatology; and received congress funding
improve serum urate. Int. Med. J. 45, 189–189 (2015). patients with gout: a systematic review and meta-​ from Lilly, Novartis and the European League Against
188. Becker, M. A. et al. Quality of life and disability in analysis. Semin. Arthritis Rheum. 47, 689–702 Rheumatism (EULAR). L.K.S. has received speaking fees from
patients with treatment-​failure gout. J. Rheumatol. (2018). Amgen and is currently principal investigator on a clinical trial
36, 1041–1048 (2009). 205. Kuo, C. F., Grainge, M. J., Mallen, C., Zhang, W. of colchicine prophylaxis (funded by the Health Research
189. Kleinman, N. L. et al. The impact of gout on work & Doherty, M. Eligibility for and prescription of Council of New Zealand). L.K.S. is a member of the Medicines
absence and productivity. Value Health 10, 231–237 urate-​lowering treatment in patients with incident Adverse Reaction Committee of New Zealand and the New
(2007). gout in England. JAMA 312, 2684–2686 (2014). Zealand Pharmaceutical Management Agency (PHARMAC)
190. Khanna, P. P. et al. Tophi and frequent gout flares 206. Duyck, S. D., Petrie, K. J. & Dalbeth, N. “You don’t Rheumatology Subcommittee. H.M. declares no competing
are associated with impairments to quality of life, have to be a drinker to get gout, but it helps”: interests.
productivity, and increased healthcare resource use: a content analysis of the depiction of gout in popular
results from a cross-​sectional survey. Health Qual. newspapers. Arthritis Care Res. 68, 1721–1725 Peer review information
Life Outcomes 10, 117 (2012). (2016). Nature Reviews Disease Primers thanks J.G. Puig,
191. Chandratre, P. et al. Health-​related quality of life in 207. Bulbin, D. et al. Improved gout outcomes in primary M. Pillinger, P. Richette, M. Andrés and R. Burgos-Vargas
gout in primary care: baseline findings from a cohort care using a novel disease management program: for their contribution to the peer review of this work.
study. Semin. Arthritis Rheum. 48, 61–69 (2018). a pilot study. Arthritis Care Res. 70, 1679–1685
192. Chandratre, P. et al. Health-​related quality of life in (2018). Publisher’s note
gout: a systematic review. Rheumatology 52, 208. Goldfien, R., Pressman, A., Jacobson, A., Ng, M. Springer Nature remains neutral with regard to jurisdictional
2031–2040 (2013). & Avins, A. A pharmacist-​staffed, virtual gout claims in published maps and institutional affiliations.


NATURE REvIEwS | DISeASe PrIMerS | Article citation ID: (2019) 5:69 17

0123456789();

You might also like