Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

PHYSIOLOGY IN MEDICINE: A SERIES OF ARTICLES LINKING MEDICINE WITH SCIENCE

Physiology in Medicine: Dennis A. Ausiello, MD, Editor; Dale J. Benos, PhD, Deputy Editor; Francois Abboud, MD,
Associate Editor; William J. Koopman, MD, Associate Editor
Review
Annals of Internal Medicine: Paul Epstein, MD, Series Editor

Pathogenesis of Gout
Hyon K. Choi, MD, DrPH; David B. Mount, MD; and Anthony M. Reginato, MD, PhD

Clinical Principles Pathophysiologic Principles

The overall disease burden of gout is substantial and may be A direct causal relationship exists between serum urate levels
increasing. and the risk for gout.

As more scientific data on the modifiable risk factors and Lifestyle factors, including adiposity and dietary habits,
comorbidities of gout become available, integration of appear to contribute to serum uric acid levels and the risk
these data into gout care strategies may become essential. for gout.

Hyperuricemia and gout are associated with the insulin Urate is extensively reabsorbed from the glomerular
resistance syndrome and related comorbid conditions. ultrafiltrate in the proximal tubule via the brush-border
urate–anion exchanger URAT1.
Lifestyle modifications that are recommended for gout
generally align with those for major chronic disorders (such Sodium-dependent reabsorption of anions increases their
as the insulin resistance syndrome, hypertension, and concentration in proximal tubule cells, resulting in increased
cardiovascular disorders); thus, these measures may be urate exchange via URAT1, increased urate reabsorption by
doubly beneficial for many patients with gout and the kidney, and hyperuricemia.
particularly for individuals with these comorbid conditions.
Genetic variation in renal urate transporters or upstream
Effective management of risk factors for gout and careful regulatory factors may explain the hereditary susceptibility
selection of certain therapies for comorbid conditions (such to conditions associated with high urate levels and a
as hypertension or the insulin resistance syndrome) may patient’s particular response to medications; these
also aid gout care. transporters may also serve as targets for future drug
development.
The urate–anion exchanger URAT1 (urate transporter-1) is a
specific target of action for both antiuricosuric and Urate crystals are able to directly initiate, to amplify, and to
uricosuric agents. sustain an intense inflammatory attack because of their
ability to stimulate the synthesis and release of humoral
The long-term health effect of hyperuricemia (beyond the and cellular inflammatory mediators.
increased risk for gout) needs to be clarified, including any
potential consequences associated with the chronic Cytokines, chemokines, proteases, and oxidants involved in
hyperuricemia that anti-inflammatory treatment does not acute urate crystal–induced inflammation also contribute to
correct. the chronic inflammation that leads to chronic gouty
synovitis, cartilage loss, and bone erosion.

G out is a type of inflammatory arthritis that is triggered by


the crystallization of uric acid within the joints and is
often associated with hyperuricemia (Figure 1). Acute gout is
The overall disease burden of gout remains substantial
and may be increasing. The prevalence of self-reported,
physician-diagnosed gout in the Third National Health
typically intermittent, constituting one of the most painful
conditions experienced by humans. Chronic tophaceous gout
usually develops after years of acute intermittent gout, al- See also:
though tophi occasionally can be part of the initial presenta- Print
tion. In addition to the morbidity that is attributable to gout Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 500
itself, the disease is associated with such conditions as the
insulin resistance syndrome, hypertension, nephropathy, and Web-Only
disorders associated with increased cell turnover (1, 2). Conversion of figures and table into slides

Ann Intern Med. 2005;143:499-516.


For author affiliations, see end of text.
For definition of terms used, see Glossary.

© 2005 American College of Physicians 499


Review Pathogenesis of Gout

and Nutrition Examination Survey was found to be greater Glossary


than 2% in men older than 30 years of age and in women
Adenosine: A condensation product of adenine and D-ribose; a nucleoside
older than 50 years of age (3). The prevalence increased found among the hydrolysis products of all nucleic acids and of the
with increasing age and reached 9% in men and 6% in various adenine nucleotides.
women older than 80 years of age (4). Furthermore, the Adenosine triphosphate: A phosphorylated nucleoside C10H16N5O13P3 of
adenine that supplies energy for many biochemical cellular processes by
incidence of primary gout (that is, patients without di- undergoing enzymatic hydrolysis (especially to adenosine diphosphate).
uretic exposure) doubled over the past 20 years, according Anion exchanger: A transport protein that mediates movement of an anion
across the plasma membrane by exchanging it with another anion on the
to the Rochester Epidemiology Project (4). Dietary and opposite side of the membrane. Urate–anion exchange plays a key role in
lifestyle trends and the increasing prevalence of obesity and the transport of urate across cell membranes.
the metabolic syndrome may explain the increasing inci- Antiuricosuric agent: A chemical or drug that results in reduced renal
excretion of urate and hyperuricemia; pyrazinamide, the classic
dence of gout. antiuricosuric drug, exerts its effect by promoting proximal tubular
Researchers have recently made great advances in de- reabsorption of urate.
Apolipoprotein: The protein component of any lipoprotein complexes that is
fining the pathogenesis of gout, including elucidating its a normal constituent of plasma chylomicrons, high-density lipoproteins,
risk factors and tracing the molecular mechanisms of renal low-density lipoproteins, and very low-density lipoproteins in humans.
urate transport and crystal-induced inflammation. This ar- Apoptosis: Disintegration of cells into membrane-bound particles that are
then phagocytosed by other cells.
ticle reviews key aspects of the pathogenesis of gout with a Brush-border membrane vesicles (BBMV): Purified from superficial renal
focus on the recent advances. cortex, BBMV are predominantly derived from the renal proximal tubule;
urate transporter-1 was initially defined as an anion exchanger activity
present in renal BBMV preparations.
ABSENCE OF URICASE IN HUMANS Calcium-binding cytoplasmic proteins S100A8, S100A9: Chemotactic factor
that stimulates neutrophil adhesion and migration by activating the
Humans are the only mammals in whom gout is ␤2-integrin CD11b/CD18.
known to develop spontaneously, probably because hyper- Chemokines: A class of polypeptide cytokines, usually 8–10 kDa, that are
chemokinetic and chemotactic, stimulating leukocyte movement and
uricemia only commonly develops in humans (5). In most attraction.
fish, amphibians, and nonprimate mammals, uric acid that Cis-inhibition: Competitive inhibition of urate exchange by a urate
transporter-1 substrate present at the same side of the plasma membrane.
has been generated from purine (see Glossary) metabolism
Chondroitin: A mucopolysaccharide occurring in sulfated form; present
undergoes oxidative degradation through the uricase en- among the ground substance materials in the extracellular matrix of
zyme, producing the more soluble compound allantoin. In connective tissue (for example, cartilage).
c-Jun N-terminal kinase: Downstream kinase activated by ERK-1/ERRK-2
humans, the uricase gene is crippled by 2 mutations that and p38 cascades, leading to autophosphorylation and regulation of
introduce premature stop codons (see Glossary) (6). The complex biological responses.
absence of uricase, combined with extensive reabsorption Cyclooxygenase-2 (COX-2): An enzyme that makes the prostaglandins that
cause inflammation, pain, and fever; nonsteroidal anti-inflammatory drugs
of filtered urate, results in urate levels in human plasma relieve symptoms as result of their ability to block COX-2 enzymes.
that are approximately 10 times those of most other mam- Cytokines: Intercellular messenger proteins; hormone-like products of many
mals (30 to 59 ␮mol/L) (7). The evolutionary advantage of different cell types that are usually active within a small radius of the cells
producing them.
these findings is unclear, but urate may serve as a primary Docosahexaenoic acid (DHA): All-cis-4,7,10,13,16,19-docosahexaenoic acid,
antioxidant in human blood because it can remove singlet an ␻-3, polyunsaturated, 22-carbon fatty acid found almost exclusively in
fish and marine animal oils; a substrate for cyclooxygenase.
oxygen and radicals as effectively as vitamin C (8). Of note, Eicosapentaenoic acid (EPA): All-cis-5,8,11,14,17-eicosapentaenoic acid, an
levels of plasma uric acid (about 300 ␮M) are approxi- ␻-3, polyunsaturated, 20-carbon fatty acid found almost exclusively in
fish and marine animal oils; a substrate for cyclooxygenase.
mately 6 times those of vitamin C in humans (8, 9). Other
E-selectin: Endothelial cell adhesion molecules consisting of a lectin-like
potential advantages of the relative hyperuricemia in pri- domain, an epidermal growth factor–like domain, and a variable number
mate species have been speculated (8, 10, 11). However, of domains that encode proteins homologous to complement-binding
proteins; their function is to mediate the binding of leukocytes to the
hyperuricemia can be detrimental in humans, as demon- vascular endothelium.
strated by its proven pathogenetic roles in gout and neph- Familial renal hypouricemia: A recessive genetic disorder caused by
rolithiasis and by its putative roles in hypertension and homozygous loss-of-function mutations in the SLC22A12 gene encoding
urate transporter-1. Patients with this disorder have hypouricemia that
other cardiovascular disorders (12). does not respond to uricosuric or antiuricosuric agents.
G proteins: A family of similar heterotrimeric proteins found in the
intracellular portion of the plasma membrane; bind activated receptor
THE ROLE OF URATE LEVELS complexes and, through conformational changes and cyclic binding and
hydrolysis of guanosine triphosphate, directly or indirectly effect
Uric acid is a weak acid (pKa, 5.8) that exists largely as alterations in channel gating and couple cell surface receptors to
urate, the ionized form, at physiologic pH. As urate con- intracellular responses.
Interleukins: A large family of hormone-like messenger proteins produced
centration increases in physiologic fluids, the risk for su- by immune cells that act on leukocytes and other cells.
persaturation and crystal formation generally increases. Kinase: An enzyme catalyzing the conversion of a proenzyme to an active
Population studies indicate a direct positive association be- enzyme (for example, enteropeptidase [enterokinase]) or catalyzing the
transfer of phosphate groups.
tween serum urate levels and a future risk for gout (13, 14), Kinin: One of a number of widely differing substances having pronounced
as shown in Figure 2. Conversely, the use of antihyperuri- and dramatic physiologic effects; kallidin and bradykinin are polypeptides,
cemic medication is associated with an 80% reduced risk formed in blood by proteolysis secondary to some pathologic process
producing vasodilation.
for recurrent gout, confirming the direct causal relation-
ship between serum uric acid levels and risk for gouty ar- Continued

500 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 www.annals.org
Pathogenesis of Gout Review
Glossary—Continued Glossary—Continued

Leptin: A helical protein secreted by adipose tissue; acts on a receptor site in Transforming growth factor-␤ (TGF-␤): A regulatory cytokine that has
the ventromedial nucleus of the hypothalamus to curb appetite and multifunctional properties and can enhance or inhibit many cellular
increase energy expenditure as body fat stores increase. functions, including interfering with the production of other cytokines and
Leukotriene: Substance produced from arachidonic acid by the lipoxygenase enhancing collagen deposition.
pathway; functions as a regulator of allergic and inflammatory reactions; Trans-stimulation: Stimulation of urate exchange by a urate transporter-1
stimulates the movement of leukocytes; identified by the letters A, B, C, substrate when present at the opposite side of the plasma membrane;
D, and E, with subscripts indicating the number of double bonds in the antiuricosuria apparently results from trans-stimulation of urate
molecule (for example, LTB4). reabsorption by anions within the cytoplasm of proximal tubular epithelial
Lipoxins: Any of several conjugated tetraene derivatives of arachidonic acid cells.
that oppose the actions of leukotrienes, have potent vasodilating effects, Tumor necrosis factor (TNF): A polypeptide cytokine, produced by
and appear to be toxic to natural killer cells. endotoxin-activated macrophages, that has the ability to modulate
Matrix metalloproteinases: A family of protein-hydrolyzing endopeptidases adipocyte metabolism, lyse tumor cells in vitro, and induce hemorrhagic
that hydrolyze extracellular proteins, especially collagens and elastin. necrosis of certain transplantable tumors in vivo.
Mitogen-activated protein kinases ERK1/ERK: One of the mitogen-activated Urate transporter-1 (URAT1): The urate–anion exchanger expressed at the
protein kinases that signals transduction pathways in eukaryotic cells and apical brush-border membrane of proximal tubular epithelial cells; URAT1
integrates diverse extracellular signals; regulates complex biological is encoded by the SLC22A12 gene.
responses, such as growth, differentiation, and death. Urate transporter/channel (UAT): Also known as galectin-9; may also be
Multidrug resistance protein-4 (MRP4): An anion transporter capable of involved in proximal tubular urate secretion.
adenosine triphosphate–driven urate efflux, expressed at the apical Uricosuric agent: A chemical or drug that results in increased renal excretion
membrane of the proximal tubule. of urate; urate transporter-1 appears to be the major target for uricosuric
Nucleotide: A combination of a nucleic acid (purine or pyrimidine), 1 sugar drugs.
(ribose or deoxyribose), and a phosphoric group. Voltage-driven organic anion transporter-1 (OATV1): A voltage-sensitive
Organic anion transporter-1 (OAT1): A basolateral anion exchanger organic anion transporter capable of transporting urate and expressed at
involved in proximal tubular transport of multiple organic anions, the apical membrane of the proximal tubule.
including urate; OAT1 is encoded by the SLC22A6 gene.
Organic anion transporter-3 (OAT3): A basolateral anion exchanger
involved in proximal tubular transport of multiple organic anions,
including urate; OAT3 is encoded by the SLC22A8 gene.
␻-3 fatty acids: Polyunsaturated fatty acids that have the final double bond
thritis (15). The solubility of urate in joint fluids, however,
in the hydrocarbon chain between the third and fourth carbon atoms is influenced by other factors in the joint, as shown in
from 1 end of the molecule; found especially in fish, fish oils, vegetable Figure 3. Such factors include temperature, pH, concen-
oils, and green leafy vegetables.
Osteocalcin: A vitamin K–dependent, calcium-binding bone protein, the tration of cations, level of articular dehydration, and the
most abundant noncollagen protein in bone; increased serum presence of such nucleating agents as nonaggregated pro-
concentrations are a marker of increased bone turnover in disease states.
teoglycans, insoluble collagens, and chondroitin sulfate (see
p38 mitogen–activated protein kinase: One of the mitogen-activated
protein kinases that signals transduction pathways in eukaryotic cells and Glossary) (16 –18). Variation in these factors may account
integrates diverse extracellular signals; regulates complex biological for some of the difference in the risk for gout associated
responses such as growth, differentiation, and death.
Peroxisome proliferator-activated receptor-␥ receptor (PPAR-␥): A nuclear
with a given elevation in serum urate level (13, 14). Fur-
receptor regulating an array of diverse functions in a variety of cell types, thermore, these factors may explain the predilection of
including regulation of genes associated with growth and differentiation. gout in the first metatarsal phalangeal joint (a peripheral
Phospholipase: An enzyme that catalyzes the hydrolysis of a phospholipid.
Prostaglandin: Any of a class of physiologically active substances present in joint with a lower temperature) and osteoarthritic joints
many tissues; causes vasodilation, vasoconstriction, and antagonism to (18) (degenerative joints with nucleating debris) and the
hormones that influence lipid metabolism.
Proteoglycan: Any of a class of glycoproteins of high molecular weight that
nocturnal onset of pain (because of intra-articular dehydra-
are found especially in the extracellular matrix of connective tissue. tion) (19).
Proximal tubule: The earliest segment of the renal tubule, responsible for
the reabsorption of urate and other solutes from the glomerular
Urate Balance
ultrafiltrate. The amount of urate in the body depends on the bal-
Purine: A double-ringed, crystalline organic base, C5H4N4, from which the
nitrogen bases adenine and guanine are derived; uric acid is a metabolic
ance between dietary intake, synthesis, and the rate of ex-
end product. cretion (20), as shown in Figure 1. Hyperuricemia results
SLC22 gene family: The “Solute Carrier-22” gene family encompasses more from urate overproduction (10%), underexcretion (90%),
than 20 different genes encoding organic anion and cation transporters,
including the urate transporter-1 (URAT1, SLC22A12), organic anion or often a combination of the two. The purine precursors
transporter-1 (OAT1, SLC22A6), and organic anion transporter-3 (OAT3, come from exogenous (dietary) sources or endogenous me-
SLC22A8).
tabolism (synthesis and cell turnover).
SLC5A8: A member of the SLC5 gene family of sodium-coupled
transporters; a leading candidate for the sodium-dependent The Relationship between Purine Intake and Urate
lactate/butyrate/pyrazinoate/nicotinate transporter that collaborates with
urate transporter-1 in proximal tubular reabsorption of urate.
Levels
Src tyrosine kinase: One of a group of enzymes of the transferase class that The dietary intake of purines contributes substantially
catalyze the phosphorylation of tyrosine residues in specific membrane to the blood uric acid. For example, the institution of an
vesicle–associated proteins.
Stop codon: Trinucleotide sequence (UAA, UGA, or UAG) that specifies the entirely purine-free diet over a period of days can reduce
end of translation or transcription. blood uric acid levels of healthy men from an average of
Synovitis: Inflammation of a synovial membrane, especially that of a joint;
in general, when unqualified, the same as arthritis.
297 ␮mol/L to 178 ␮mol/L (21, 22). The bioavailable
Transcription: Transfer of genetic code information from one kind of nucleic purine content of particular foods would depend on their
acid to another; commonly used to refer to transfer of genetic relative cellularity and the transcriptional (see Glossary)
information from DNA to RNA.
and metabolic activity of the cellular content (20). Little
Continued is known, however, about the precise identity and quantity
www.annals.org 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 501
Review Pathogenesis of Gout

Figure 1. Overview of the pathogenesis of gout. increased risk for gout (27). The variation in the risk for
gout associated with different purine-rich foods may be
explained by varying amounts and type of purine content
and their bioavailability for metabolizing purine to uric
acid (28). At the practical level, these data suggest that
dietary purine restriction in patients with gout or hyper-
uricemia (29, 30) may be applicable to purines of animal
origin but not to purine-rich vegetables, which are excel-
lent sources of protein, fiber, vitamins, and minerals. Sim-
ilarly, implications of the recent findings (27, 28, 31) in
the management of hyperuricemia or gout were consistent
with the new dietary recommendations for the general
public (32), with the exception of the guidelines for fish
intake (Figure 4). Thus, among patients with gout or hy-
peruricemia, the use of plant-derived ␻-3 fatty acids or
supplements of eicosapentaenoic acid and docosahexaenoic
acid (see Glossary) instead of fish consumption could be
considered to provide the benefit of these fatty acids with-
out increasing the risk for gout.

PURINE METABOLISM AND GOUT


The steps in the urate production pathways implicated
in the pathogenesis of gout are displayed in Figure 5. The
vast majority of patients with endogenous overproduction
Gout is mediated by the supersaturation and crystallization of uric acid
within the joints. The amount of urate in the body depends on the of urate have the condition as a result of salvaged purines
balance between dietary intake, synthesis, and excretion. Hyperuricemia arising from increased cell turnover in proliferative and
results from the overproduction of urate (10%), from underexcretion of inflammatory disorders (for example, hematologic cancer
urate (90%), or often a combination of the two. Approximately one third
of urate elimination in humans occurs in the gastrointestinal tract, with
the remainder excreted in the urine.
Figure 2. The relationship between serum uric acid levels and
the incidence of gout.

of individual purines in most foods, especially when


cooked or processed (23). When a purine precursor is in-
gested, pancreatic nucleases break its nucleic acids into nu-
cleotides (see Glossary), phosphodiesterases break oligonu-
cleotides into simple nucleotides, and pancreatic and
mucosal enzymes remove phosphates and sugars from nu-
cleotides (20). The addition of dietary purines to purine-
free dietary protocols has revealed a variable increase in
blood uric acid levels, depending on the formulation and
dose of purines administered (21). For example, RNA has
a greater effect than an equivalent amount of DNA (24),
ribomononucleotides have a greater effect than nucleic acid
(21), and adenine has a greater effect than guanine (25, 26).
A recent large prospective study showed that men in
the highest quintile of meat intake had a 41% higher risk
for gout compared with the lowest quintile, and men in the
highest quintile of seafood intake had a 51% higher risk
compared with the lowest quintile (27). Correspondingly,
in a nationally representative sample of men and women in
Annual incidence of gout was less than 0.1% for men with serum uric
the United States, higher levels of meat and seafood con- acid levels less than 416 ␮mol/L, 0.4% for men with levels of 416 to 475
sumption were associated with higher serum uric acid lev- ␮mol/L, 0.8% for men with levels of 476 to 534 ␮mol/L, 4.3% for men
els (28). However, consumption of oatmeal and purine- with levels of 535 to 594 ␮mol/L, and 7.0% for men with levels greater
than 595 ␮mol/L, according to the Normative Aging Study (13). The
rich vegetables (for example, peas, beans, lentils, spinach, solid line denotes these data points; the dotted line shows an exponential
mushrooms, and cauliflower) was not associated with an projection of the data points.
502 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 www.annals.org
Pathogenesis of Gout Review
Figure 3. Mechanisms of monosodium urate crystal formation and induction of crystal-induced inflammation.

Urate crystallizes as a monosodium salt in oversaturated tissue fluids. Its crystallization depends on the concentrations of both urate and cation levels.
Several other factors contribute to the decreased solubility of sodium urate and crystallization. Alteration in the extracellular matrix leading to an increase
in nonaggregated proteoglycans, chondroitin sulfate, insoluble collagen fibrils, and other molecules in the affected joint may serve as nucleating agents.
Furthermore, monosodium urate (MSU) crystals can undergo spontaneous dissolution depending on their physiochemical environments. Chronic
cumulative urate crystal formation in tissue fluids leads to MSU crystal deposition (tophus) in the synovium and cell surface layer of cartilage. Synovial
tophi are usually walled off, but changes in the size and packing of the crystal from microtrauma or from changes in uric acid levels may loosen them
from the organic matrix. This activity leads to “crystal shedding” and facilitates crystal interaction with synovial cell lining and residential inflammatory
cells, leading to an acute gouty flare.

and psoriasis), from pharmacologic intervention resulting tion by net ATP degradation to AMP (41, 44). In addi-
in increased urate production (such as chemotherapy), or tion, decreased urinary excretion as a result of dehydration
from tissue hypoxia. Only a small proportion of those with and metabolic acidosis may contribute to the hyperurice-
urate overproduction (10%) have the well-characterized in- mia that is associated with ethanol ingestion, as discussed
born errors of metabolism (for example, superactivity of later in this review (34, 45).
5’-phosphoribosyl-1-pyrophosphate synthetase and defi- Recently, a large-scale prospective study confirmed
ciency of hypoxanthine– guanine phosphoribosyl trans- that the effect of ethanol on urate levels can be translated
ferase). These genetic disorders have been extensively re- into the risk for gout (31). Compared with abstinence,
viewed in textbooks (20, 33, 34), and the involved daily alcohol consumption of 10 to 14.9 g increased the
pathways are depicted in Figure 5. risk for gout by 32%; daily consumption of 15 to 29.9 g,
Conditions associated with net adenosine triphosphate 30 to 49.9 g, and 50 g or greater increased the risk by 49%,
(ATP) (see Glossary) degradation lead to accumulation of 96%, and 153%, respectively. Furthermore, the study also
adenosine diphosphate (ADP) and adenosine monophos- found that this risk varied according to type of alcoholic
phate (AMP), which can be rapidly degraded to uric acid beverage: Beer conferred a larger risk than liquor, whereas
(35– 44), as shown in Figure 5. For example, ethanol ad- moderate wine drinking did not increase risk (31). Corre-
ministration has been shown to increase uric acid produc- spondingly, a national U.S. survey demonstrated parallel
www.annals.org 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 503
Review Pathogenesis of Gout

Figure 4. Dietary influences on the risk for gout and their implications within the Harvard Healthy Eating Pyramid.

Data on the relationship between diet and the risk for gout are primarily derived from the recent Health Professionals Follow-Up Study (27, 28, 31).
Implications of these findings in the management of hyperuricemia or gout are generally consistent with the new Healthy Eating Pyramid (32), except
for fish intake. The use of plant-derived ␻-3 fatty acids or supplements of eicosapentaenoic acid and docosahexaenoic acid in place of fish consumption
could be considered to provide patients the benefit of these fatty acids without increasing the risk for gout. Use of ␻-3 fatty acids may have
anti-inflammatory effect against gouty flares. Vitamin C intake exerts a uricosuric effect. (Adapted with permission from reference 32: Willett WC,
Stampfer MJ. Rebuilding the food pyramid. Sci Am. 2003;288:64-71.) Red arrows denote an increased risk for gout, solid green arrows denote a
decreased risk, and yellow arrows denote no influence on risk. Broken green arrows denote potential effect but without prospective evidence for the
outcome of gout.

associations between these alcoholic beverages and serum panying phosphate depletion limits regeneration of ATP
urate levels (46). These findings suggest that certain non- from ADP. The subsequent catabolism of AMP serves as a
alcoholic components that vary among these alcoholic bev- substrate for uric acid formation (48). Thus, within min-
erages play an important role in urate metabolism. Ingested utes after fructose infusion, plasma (and later urinary) uric
purines in beer, such as highly absorbable guanosine (23, acid concentrations are increased (42). In conjunction with
47), may produce an effect on blood uric acid levels that is purine nucleotide depletion, rates of purine synthesis de
sufficient to augment the hyperuricemic effect of alcohol novo are accelerated, thus potentiating uric acid produc-
itself, thereby producing a greater risk for gout than liquor tion (43). Oral fructose may also increase blood uric acid
or wine. Whether other nonalcoholic offending factors ex- levels, especially in patients with hyperuricemia (49) or a
ist remains unclear, particularly in regard to beer; instead, history of gout (50). Fructose has also been implicated in
protective factors in wine may be mitigating the alcohol the risk for the insulin resistance syndrome and obesity, which
effect on the risk for gout (28). are closely associated with gout (51, 52). Furthermore, hyper-
Fructose is the only carbohydrate that has been shown uricemia resulting from ATP degradation can occur in acute,
to exert a direct effect on uric acid metabolism (23). Fruc- severe illnesses, such as the adult respiratory distress syndrome,
tose phosphorylation in the liver uses ATP, and the accom- myocardial infarction, or status epilepticus (34 –36).
504 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 www.annals.org
Pathogenesis of Gout Review

ADIPOSITY, INSULIN RESISTANCE, AND GOUT absorption. Finally, in the insulin resistance syndrome, im-
Increased adiposity and the insulin resistance syn- paired oxidative phosphorylation may increase systemic
drome are both associated with hyperuricemia (53–56). adenosine (see Glossary) concentrations by increasing the
Body mass index, waist-to-hip ratio, and weight gain have intracellular levels of coenzyme A esters of long-chain fatty
all been associated with the risk for incident gout in men acids. Increased adenosine, in turn, can result in renal re-
(28, 57). Conversely, small, open-label interventional stud- tention of sodium, urate, and water (66 – 69). Some re-
ies showed that weight reduction was associated with a searchers have speculated that increased extracellular aden-
decline in urate levels and risk for gout (58, 59). osine concentrations over the long term may also
Reduced de novo purine synthesis was observed in pa- contribute to hyperuricemia by increasing urate production
tients after weight loss, resulting in decreased serum urate (66). The growing “epidemic” of obesity (70, 71) and the
levels (60). Exogenous insulin can reduce the renal excre- insulin resistance syndrome (72) present a substantial chal-
tion of urate in both healthy and hypertensive persons (54, lenge in the prevention and management of gout.
61, 62). Insulin may enhance renal urate reabsorption
through stimulation of the urate–anion exchanger urate
transporter-1 (URAT1) (see Glossary) (63) or through the HYPERTENSION, CARDIOVASCULAR DISORDERS, AND
sodium-dependent anion cotransporter in brush-border GOUT
membranes of the renal proximal tubule (discussed later in Associations between hypertension and the incidence
this review). Because serum levels of leptin (see Glossary) of gout have been observed (13, 57), but researchers were
and urate tend to increase together (64, 65), some investi- previously unable to determine whether hypertension was
gators have also suggested that leptin may affect renal re- independently associated or if it only served as a marker for

Figure 5. Urate production pathways implicated in the pathogenesis of gout.

The de novo synthesis starts with 5’-phosphoribosyl 1-pyrophosphate (PRPP), which is produced by addition of a further phosphate group from
adenosine triphosphate (ATP) to the modified sugar ribose-5-phosphate. This step is performed by the family of PRPP synthetase (PRS) enzymes. In
addition, purine bases derived from tissue nucleic acids are reutilized through the salvage pathway. The enzyme hypoxanthine– guanine phosphoribosyl
transferase (HPRT) salvages hypoxanthine to inosine monophosphate (IMP) and guanine to guanosine monophosphate (GMP). Only a small proportion
of patients with urate overproduction have the well-characterized inborn errors of metabolism, such as superactivity of PRS and deficiency of HPRT.
Furthermore, conditions associated with net ATP degradation lead to the accumulation of adenosine diphosphate (ADP) and adenosine monophosphate
(AMP), which can be rapidly degraded to uric acid. These conditions are displayed in left upper corner. Plus sign denotes stimulation, and minus sign
denotes inhibition. APRT ⫽ adenine phosphoribosyl transferase; PNP ⫽ purine nucleotide phosphorylase.
www.annals.org 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 505
Review Pathogenesis of Gout

Table. Substances Affecting Urate Levels and Their Underlying Mechanisms*

Substance Implicated Mechanism


Urate-increasing agents
Pyrazinamide Trans-stimulation of URAT1 (63)
Nicotinate Trans-stimulation of URAT1 (63)
Lactate, ␤-hydroxybutyrate, acetoacetate Trans-stimulation of URAT1 (63)
Salicylate (low dose) Decreased renal urate excretion (78)
Diuretics Increased renal tubular reabsorption associated with volume depletion (79, 80), may
stimulate URAT1 (63)
Cyclosporine Increased renal tubular reabsorption associated with decreased glomerular filtration (81–85),
hypertension (86), interstitial nephropathy
Tacrolimus Similar to cyclosporine (87, 88)
Ethambutol Decreased renal urate excretion
␤-Blockers Unknown (no change in renal urate excretion) (89)

Urate-decreasing agents
Uricosurics
Probenecid Inhibition of URAT1 (63, 90)
Sulfinpyrazone Inhibition of URAT1 (63, 90)
Benzbromarone Inhibition of URAT1 (63, 90)
Losartan Inhibition of URAT1 (63)
Salicylate (high-dose) Inhibition of URAT1 (63)
Fenofibrate May inhibit URAT1
Amlodipine Increased renal urate excretion (86)
Xanthine oxidase inhibitors
Allopurinol Inhibition of xanthine oxidase
Febuxostat Inhibition of xanthine oxidase
Uricase Oxidation of urate to allantoin

* Numbers in parentheses are reference numbers. URAT1 ⫽ urate transporter-1.

associated risk factors, such as dietary factors, obesity, di- by pyrazinoate, the relevant metabolite, has never been
uretic use, and renal failure. A recent prospective study, demonstrated. Indeed, pyrazinamide has no effect in ani-
however, has confirmed that hypertension is associated mal species that eliminate urate through net secretion (92),
with an increased risk for gout independent of these po- and direct effects of the drug on human urate secretion are
tential confounders (28). Renal urate excretion was found largely unsubstantiated (91). Rather, studies utilizing renal
to be inappropriately low relative to glomerular filtration brush-border membrane vesicles (see Glossary) (93, 94)
rates in patients with essential hypertension (73, 74). Re- have shown that pyrazinoate activates the reabsorption of
duced renal blood flow with increased renal and systemic urate through indirect stimulation of apical urate exchange
vascular resistance may also contribute to elevated serum (Figure 5). Similar mechanisms underlie the clinically rel-
uric acid levels (75). Hyperuricemia in patients with essen- evant hyperuricemic effects of lactate (45), ketoacids (95),
tial hypertension may reflect early nephrosclerosis, thus im- and nicotinate (96), as shown in the Table. Recent ad-
plying renal morbidity in these patients. Furthermore, vances in the understanding of the relevant physiology are
studies have suggested that hyperuricemia may be associ- reviewed in the following sections.
ated with incident hypertension or cardiovascular disor-
The Renal Urate–Anion Exchanger URAT1
ders. The proposed role of urate in the pathogenesis of
Enomoto and colleagues (63) recently identified the
these disorders has recently been reviewed in the Physiol-
molecular target for uricosuric agents (see Glossary), an
ogy in Medicine series (12).
anion exchanger responsible for the reabsorption of filtered
urate by the renal proximal tubule (Table). The authors
RENAL TRANSPORT OF URATE searched the human genome database for novel gene se-
Renal urate transport is typically explained by a quences within the organic anion transporter (OAT) gene
4-component model: glomerular filtration, a near-complete family and identified URAT1 (SLC22A12) (see Glossary),
reabsorption of filtered urate, subsequent secretion, and a novel transporter expressed at the apical brush border of
postsecretory reabsorption in the remaining proximal tu- the proximal nephron (63). Urate–anion exchange activity
bule (see Glossary) (76, 77). This model evolved from an similar to that of URAT1 was initially described in brush-
interpretation of the effects of “uricosuric” and “antiurico- border membrane vesicles from urate-reabsorbing species,
suric” agents; drugs and compounds known to affect serum such as rats and dogs (97–100), and was subsequently con-
urate levels are summarized in the Table. The urate secre- firmed in human kidneys (101). Frog eggs (Xenopus oo-
tion step was incorporated into the model to explain the cytes) injected with URAT1-encoding RNA transport
potent antiuricosuric effect of pyrazinamide (91). How- urate and exhibit pharmacologic properties consistent with
ever, direct inhibition of proximal tubular urate secretion data from human brush-border membrane vesicles (63,
506 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 www.annals.org
Pathogenesis of Gout Review

101). These and other experiments indicate that uricosuric essential for the effect of both uricosuric and antiuricosuric
compounds (for example, probenecid, benzbromarone, agents (see Glossary) (90).
sulfinpyrazone, and losartan) directly inhibit URAT1 from
the apical side of tubular cells (“cis-inhibition” [see Glos- Secondary Sodium Dependency of Urate Reabsorption
sary]). Conversely, antiuricosuric substances (for example, Antiuricosuric agents exert their effect by stimulating re-
pyrazinoate, nicotinate, and lactate) serve as the exchang- nal reabsorption rather than inhibiting tubular secretion (91).
ing anion from inside cells (Figure 6 and Table), thereby The mechanism appears to involve a “priming” of renal urate
stimulating anion exchange and urate reabsorption (“trans- reabsorption through the sodium-dependent loading of prox-
stimulation” [see Glossary]) (9, 63). In addition to urate, imal tubular epithelial cells with anions capable of a trans-
URAT1 has particular affinity for aromatic organic anions, stimulation of urate reabsorption (Figure 6). Studies from sev-
such as nicotinate and pyrazinoate, followed by lactate, eral laboratories have indicated that a transporter in the
␤-hydroxybutyrate, acetoacetate, and inorganic anions, proximal tubule brush border mediates sodium-dependent
such as chloride and nitrate (63). reabsorption of pyrazinoate, nicotinate, lactate, pyruvate,
Enomoto and colleagues (63) provided unequivocal ␤-hydroxybutyrate, and acetoacetate (102–104), monovalent
genetic proof that URAT1 is crucial for urate homeostasis: anions that are also substrates for URAT1 (63). Increased
A handful of patients with “familial renal hypouricemia” plasma concentrations of these antiuricosuric anions result in
(OMIM [Online Mendelian Inheritance in Man] accession their increased glomerular filtration and greater reabsorption
number 220150; see Glossary) were shown to carry loss- by the proximal tubule. The augmented intraepithelial con-
of-function mutations in the human SLC22A12 gene en- centrations in turn induce the reabsorption of urate by pro-
coding URAT1, indicating that this exchanger is essential moting the URAT1-dependent anion exchange of filtered
for proximal tubular reabsorption. Furthermore, pyrazin- urate (trans-stimulation) (Figure 6).
amide, benzbromarone, and probenecid failed to affect Urate reabsorption by the proximal tubule thus exhib-
urate clearance in patients with homozygous loss-of-func- its a form of secondary sodium dependency, in that sodium-
tion mutations in SLC22A12, indicating that URAT1 is dependent loading of proximal tubular cells stimulates

Figure 6. Urate transport mechanisms in human proximal tubule.

Urate transporter-1 (URAT1) is located in the apical membrane of proximal tubular cells in human kidneys and transports urate from lumen to proximal
tubular cells in exchange for anions in order to maintain electrical balance. This exchanger is essential for proximal tubular reabsorption of urate and is
targeted by both uricosuric and antiuricosuric agents. Sodium-dependent entry of monovalent anions (such as pyrazinoate, nicotinate, lactate, pyruvate,
␤-hydroxybutyrate, and acetoacetate), presumptively through the sodium–anion cotransporter, fuels the absorption of luminal urate via the anion
exchanger URAT1. Basolateral entry of urate during urate secretion by the proximal tubule is stimulated by sodium-dependent uptake of the divalent
anion ␣-ketoglutarate, leading to urate-␣-ketoglutarate exchange via organic anion transporter-1 (OAT1) or organic anion transporter-3 (OAT3). These
proteins or similar transporters may facilitate the basolateral influx or efflux of urate. As discussed in the text, although the quantitative role of human
urate secretion remains unclear, several molecular candidates have been proposed for the electrogenic urate secretion pathway in apical membrane of
proximal tubules, including URAT1, ATP-driven efflux pathway (MRP4), and voltage-driven organic anion transporter-1 (OATV1). FEu ⫽ renal
clearance of urate/glomerular filtration rate.
www.annals.org 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 507
Review Pathogenesis of Gout

Figure 7. Dual effects of pyrazinoate on urate transport. 110), and parathyroid hormone (111); URAT1 and the
sodium-dependent anion cotransporter or cotransporters
may be targets for these stimuli.

Dose-Dependent Dual Response in Urate Excretion


A conundrum in the pathophysiology of gout has been
how certain anions can exhibit either uricosuric or antiuri-
cosuric properties, depending on the dose administered.
Monovalent anions that interact with URAT1 have the
dual potential to increase or decrease renal urate excretion
(93, 112) because they can both trans-stimulate and cis-
inhibit apical urate exchange in the proximal tubule (101).
For example, a low concentration of pyrazinoate stimulates
urate reabsorption as a consequence of trans-stimulation,
whereas a higher concentration reduces urate reabsorption
through extracellular cis-inhibition of URAT1 (63, 93,
113) (Figure 7). Dissenting opinions notwithstanding
(114), these observations remain consistent with the basic
scheme of apical urate transport shown in Figure 6. Bipha-
sic effects on urate excretion (that is, antiuricosuria at low
doses and uricosuria at high doses) are particularly well
described for salicylate (115). Salicylate cis-inhibits
URAT1 (63, 116), explaining the high-dose uricosuric ef-
fect; low antiuricosuria reflects a trans-stimulation of
URAT1 by intracellular salicylate, which is evidently a sub-
strate for the sodium–pyrazinoate transporter (103). Min-
The anti-uricosuric agent pyrazinoate (PZA), a metabolite of pyrazin- imal doses of salicylate—75, 150, and 325 mg daily—were
amide, has dual effects on urate transport by the proximal tubule. Urate shown to increase serum uric acid levels by 16, 12, and 2
uptake by brush-border membrane vesicles isolated from canine kidney ␮mol/L, respectively (78). However, the effect on the risk
cortex is shown, in the presence of 100 mM sodium (Na⫹) with 0.1 mM
PZA, 0 PZA, or 5 mM PZA. The concentration results in Na⫹-depen- for gout of this salicylate-induced increase in the serum
dent uptake of PZA and a potentiation of urate uptake via urate trans- uric acid level has not been determined.
porter-1 (URAT1); in contrast, the higher concentration cis-inhibits
URAT1, thus reducing urate uptake by the membrane vesicles. (Repro-
duced with permission from reference 93: Guggino SE, Aronson PS. Other Renal Urate Transporters
Paradoxical effects of pyrazinoate and nicotinate on urate transport in At the basolateral membrane of proximal tubular cells,
dog renal microvillus membranes. J Clin Invest. 1985;76:543-7.) the entry of urate from the surrounding interstitium ap-
pears to be driven by sodium-dependent uptake of divalent
anions, such as ␣-ketoglutarate, rather than monovalent
brush-border urate exchange; urate itself is not a substrate
for the sodium–anion transporter. The molecular identity carboxylates, such as pyrazinoate and lactate (117, 118)
of the relevant sodium-dependent anion cotransporter or (Figure 6). Candidate proteins for this basolateral urate
cotransporters remains unclear; however, a leading candi- exchange activity include both OAT1 (119) and OAT3
date gene is SLC5A8 (see Glossary), which encodes a sodi- (120,⫺ 121) (see Glossary),

each of which function as an-
um-dependent lactate and butyrate cotransporter (105). ion1 – dicarboxylate2 exchangers (121–123) at the baso-
Preliminary data indicate that the SLC5A8 protein can also lateral membrane of the proximal tubule. These proteins
transport both pyrazinoate and nicotinate, potentiating (or similar transporters) conceivably facilitate the basolat-
urate transport in Xenopus oocytes that co-express URAT1 eral influx or efflux of urate.
(106). As mentioned previously, the quantitative role of hu-
The antiuricosuric mechanism explains the long- man urate secretion remains unclear. Nonetheless, several
standing clinical observation that hyperuricemia is induced molecular candidates have been proposed for the electro-
by increased ␤-hydroxybutyrate and acetoacetate levels in genic urate secretion pathway across the apical membrane
diabetic ketoacidosis (95), increased lactic acid levels in of proximal tubules, including the urate transporter/chan-
alcohol intoxication (45), or increased nicotinate and nel (UAT, also known as galectin-9) (124) and the voltage-
pyrazinoate levels in niacin and pyrazinamide therapy, re- driven organic anion transporter-1 (OATV1) (125). The
spectively (96). Urate retention is also known to be pro- apical ATP-driven anion transporter multidrug resistance
voked by a reduction in extracellular fluid volume (107) protein 4 (MRP4) (see Glossary) has also been shown to
and by excesses of angiotensin II (108, 109), insulin (62, mediate urate efflux (126).
508 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 www.annals.org
Pathogenesis of Gout Review
Figure 8. Putative mechanisms for initiation, perpetuation, and termination of an acute monosodium urate crystal–induced gouty
inflammation.

Recent advances in the understanding of acute gouty attack are illustrated (left). The attack is primarily neutrophil-dependent and initiated by the
capacity of urate crystals to activate complements and to stimulate synovial lining cells and resident inflammatory cells to induce a variety of inflammatory
mediators. As depicted (right), self-resolution of acute gout is mediated by several mechanisms, including coating of monosodium urate crystals with
proteins and clearance by differentiated macrophages, neutrophil apoptosis, clearance of apoptotic cells, inactivation of inflammatory mediators, and the
release of anti-inflammatory mediators. Dots represent humoral inflammatory mediators, including cytokines and chemokines. Apo B ⫽ apolipoprotein
B; Apo E ⫽ apolipoprotein E; C1q, C3a, C3b, C5a, C5b-9 ⫽ complement membrane attack complex; IL ⫽ interleukin; LDL ⫽ low-density
lipoprotein; LTB4 ⫽ leukotriene B4; MCP-1 ⫽ monocyte chemoattractant protein-1/CCL2; MIP-1 ⫽ macrophage inflammatory protein-1/CCL3;
MMP-3 ⫽ matrix metalloproteinase-3; NO ⫽ nitrous oxide; PAF ⫽ platelet-activating factor; PGE2 ⫽ prostaglandin E2; PLA2 ⫽ phospholipase A2;
PPAR-␣ ⫽ peroxisome proliferator-activated receptor-␣ ligand; PPAR␥ ⫽ peroxisome proliferator-activated receptor-␥ ligand; TGF-␤ ⫽ transcription
growth factor-␤; TNF-␣ ⫽ tumor necrosis factor-␣; S100A8/A9 ⫽ myeloid-related protein; sTNFr ⫽ soluble tumor necrosis factor receptor.

URATE CRYSTAL–INDUCED INFLAMMATION ticular properties of the urate crystal to interact directly
Urate crystals are directly able to initiate, to amplify, with lipid membranes and proteins through cell membrane
and to sustain an intense inflammatory attack because of perturbation and cross-linking of membrane glycoproteins
their ability to stimulate the synthesis and release of hu- in the phagocyte. This interaction leads to the activation of
moral and cellular inflammatory mediators (Figure 8). several signal transduction pathways, including G proteins,
phospholipase C and D, Src tyrosine kinases, the mitogen-
Urate Crystal–Induced Cell Activation and Signaling activated protein kinases ERK1/ERK2, c-Jun N-terminal
Urate crystals interact with the phagocyte through 2 kinase, and p38 mitogen-activated protein kinase (see
broad mechanisms. First, they activate the cells through Glossary) (127–130). These steps are critical for crystal-
the conventional route as opsonized and phagocytosed induced interleukin (IL)– 8 (see Glossary) expression in
particles, eliciting the stereotypical phagocyte response of monocytic cells (130 –132), which plays a key role in the
lysosomal fusion, respiratory burst, and release of inflam- neutrophil accumulation that is discussed later in this re-
matory mediators. The other mechanism involves the par- view (133).
www.annals.org 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 509
Review Pathogenesis of Gout

Crystal-Induced Cellular Response mately 90% of the neutrophil chemotactic activity of hu-
Cellular kinetic analyses using experimental animal man monocytes in response to urate crystals (133). Neu-
models of gout (134, 135) indicate that monocytes and tralization of IL-8 or its receptor may substantially reduce
mast cells participate during the early phase of inflamma- the IL-8 –induced neutrophilic inflammatory process (148)
tion, whereas neutrophil infiltrates occur later during in- and provide a potential therapeutic target in gout. Several
flammation (Figure 8). Phagocytes from noninflamed other neutrophil chemotactic factors, including the calci-
joints may contain urate crystals (136), and most of these um-binding proteins (calgranulins) S100A8 and S100A9
phagocytes are macrophages (137). The state of differenti- (see Glossary) (149, 150), have also been shown to be
ation of mononuclear phagocytes determines whether the involved in neutrophil migration induced by urate crystals
crystals will trigger an inflammatory response. In less dif- (Figure 8).
ferentiated cell lines, synthesis of tumor necrosis factor–␣
(TNF-␣) (see Glossary) and endothelial cell activation oc- SPONTANEOUS RESOLUTION OF ACUTE GOUT
curred after urate crystal phagocytosis, whereas well-differ-
The self-limited nature of acute gout is thought to
entiated macrophages failed to induce TNF-␣ synthesis or
involve several mechanisms (151), as shown in Figure 8.
to activate endothelial cells (137). Similarly, freshly isolated
Clearance of urate crystals by differentiated macrophages
human monocytes lead to a vigorous response by induction
in vitro has been linked to inhibition of leukocyte and
of TNF-␣, IL-1␤, IL-6, IL-8, and cyclooxygenase-2 secre-
endothelial activation (137, 138, 152). Neutrophil apopto-
tion (see Glossary), whereas human macrophages differen-
sis (see Glossary) and other apoptotic cell clearance repre-
tiated in vitro for 7 days failed to secrete cytokines (see
sent a fundamental mechanism in the resolution of acute
Glossary) or to induce endothelial cell activation (138).
inflammation. Furthermore, transforming growth factor–␤
These findings indicate that monocytes play a central role
(see Glossary) becomes abundant in acute gouty synovial
in stimulating an acute attack of gout, whereas differenti-
fluid and inhibits IL-1 receptor expression and IL-1–
ated macrophages play an anti-inflammatory role in termi-
driven cellular inflammatory responses (153, 154).
nating an acute attack and preserving the asymptomatic
Upregulation of IL-10 expression has been shown to
state (Figure 8).
limit experimental urate-induced inflammation and may
Experimental animal models suggest that mast cells are
function as a native inhibitor of gouty inflammation (155).
involved in the early phase of crystal-induced inflammation
Similarly, urate crystals induce peroxisome proliferator–ac-
(134), and they also release inflammatory mediators, such tivated receptor-␥ (PPAR-␥) (see Glossary) expression in
as histamine (139), in response to C3a, C5a, and IL-1. The human monocytes and promote neutrophil and macro-
vasodilatation, increased vascular permeability, and pain phage apoptosis (156). Research has yet to determine if the
are also mediated by kinins, complement cleavage peptides, PPAR-␥– based therapy currently available for type 2 dia-
and other vasoactive prostaglandins (see Glossary) (140). betes would also be useful in gout management.
Neutrophilic Influx and Amplification Inactivation of inflammatory mediators by proteolytic
Neutrophilic synovitis (see Glossary) is the hallmark cleavage, cross-desensitization of receptors for chemokines,
of an acute gouty attack (Figure 8). Neutrophilic– endo- release of lipoxins (see Glossary), IL-1 receptor antagonist,
thelial cell interaction leading to neutrophilic influx ap- and other anti-inflammatory mediators all facilitate the res-
pears to be an important event in this inflammation and olution of acute gout. As shown in Figure 8, increased
represents a major locus for the pharmacologic effect of vascular permeability allows the entry of large molecules
(such as apolipoproteins B and E [see Glossary]) and other
colchicine. Neutrophil influx is believed to be promoted by
plasma proteins into the synovial cavity, which also con-
the endothelial–neutrophil adhesion that is triggered by
tributes to the spontaneous resolution of acute flares (157,
IL-1, TNF-␣, and several chemokines (see Glossary), such
158).
as IL-8 and neutrophil chemoattractant protein-1 (MCP-
1). Neutrophil migration involves neutrophilic– endothe-
lial interaction mediated by cytokine-induced clustering of CHRONIC GOUTY ARTHRITIS
E-selectin (see Glossary) on endothelial cells. Colchicine Chronic gouty arthritis typically develops in patients
interferes with the interactions by altering the number and who have had gout for years (Figure 9). Cytokines, che-
distribution of selectins on endothelial cells and neutro- mokines, proteases, and oxidants involved in acute urate
phils in response to IL-1 or TNF-␣ (141). crystal–induced inflammation also contribute to the
Once in the synovial tissue, the neutrophils follow chronic inflammation, leading to chronic synovitis, carti-
concentration gradients of chemoattractants such as C5a, lage loss, and bone erosion. Even during remissions of
leukotriene B4 (see Glossary), platelet-activating factor, acute flares, low-grade synovitis in involved joints may per-
IL-1, and IL-8 (142). Among these factors, IL-8 and sist with ongoing intra-articular phagocytosis of crystals
growth-related gene chemokines play a central role in neu- by leukocytes (136). Tophi on the cartilage surface, which
trophil invasion in experimental models of acute gout can be observed through arthroscopy (159), may contrib-
(143–147). For example, IL-8 alone accounts for approxi- ute to chondrolysis despite adequate treatment of both
510 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 www.annals.org
Pathogenesis of Gout Review
Figure 9. Putative mechanisms for chronic monosodium urate–induced inflammation and cartilage and bone destruction.

Low-level inflammation persists during the remissions of acute flares. Cytokines, chemokines, proteases, and oxidants involved in acute inflammation
contribute to chronic inflammation leading to chronic synovitis, cartilage loss, and bone erosion. Monosodium urate (MSU) crystals are able to activate
chondrocytes to release interleukin-1, inducible nitric oxide synthetase, and matrix metalloproteinases, leading to cartilage destruction. Similarly, MSU
crystal activation of osteoblasts, release of cytokines by activated osteoblast, and decreased anabolic function contribute to the juxta-articular bone damage
seen in chronic MSU inflammation. IL ⫽ interleukin; iNOs ⫽ inducible nitrous oxide synthase; MMP-9 ⫽ matrix metalloproteinase-9; PGE2 ⫽
prostaglandin E2.

hyperuricemia and acute gouty attacks (160). Adherent Weight control, limits on red meat consumption, and
chondrocytes phagocytize microcrystals and produce active daily exercise are important foundations of lifestyle modi-
metalloproteinases. Furthermore, crystal– chondrocyte cell fication recommendations for patients with gout or hyper-
membrane interactions can trigger chondrocyte activation, uricemia and parallel recommendations related to preven-
gene expression of IL-1␤ and inducible nitric oxide synthase, tion of coronary heart disease, diabetes, and certain types
nitric oxide release, and the overexpression of matrix metallo- of cancer. Patients with gout could consider using plant-
proteinases (see Glossary) that leads to cartilage destruction derived ␻-3 fatty acids or supplements of eicosapentaenoic
(161). The crystals can also suppress the 1,25-dihydroxychole- acid and docosahexanoic acid instead of consuming fish for
calciferol–induced activity of alkaline phosphatase and osteo- cardiovascular benefits. The recent recommendation on
calcin (see Glossary). Thus, crystals can reduce the anabolic dairy consumption for the general public would also be
effects of osteoblasts, thereby contributing to damage to the applicable for most patients with gout or hyperuricemia
juxta-articular bone (162) (Figure 9). and may offer added benefit to individuals with hyperten-
sion, diabetes, and cardiovascular disorders. Further risk–
SUMMARY benefit assessments in each specific clinical context would
The disease burden of gout remains substantial and be helpful. Daily consumption of nuts and legumes as rec-
may be increasing. As more scientific data on modifiable ommended by the Harvard Healthy Eating Pyramid (32)
risk factors and comorbidities of gout become available, may also provide important health benefits without in-
integration of these data into gout care strategy may be- creasing the risk for gout. Similarly, a daily glass of wine
come essential, similar to the current care strategies for may benefit health without imposing an elevated risk for
hypertension (163) and type 2 diabetes (164). Recommen- gout, especially in contrast to beer or liquor consumption.
dations for lifestyle modification to treat or to prevent gout These lifestyle modifications are inexpensive and safe and,
are generally in line with those for the prevention or treat- when combined with drug therapy, may result in better
ment of other major chronic disorders (32). Thus, the net control of gout.
health benefits from these general healthy lifestyle recom- Effective management of gout risk factors (for exam-
mendations (32) are expected to be even larger among ple, hypertension) and the strategic choice of certain ther-
many patients with gout, particularly those with coexisting apies for comorbid conditions may also aid gout care. For
insulin resistance syndrome, diabetes, obesity, and hyper- example, antihypertensive agents with uricosuric properties
tension. (for example, losartan [165] or amlodipine [86]) could
www.annals.org 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 511
Review Pathogenesis of Gout

have a better risk– benefit ratio than diuretics for hyperten- References
sion in hypertensive patients with gout. Similarly, the uri- 1. Terkeltaub RA. Clinical practice. Gout. N Engl J Med. 2003;349:1647-55.
[PMID: 14573737].
cosuric property of fenofibrate (165) may be associated
2. Schlesinger N, Schumacher HR Jr. Gout: can management be improved?
with a favorable risk– benefit ratio among patients with Curr Opin Rheumatol. 2001;13:240-4. [PMID: 11333356]
gout and the metabolic syndrome. 3. Kramer HM, Curhan G. The association between gout and nephrolithiasis:
The recently elucidated molecular mechanism of renal the National Health and Nutrition Examination Survey III, 1988-1994. Am J
Kidney Dis. 2002;40:37-42. [PMID: 12087559]
urate transport has several important implications in con- 4. Arromdee E, Michet CJ, Crowson CS, O’Fallon WM, Gabriel SE. Epide-
ditions that are associated with high urate levels. In partic- miology of gout: is the incidence rising? J Rheumatol. 2002;29:2403-6. [PMID:
ular, the molecular characterization of the URAT1 anion 12415600]
5. Johnson RJ, Rideout BA. Uric acid and diet—insights into the epidemic of
exchanger has provided a specific target of action for well- cardiovascular disease. N Engl J Med. 2004;350:1071-3. [PMID: 15014177]
known substances affecting urate levels. Genetic variation 6. Wu XW, Lee CC, Muzny DM, Caskey CT. Urate oxidase: primary structure
in these renal transporters or upstream regulatory factors and evolutionary implications. Proc Natl Acad Sci U S A. 1989;86:9412-6.
[PMID: 2594778]
may explain the genetic tendency to develop conditions 7. Wu XW, Muzny DM, Lee CC, Caskey CT. Two independent mutational
associated with high urate levels and a patient’s particular events in the loss of urate oxidase during hominoid evolution. J Mol Evol. 1992;
response to medications. Furthermore, the transporters 34:78-84. [PMID: 1556746]
themselves may serve as targets for future drug develop- 8. Ames BN, Cathcart R, Schwiers E, Hochstein P. Uric acid provides an
antioxidant defense in humans against oxidant- and radical-caused aging and
ment. cancer: a hypothesis. Proc Natl Acad Sci U S A. 1981;78:6858-62. [PMID:
Finally, advances in our understanding of crystal-in- 6947260]
duced inflammation indicate that gout shares many patho- 9. Hediger MA. Kidney function: gateway to a long life? Nature. 2002;417:393,
395. [PMID: 12024201]
genetic features with other chronic inflammatory disorders. 10. Oda M, Satta Y, Takenaka O, Takahata N. Loss of urate oxidase activity in
Some newly available potent anti-inflammatory medica- hominoids and its evolutionary implications. Mol Biol Evol. 2002;19:640-53.
tions (including biological agents that are indicated for [PMID: 11961098]
11. Watanabe S, Kang DH, Feng L, Nakagawa T, Kanellis J, Lan H, et al. Uric
other conditions) may have therapeutic potential in se- acid, hominoid evolution, and the pathogenesis of salt-sensitivity. Hypertension.
lected subsets of patients with gout, although the high costs 2002;40:355-60. [PMID: 12215479]
of biological agents would probably prevent their wide- 12. Oparil S, Zaman MA, Calhoun DA. Pathogenesis of hypertension. Ann
spread use in gout. Anti-inflammatory agents for gout (in- Intern Med. 2003;139:761-76. [PMID: 14597461]
13. Campion EW, Glynn RJ, DeLabry LO. Asymptomatic hyperuricemia. Risks
cluding colchicine) are typically used to treat acute gout or and consequences in the Normative Aging Study. Am J Med. 1987;82:421-6.
to reduce the risk for rebound gout attacks during the [PMID: 3826098]
initiation of urate-lowering therapy but do not lower se- 14. Lin KC, Lin HY, Chou P. The interaction between uric acid level and other
risk factors on the development of gout among asymptomatic hyperuricemic men
rum levels of uric acid. The long-term safety profile of in a prospective study. J Rheumatol. 2000;27:1501-5. [PMID: 10852278]
these agents needs to be clarified, including the potential 15. Shoji A, Yamanaka H, Kamatani N. A retrospective study of the relationship
consequences of chronic hyperuricemia with such anti-in- between serum urate level and recurrent attacks of gouty arthritis: evidence for
reduction of recurrent gouty arthritis with antihyperuricemic therapy. Arthritis
flammatory treatment. Rheum. 2004;51:321-5. [PMID: 15188314]
16. Burt HM, Dutt YC. Growth of monosodium urate monohydrate crystals:
effect of cartilage and synovial fluid components on in vitro growth rates. Ann
From Arthritis Research Centre of Canada, University of British Colum- Rheum Dis. 1986;45:858-64. [PMID: 3098195]
bia, Vancouver, British Columbia, Canada; Massachusetts General Hos- 17. McGill NW, Dieppe PA. The role of serum and synovial fluid components
pital, Brigham and Women’s Hospital, Harvard Medical School, and VA in the promotion of urate crystal formation. J Rheumatol. 1991;18:1042-5.
[PMID: 1717687]
Boston Healthcare System, Boston, Massachusetts.
18. Fam AG, Stein J, Rubenstein J. Gouty arthritis in nodal osteoarthritis. J
Rheumatol. 1996;23:684-9. [PMID: 8730127]
19. Simkin PA, Pizzorno JE. Transynovial exchange of small molecules in nor-
Acknowledgments: The authors thank Dr. John Seeger for his critical mal human subjects. J Appl Physiol. 1974;36:581-7. [PMID: 4826322]
review of the manuscript. 20. Hochberg MC, Silman AJ, Smolen JS, Weinblatt ME, Weisman M.
Rheumatology. 3rd ed. New York: Mosby; 2003.
21. Griebsch A, Zollner N. Effect of ribomononucleotides given orally on uric
Potential Financial Conflicts of Interest: Consultancies: H.K. Choi acid production in man. Adv Exp Med Biol. 1974;41:443-9. [PMID: 4832569]
22. Coe FL, Moran E, Kavalich AG. The contribution of dietary purine over-
(TAP Pharmaceutical Products); Honoraria: H.K. Choi (TAP Pharma-
consumption to hyperpuricosuria in calcium oxalate stone formers. J Chronic
ceutical Products); Grants received: H.K. Choi (TAP Pharmaceutical Dis. 1976;29:793-800. [PMID: 1010873]
Products). 23. Gibson T, Rodgers AV, Simmonds HA, Court-Brown F, Todd E, Meilton
V. A controlled study of diet in patients with gout. Ann Rheum Dis. 1983;42:
123-7. [PMID: 6847259]
Requests for Single Reprints: Hyon K. Choi, MD, DrPH, Division of 24. Zollner N, Griebsch A. Diet and gout. Adv Exp Med Biol. 1974;41:435-42.
Rheumatology, Department of Medicine, University of British Colum- [PMID: 4832568]
bia, Arthritis Research Centre of Canada, 895 West 10th Avenue, Van- 25. Clifford AJ, Riumallo JA, Young VR, Scrimshaw NS. Effects of oral purines
couver, BC V5Z 1L7; e-mail, hchoi@partners.org. on serum and urinary uric acid of normal, hyperuricaemic and gouty humans
[Abstract]. J Nutr. 1976;106:428-50.
26. Watson AR, Simmonds HA, Webster DR, Layward L, Evans DI. Purine
Current author addresses are available at www.annals.org. nucleoside phosphorylase (PNP) deficiency: a therapeutic challenge. Adv Exp

512 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 www.annals.org
Pathogenesis of Gout Review
Med Biol. 1984;165 Pt A:53-9. [PMID: 6426259] Dis. 1998;57:509-10. [PMID: 9849306]
27. Choi HK, Atkinson K, Karlson EW, Willett W, Curhan G. Purine-rich 55. Lee J, Sparrow D, Vokonas PS, Landsberg L, Weiss ST. Uric acid and
foods, dairy and protein intake, and the risk of gout in men. N Engl J Med. coronary heart disease risk: evidence for a role of uric acid in the obesity-insulin
2004;350:1093-103. [PMID: 15014182] resistance syndrome. The Normative Aging Study. Am J Epidemiol. 1995;142:
28. Choi HK, Atkinson K, Karlson EW, Curhan G. Obesity, weight change, 288-94. [PMID: 7631632]
hypertension, diuretic use, and risk of gout in men: the Health Professionals 56. Rathmann W, Funkhouser E, Dyer AR, Roseman JM. Relations of
Follow-up Study. Arch Intern Med. 2005;165:742-8. [PMID: 15824292] hyperuricemia with the various components of the insulin resistance syn-
29. Emmerson BT. The management of gout. N Engl J Med. 1996;334:445-51. drome in young black and white adults: the CARDIA study. Coronary
[PMID: 8552148] Artery Risk Development in Young Adults. Ann Epidemiol. 1998;8:250-
30. Fam AG. Gout, diet, and the insulin resistance syndrome [Editorial]. J Rheu- 61. [PMID: 9590604]
matol. 2002;29:1350-5. [PMID: 12136887] 57. Roubenoff R, Klag MJ, Mead LA, Liang KY, Seidler AJ, Hochberg MC.
31. Choi HK, Atkinson K, Karlson EW, Willett W, Curhan G. Alcohol intake Incidence and risk factors for gout in white men. JAMA. 1991;266:3004-7.
and risk of incident gout in men: a prospective study. Lancet. 2004;363:1277-81. [PMID: 1820473]
[PMID: 15094272] 58. Dessein PH, Shipton EA, Stanwix AE, Joffe BI, Ramokgadi J. Beneficial
32. Willett WC, Stampfer MJ. Rebuilding the food pyramid. Sci Am. 2003;288: effects of weight loss associated with moderate calorie/carbohydrate restriction,
64-71. [PMID: 12506426] and increased proportional intake of protein and unsaturated fat on serum urate
33. Klippel JH. Primer on the Rheumatic Diseases. 12th ed. Atlanta, GA: Ar- and lipoprotein levels in gout: a pilot study. Ann Rheum Dis. 2000;59:539-43.
thritis Foundation; 2001. [PMID: 10873964]
34. Koopman WJ. Arthritis & Allied Conditions: A Textbook of Rheumatology. 59. Yamashita S, Matsuzawa Y, Tokunaga K, Fujioka S, Tarui S. Studies on the
12th ed. New York: Lippincott Williams & Wilkins; 2001. impaired metabolism of uric acid in obese subjects: marked reduction of renal
35. Woolliscroft JO, Colfer H, Fox IH. Hyperuricemia in acute illness: a poor urate excretion and its improvement by a low-calorie diet. Int J Obes. 1986;10:
prognostic sign. Am J Med. 1982;72:58-62. [PMID: 7058824] 255-64. [PMID: 3771090]
36. Woolliscroft JO, Fox IH. Increased body fluid purine levels during hypoten- 60. Emmerson BT. Alteration of urate metabolism by weight reduction. Aust N
sive events. Evidence for ATP degradation. Am J Med. 1986;81:472-8. [PMID: Z J Med. 1973;3:410-2. [PMID: 4519128]
3752148] 61. Ter Maaten JC, Voorburg A, Heine RJ, Ter Wee PM, Donker AJ,
37. Mineo I, Kono N, Hara N, Shimizu T, Yamada Y, Kawachi M, et al. Gans RO. Renal handling of urate and sodium during acute physiological
Myogenic hyperuricemia. A common pathophysiologic feature of glycogenosis hyperinsulinaemia in healthy subjects. Clin Sci (Lond). 1997;92:51-8.
types III, V, and VII. N Engl J Med. 1987;317:75-80. [PMID: 3473284] [PMID: 9038591]
38. Fox IH. Adenosine triphosphate degradation in specific disease. J Lab Clin 62. Muscelli E, Natali A, Bianchi S, Bigazzi R, Galvan AQ, Sironi AM, et al.
Med. 1985;106:101-10. [PMID: 3860585] Effect of insulin on renal sodium and uric acid handling in essential hypertension.
39. Jinnai K, Kono N, Yamamoto Y, Kanda F, Ohno S, Tsutsumi M, et al. Am J Hypertens. 1996;9:746-52. [PMID: 8862220]
Glycogenosis type V (McArdle’s disease) with hyperuricemia. A case report and 63. Enomoto A, Kimura H, Chairoungdua A, Shigeta Y, Jutabha P, Cha SH,
clinical investigation. Eur Neurol. 1993;33:204-7. [PMID: 8467838] et al. Molecular identification of a renal urate anion exchanger that regulates
40. Yamanaka H, Kawagoe Y, Taniguchi A, Kaneko N, Kimata S, Hosoda S, et blood urate levels. Nature. 2002;417:447-52. [PMID: 12024214]
al. Accelerated purine nucleotide degradation by anaerobic but not by aerobic 64. Bedir A, Topbas M, Tanyeri F, Alvur M, Arik N. Leptin might be a
ergometer muscle exercise. Metabolism. 1992;41:364-9. [PMID: 1556942] regulator of serum uric acid concentrations in humans. Jpn Heart J. 2003;44:
41. Faller J, Fox IH. Ethanol-induced hyperuricemia: evidence for increased 527-36. [PMID: 12906034]
urate production by activation of adenine nucleotide turnover. N Engl J Med. 65. Fruehwald-Schultes B, Peters A, Kern W, Beyer J, Pfutzner A. Serum leptin
1982;307:1598-602. [PMID: 7144847] is associated with serum uric acid concentrations in humans. Metabolism. 1999;
42. Fox IH, Kelley WN. Studies on the mechanism of fructose-induced hyper- 48:677-80. [PMID: 10381138]
uricemia in man. Metabolism. 1972;21:713-21. [PMID: 5047915] 66. Bakker SJ, Gans RO, ter Maaten JC, Teerlink T, Westerhoff HV, Heine
43. Raivio KO, Becker A, Meyer LJ, Greene ML, Nuki G, Seegmiller JE. RJ. The potential role of adenosine in the pathophysiology of the insulin resis-
Stimulation of human purine synthesis de novo by fructose infusion. Metabo- tance syndrome. Atherosclerosis. 2001;155:283-90. [PMID: 11254897]
lism. 1975;24:861-9. [PMID: 166270] 67. Balakrishnan VS, Coles GA, Williams JD. Effects of intravenous adenosine
44. Puig JG, Fox IH. Ethanol-induced activation of adenine nucleotide on renal function in healthy human subjects. Am J Physiol. 1996;271:F374-81.
turnover. Evidence for a role of acetate. J Clin Invest. 1984;74:936-41. [PMID: 8770169]
[PMID: 6470146] 68. Balakrishnan VS, Coles GA, Williams JD. A potential role for endogenous
45. Lieber CS, Jones DP, Losowsky MS, Davidson CS. Interrelation of uric acid adenosine in control of human glomerular and tubular function. Am J Physiol.
and ethanol metabolism in man. J Clin Invest. 1962;41:1863-70. [PMID: 1993;265:F504-10. [PMID: 8238379]
13930523] 69. Fransen R, Koomans HA. Adenosine and renal sodium handling: direct
46. Choi HK, Curhan G. Beer, liquor, and wine consumption and serum uric natriuresis and renal nerve-mediated antinatriuresis. J Am Soc Nephrol. 1995;6:
acid level: the Third National Health and Nutrition Examination Survey. Arthri- 1491-7. [PMID: 8589328]
tis Rheum. 2004;51:1023-9. [PMID: 15593346] 70. Flegal KM, Carroll MD, Ogden CL, Johnson CL. Prevalence and trends in
47. Gibson T, Rodgers AV, Simmonds HA, Toseland P. Beer drinking and its obesity among US adults, 1999-2000. JAMA. 2002;288:1723-7. [PMID:
effect on uric acid. Br J Rheumatol. 1984;23:203-9. [PMID: 6743968] 12365955]
48. Fox IH, Palella TD, Kelley WN. Hyperuricemia: a marker for cell energy 71. Freedman DS, Khan LK, Serdula MK, Galuska DA, Dietz WH. Trends
crisis [Editorial]. N Engl J Med. 1987;317:111-2. [PMID: 3473283] and correlates of class 3 obesity in the United States from 1990 through 2000.
49. Emmerson BT. Effect of oral fructose on urate production. Ann Rheum Dis. JAMA. 2002;288:1758-61. [PMID: 12365960]
1974;33:276-80. [PMID: 4843132] 72. Ford ES, Giles WH, Dietz WH. Prevalence of the metabolic syndrome
50. Stirpe F, Della Corte E, Bonetti E, Abbondanza A, Abbati A, De Stefano F. among US adults: findings from the third National Health and Nutrition Exam-
Fructose-induced hyperuricaemia. Lancet. 1970;2:1310-1. [PMID: 4098798] ination Survey. JAMA. 2002;287:356-9. [PMID: 11790215]
51. Gross LS, Li L, Ford ES, Liu S. Increased consumption of refined carbohy- 73. Wyngaarden JB, Kelley WN. Gout and Hyperuricemia. New York: Grune
drates and the epidemic of type 2 diabetes in the United States: an ecologic & Stratton; 1976.
assessment. Am J Clin Nutr. 2004;79:774-9. [PMID: 15113714] 74. Prebis JW, Gruskin AB, Polinsky MS, Baluarte HJ. Uric acid in childhood
52. Bray GA, Nielsen SJ, Popkin BM. Consumption of high-fructose corn syrup essential hypertension. J Pediatr. 1981;98:702-7. [PMID: 7229748]
in beverages may play a role in the epidemic of obesity. Am J Clin Nutr. 2004; 75. Messerli FH, Frohlich ED, Dreslinski GR, Suarez DH, Aristimuno GG.
79:537-43. [PMID: 15051594] Serum uric acid in essential hypertension: an indicator of renal vascular involve-
53. Glynn RJ, Campion EW, Silbert JE. Trends in serum uric acid levels 1961– ment. Ann Intern Med. 1980;93:817-21. [PMID: 7447188]
1980. Arthritis Rheum. 1983;26:87-93. [PMID: 6824508] 76. Diamond HS, Paolino JS. Evidence for a postsecretory reabsorptive site for
54. Emmerson B. Hyperlipidaemia in hyperuricaemia and gout. Ann Rheum uric acid in man. J Clin Invest. 1973;52:1491-9. [PMID: 4703233]

www.annals.org 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 513
Review Pathogenesis of Gout

77. Giebisch G, Windhager E. Transport of urea, glucose, phosphate, calcium, 101. Roch-Ramel F, Werner D, Guisan B. Urate transport in brush-border
magnesium, and organic solutes. In: Boron W, Boulpaep E, eds. Medical Phys- membrane of human kidney. Am J Physiol. 1994;266:F797-805. [PMID:
iology. Philadelphia: WB Saunders; 2003:790-813. 8203564]
78. Caspi D, Lubart E, Graff E, Habot B, Yaron M, Segal R. The effect of 102. Garcia ML, Benavides J, Valdivieso F. Ketone body transport in renal
mini-dose aspirin on renal function and uric acid handling in elderly patients. brush border membrane vesicles. Biochim Biophys Acta. 1980;600:922-30.
Arthritis Rheum. 2000;43:103-8. [PMID: 10643705] [PMID: 7407151]
79. Steele TH. Evidence for altered renal urate reabsorption during changes in 103. Manganel M, Roch-Ramel F, Murer H. Sodium-pyrazinoate cotransport in
volume of the extracellular fluid. J Lab Clin Med. 1969;74:288-99. [PMID: rabbit renal brush border membrane vesicles. Am J Physiol. 1985;249:F400-8.
5799512] [PMID: 4037092]
80. Steele TH, Oppenheimer S. Factors affecting urate excretion following di- 104. Boumendil-Podevin EF, Podevin RA. Nicotinic acid transport by brush
uretic administration in man. Am J Med. 1969;47:564-74. [PMID: 4309843] border membrane vesicles from rabbit kidney. Am J Physiol. 1981;240:F185-91.
81. Lin HY, Rocher LL, McQuillan MA, Schmaltz S, Palella TD, Fox IH. [PMID: 7212065]
Cyclosporine-induced hyperuricemia and gout. N Engl J Med. 1989;321:287- 105. Coady MJ, Chang MH, Charron FM, Plata C, Wallendorff B, Sah JF, et
92. [PMID: 2664517] al. The human tumour suppressor gene SLC5A8 expresses a Na⫹-monocarboxy-
82. Gupta AK, Rocher LL, Schmaltz SP, Goldfarb MT, Brown MD, Ellis CN, late cotransporter. J Physiol. 2004;557:719-31. [PMID: 15090606]
et al. Short-term changes in renal function, blood pressure, and electrolyte levels 106. Zandi-Nejad K, Plata C, Enck AH, Mercado A, Romero MF, Mount DB.
in patients receiving cyclosporine for dermatologic disorders. Arch Intern Med. Slc5a8 functions as a sodium-dependent pyrazinoate and nicotinate cotrans-
1991;151:356-62. [PMID: 1992963]. porter; implications for renal urate transport. J Am Soc Nephrol.
83. Hansen JM, Fogh-Andersen N, Leyssac PP, Strandgaard S. Glomerular and 2004;15:89A.
tubular function in renal transplant patients treated with and without ciclosporin 107. Weinman EJ, Eknoyan G, Suki WN. The influence of the extracellular
A. Nephron. 1998;80:450-7. [PMID: 9832645] fluid volume on the tubular reabsorption of uric acid. J Clin Invest. 1975;55:
84. Clive DM. Renal transplant-associated hyperuricemia and gout. J Am Soc 283-91. [PMID: 1127100]
Nephrol. 2000;11:974-9. [PMID: 10770978] 108. Ferris TF, Gorden P. Effect of angiotensin and norepinephrine upon urate
85. Ahn KJ, Kim YS, Lee HC, Park K, Huh KB. Cyclosporine-induced hyper- clearance in man. Am J Med. 1968;44:359-65. [PMID: 4295950]
uricemia after renal transplant: clinical characteristics and mechanisms. Trans- 109. Moriwaki Y, Yamamoto T, Tsutsumi Z, Takahashi S, Hada T. Effects of
plant Proc. 1992;24:1391-2. [PMID: 1496597] angiotensin II infusion on renal excretion of purine bases and oxypurinol. Me-
86. Chanard J, Toupance O, Lavaud S, Hurault de Ligny B, Bernaud C, tabolism. 2002;51:893-5. [PMID: 12077737]
Moulin B. Amlodipine reduces cyclosporin-induced hyperuricaemia in hyperten- 110. Quinones Galvan A, Natali A, Baldi S, Frascerra S, Sanna G, Ciociaro D,
sive renal transplant recipients. Nephrol Dial Transplant. 2003;18:2147-53. et al. Effect of insulin on uric acid excretion in humans. Am J Physiol. 1995;268:
E1-5. [PMID: 7840165]
[PMID: 13679494]
111. Mintz DH, Canary JJ, Carreon G, Kyle LH. Hyperuricemia in hyperpara-
87. Starzl TE, Fung J, Jordan M, Shapiro R, Tzakis A, McCauley J, et al.
thyroidism. N Engl J Med. 1961;265:112-5. [PMID: 13771118]
Kidney transplantation under FK 506. JAMA. 1990;264:63-7. [PMID:
112. Kahn AM, Weinman EJ. Urate transport in the proximal tubule: in vivo
1693970]
and vesicle studies. Am J Physiol. 1985;249:F789-98. [PMID: 3000189]
88. Boots JM, van Duijnhoven EM, Christiaans MH, Nieman FH, van Suylen
113. Fanelli GM Jr, Weiner IM. Pyrazinoate excretion in the chimpanzee. Re-
RJ, van Hooff JP. Single-center experience with tacrolimus versus cyclosporine-
lation to urate disposition and the actions of uricosuric drugs. J Clin Invest.
Neoral in renal transplant recipients. Transpl Int. 2001;14:370-83. [PMID:
1973;52:1946-57. [PMID: 4719671]
11793034]
114. Steele TH. Hyperuricemic nephropathies. Nephron. 1999;81 Suppl 1:45-9.
89. Reyes AJ. Cardiovascular drugs and serum uric acid. Cardiovasc Drugs Ther.
[PMID: 9873214]
2003;17:397-414. [PMID: 15107595]
115. Yu TF, Gutman AB. Study of the paradoxical effects of salicylate in low,
90. Ichida K, Hosoyamada M, Hisatome I, Enomoto A, Hikita M, Endou H,
intermediate and high dosage on the renal mechanisms for excretion of urate in
et al. Clinical and molecular analysis of patients with renal hypouricemia in man. J Clin Invest. 1959;38:1298-315. [PMID: 13673086]
Japan—influence of URAT1 gene on urinary urate excretion. J Am Soc Nephrol. 116. Roch-Ramel F, Guisan B, Diezi J. Effects of uricosuric and antiuricosuric
2004;15:164-73. [PMID: 14694169] agents on urate transport in human brush-border membrane vesicles. J Pharmacol
91. Roch-Ramel F, Guisan B. Renal transport of urate in humans. News Physiol Exp Ther. 1997;280:839-45. [PMID: 9023298]
Sci. 1999;14:80-84. [PMID: 11390825] 117. Kahn AM, Shelat H, Weinman EJ. Urate and p-aminohippurate transport
92. Simmonds HA, Hatfield PJ, Cameron JS, Cadenhead A. Uric acid excretion in rat renal basolateral vesicles. Am J Physiol. 1985;249:F654-61. [PMID:
by the pig kidney. Am J Physiol. 1976;230:1654-61. [PMID: 7142] 4061653]
93. Guggino SE, Aronson PS. Paradoxical effects of pyrazinoate and nicotinate 118. Werner D, Roch-Ramel F. Indirect Na⫹ dependency of urate and p-
on urate transport in dog renal microvillus membranes. J Clin Invest. 1985;76: aminohippurate transport in pig basolateral membrane vesicles. Am J Physiol.
543-7. [PMID: 4031062] 1991;261:F265-72. [PMID: 1877650]
94. Roch-Ramel F, Guisan B, Schild L. Indirect coupling of urate and p-amino- 119. Ichida K, Hosoyamada M, Kimura H, Takeda M, Utsunomiya Y, Hosoya
hippurate transport to sodium in human brush-border membrane vesicles. Am J T, et al. Urate transport via human PAH transporter hOAT1 and its gene struc-
Physiol. 1996;270:F61-8. [PMID: 8769823] ture. Kidney Int. 2003;63:143-55. [PMID: 12472777]
95. Padova J, Bendersky G. Hyperuricemia in diabetic ketoacidosis. N Engl J 120. Cha SH, Sekine T, Fukushima JI, Kanai Y, Kobayashi Y, Goya T, et al.
Med. 1962;267:530-4. [PMID: 14483098] Identification and characterization of human organic anion transporter 3 express-
96. Gershon SL, Fox IH. Pharmacologic effects of nicotinic acid on human ing predominantly in the kidney. Mol Pharmacol. 2001;59:1277-86. [PMID:
purine metabolism. J Lab Clin Med. 1974;84:179-86. [PMID: 4367231] 11306713]
97. Blomstedt JW, Aronson PS. pH gradient-stimulated transport of urate and 121. Bakhiya A, Bahn A, Burckhardt G, Wolff N. Human organic anion trans-
p-aminohippurate in dog renal microvillus membrane vesicles. J Clin Invest. porter 3 (hOAT3) can operate as an exchanger and mediate secretory urate flux.
1980;65:931-4. [PMID: 7358852] Cell Physiol Biochem. 2003;13:249-56. [PMID: 14586168]
98. Guggino SE, Martin GJ, Aronson PS. Specificity and modes of the anion 122. Sweet DH, Chan LM, Walden R, Yang XP, Miller DS, Pritchard JB.
exchanger in dog renal microvillus membranes. Am J Physiol. 1983;244:F612- Organic anion transporter 3 (Slc22a8) is a dicarboxylate exchanger indirectly
21. [PMID: 6859253] coupled to the Na⫹ gradient. Am J Physiol Renal Physiol. 2003;284:F763-9.
99. Kahn AM, Aronson PS. Urate transport via anion exchange in dog renal [PMID: 12488248]
microvillus membrane vesicles. Am J Physiol. 1983;244:F56-63. [PMID: 123. Aslamkhan A, Han YH, Walden R, Sweet DH, Pritchard JB. Stoichiom-
6849384] etry of organic anion/dicarboxylate exchange in membrane vesicles from rat renal
100. Kahn AM, Branham S, Weinman EJ. Mechanism of urate and p-amino- cortex and hOAT1-expressing cells. Am J Physiol Renal Physiol. 2003;285:F775-
hippurate transport in rat renal microvillus membrane vesicles. Am J Physiol. 83. [PMID: 12837685]
1983;245:F151-8. [PMID: 6309010] 124. Lipkowitz MS, Leal-Pinto E, Rappoport JZ, Najfeld V, Abramson RG.

514 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 www.annals.org
Pathogenesis of Gout Review
Functional reconstitution, membrane targeting, genomic structure, and chromo- 141. Cronstein BN, Molad Y, Reibman J, Balakhane E, Levin RI, Weissmann
somal localization of a human urate transporter. J Clin Invest. 2001;107:1103- G. Colchicine alters the quantitative and qualitative display of selectins on endo-
15. [PMID: 11342574] thelial cells and neutrophils. J Clin Invest. 1995;96:994-1002.
125. Jutabha P, Kanai Y, Hosoyamada M, Chairoungdua A, Kim do K, Iribe [PMID: 7543498]
Y, et al. Identification of a novel voltage-driven organic anion transporter present 142. Springer TA. Traffic signals for lymphocyte recirculation and leukocyte
at apical membrane of renal proximal tubule. J Biol Chem. 2003;278:27930-8. emigration: the multistep paradigm. Cell. 1994;76:301-14. [PMID: 7507411]
[PMID: 12740363] 143. Terkeltaub R, Baird S, Sears P, Santiago R, Boisvert W. The murine
126. Van Aubel RA, Smeets PH, van den Heuvel JJ, Russel FG. Human homolog of the interleukin-8 receptor CXCR-2 is essential for the occurrence of
organic anion transporter MRP4 (ABCC4) is an efflux pump for the purine end neutrophilic inflammation in the air pouch model of acute urate crystal-induced
metabolite urate with multiple allosteric substrate binding sites. Am J Physiol gouty synovitis. Arthritis Rheum. 1998;41:900-9. [PMID: 9588743]
Renal Physiol. 2005;288:F327-33. [PMID: 15454390] 144. Fujiwara K, Ohkawara S, Takagi K, Yoshinaga M, Matsukawa A. Involve-
127. Terkeltaub RA, Sklar LA, Mueller H. Neutrophil activation by inflamma- ment of CXC chemokine growth-related oncogene-alpha in monosodium urate
tory microcrystals of monosodium urate monohydrate utilizes pertussis toxin- crystal-induced arthritis in rabbits. Lab Invest. 2002;82:1297-304. [PMID:
insensitive and -sensitive pathways. J Immunol. 1990;144:2719-24. [PMID: 12379764]
2108211] 145. Matsukawa A, Miyazaki S, Maeda T, Tanase S, Feng L, Ohkawara S, et
al. Production and regulation of monocyte chemoattractant protein-1 in lipo-
128. Bomalaski JS, Baker DG, Brophy LM, Clark MA. Monosodium urate
polysaccharide- or monosodium urate crystal-induced arthritis in rabbits: roles of
crystals stimulate phospholipase A2 enzyme activities and the synthesis of a
tumor necrosis factor alpha, interleukin-1, and interleukin-8. Lab Invest. 1998;
phospholipase A2-activating protein. J Immunol. 1990;145:3391-7. [PMID:
78:973-85. [PMID: 9714185]
2230125]
146. Matsukawa A, Yoshimura T, Maeda T, Takahashi T, Ohkawara S, Yo-
129. Gaudry M, Gilbert C, Barabe F, Poubelle PE, Naccache PH. Activation of
shinaga M. Analysis of the cytokine network among tumor necrosis factor alpha,
Lyn is a common element of the stimulation of human neutrophils by soluble interleukin-1beta, interleukin-8, and interleukin-1 receptor antagonist in
and particulate agonists. Blood. 1995;86:3567-74. [PMID: 7579465] monosodium urate crystal-induced rabbit arthritis. Lab Invest. 1998;78:559-69.
130. Liu R, O’Connell M, Johnson K, Pritzker K, Mackman N, Terkeltaub R. [PMID: 9605181]
Extracellular signal-regulated kinase 1/extracellular signal-regulated kinase 2 mi- 147. Jaramillo M, Godbout M, Naccache PH, Olivier M. Signaling events
togen-activated protein kinase signaling and activation of activator protein 1 and involved in macrophage chemokine expression in response to monosodium urate
nuclear factor kappaB transcription factors play central roles in interleukin-8 crystals. J Biol Chem. 2004;279:52797-805. [PMID: 15471869]
expression stimulated by monosodium urate monohydrate and calcium pyro- 148. Nishimura A, Akahoshi T, Takahashi M, Takagishi K, Itoman M, Kondo
phosphate crystals in monocytic cells. Arthritis Rheum. 2000;43:1145-55. H, et al. Attenuation of monosodium urate crystal-induced arthritis in rabbits by
[PMID: 10817569] a neutralizing antibody against interleukin-8. J Leukoc Biol. 1997;62:444-9.
131. Barabe F, Gilbert C, Liao N, Bourgoin SG, Naccache PH. Crystal-induced [PMID: 9335313]
neutrophil activation VI. Involvement of FcgammaRIIIB (CD16) and CD11b in 149. Rouleau P, Vandal K, Ryckman C, Poubelle PE, Boivin A, Talbot M, et
response to inflammatory microcrystals. FASEB J. 1998;12:209-20. [PMID: al. The calcium-binding protein S100A12 induces neutrophil adhesion, migra-
9472986] tion, and release from bone marrow in mouse at concentrations similar to those
132. Liu R, Aupperle K, Terkeltaub R. Src family protein tyrosine kinase sig- found in human inflammatory arthritis. Clin Immunol. 2003;107:46-54.
naling mediates monosodium urate crystal-induced IL-8 expression by monocytic [PMID: 12738249]
THP-1 cells. J Leukoc Biol. 2001;70:961-8. [PMID: 11739559] 150. Ryckman C, McColl SR, Vandal K, de Medicis R, Lussier A, Poubelle PE,
133. Terkeltaub R, Zachariae C, Santoro D, Martin J, Peveri P, Matsushima et al. Role of S100A8 and S100A9 in neutrophil recruitment in response to
K. Monocyte-derived neutrophil chemotactic factor/interleukin-8 is a potential monosodium urate monohydrate crystals in the air-pouch model of acute gouty
mediator of crystal-induced inflammation. Arthritis Rheum. 1991;34:894-903. arthritis. Arthritis Rheum. 2003;48:2310-20. [PMID: 12905486]
[PMID: 2059236] 151. Ortiz-Bravo E, Sieck MS, Schumacher HR Jr. Changes in the proteins
134. Schiltz C, Liote F, Prudhommeaux F, Meunier A, Champy R, Callebert J, coating monosodium urate crystals during active and subsiding inflammation.
et al. Monosodium urate monohydrate crystal-induced inflammation in vivo: Immunogold studies of synovial fluid from patients with gout and of fluid ob-
quantitative histomorphometric analysis of cellular events. Arthritis Rheum. tained using the rat subcutaneous air pouch model. Arthritis Rheum. 1993;36:
2002;46:1643-50. [PMID: 12115197] 1274-85. [PMID: 8216421]
135. Tramontini NL, Kuipers PJ, Huber CM, Murphy K, Naylor KB, Broady 152. Haskard DO, Landis RC. Interactions between leukocytes and endothelial
AJ, et al. Modulation of leukocyte recruitment and IL-8 expression by the mem- cells in gout: lessons from a self-limiting inflammatory response. Arthritis Res.
brane attack complex of complement (C5b-9) in a rabbit model of antigen- 2002;4 Suppl 3:S91-7. [PMID: 12110127]
induced arthritis. Inflammation. 2002;26:311-9. [PMID: 12546141] 153. Liote F, Prudhommeaux F, Schiltz C, Champy R, Herbelin A, Ortiz-
Bravo E, et al. Inhibition and prevention of monosodium urate monohydrate
136. Pascual E, Batlle-Gualda E, Martinez A, Rosas J, Vela P. Synovial fluid
crystal-induced acute inflammation in vivo by transforming growth factor beta1.
analysis for diagnosis of intercritical gout. Ann Intern Med. 1999;131:756-9.
Arthritis Rheum. 1996;39:1192-8. [PMID: 8670330]
[PMID: 10577299]
154. Huynh ML, Fadok VA, Henson PM. Phosphatidylserine-dependent inges-
137. Yagnik DR, Hillyer P, Marshall D, Smythe CD, Krausz T, Haskard DO,
tion of apoptotic cells promotes TGF-beta1 secretion and the resolution of in-
et al. Noninflammatory phagocytosis of monosodium urate monohydrate crystals
flammation. J Clin Invest. 2002;109:41-50. [PMID: 11781349]
by mouse macrophages. Implications for the control of joint inflammation in 155. Murakami Y, Akahoshi T, Kawai S, Inoue M, Kitasato H. Antiinflamma-
gout. Arthritis Rheum. 2000;43:1779-89. [PMID: 10943868] tory effect of retrovirally transfected interleukin-10 on monosodium urate mono-
138. Landis RC, Yagnik DR, Florey O, Philippidis P, Emons V, Mason JC, hydrate crystal-induced acute inflammation in murine air pouches. Arthritis
et al. Safe disposal of inflammatory monosodium urate monohydrate crystals by Rheum. 2002;46:2504-13. [PMID: 12355499]
differentiated macrophages. Arthritis Rheum. 2002;46:3026-33. [PMID: 156. Akahoshi T, Namai R, Murakami Y, Watanabe M, Matsui T, Nishimura
12428246] A, et al. Rapid induction of peroxisome proliferator-activated receptor gamma
139. Getting SJ, Flower RJ, Parente L, de Medicis R, Lussier A, Woliztky BA, expression in human monocytes by monosodium urate monohydrate crystals.
et al. Molecular determinants of monosodium urate crystal-induced murine peri- Arthritis Rheum. 2003;48:231-9. [PMID: 12528124]
tonitis: a role for endogenous mast cells and a distinct requirement for endo- 157. Terkeltaub R, Curtiss LK, Tenner AJ, Ginsberg MH. Lipoproteins con-
thelial-derived selectins. J Pharmacol Exp Ther. 1997;283:123-30. [PMID: taining apoprotein B are a major regulator of neutrophil responses to mono-
9336316] sodium urate crystals. J Clin Invest. 1984;73:1719-30. [PMID: 6725556]
140. Webster ME, Maling HM, Zweig MH, Williams MA, Anderson W Jr. 158. Terkeltaub RA, Dyer CA, Martin J, Curtiss LK. Apolipoprotein (apo) E
Urate crystal induced inflammation in the rat: evidence for the combined actions inhibits the capacity of monosodium urate crystals to stimulate neutrophils.
of kinins, histamine and components of complement. Immunol Commun. 1972; Characterization of intraarticular apo E and demonstration of apo E binding to
1:185-98. [PMID: 4663514] urate crystals in vivo. J Clin Invest. 1991;87:20-6. [PMID: 1985096]

www.annals.org 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 515
Review Pathogenesis of Gout

159. Yu KH. Intraarticular tophi in a joint without a previous gouty attack. J in vitro: synergism with IL-1 to overexpress cyclooxygenase-2. J Immunol. 2002;
Rheumatol. 2003;30:1868-70. [PMID: 12913949] 168:5310-7. [PMID: 11994489]
160. McCarthy GM, Barthelemy CR, Veum JA, Wortmann RL. Influence of 163. August P. Initial treatment of hypertension. N Engl J Med. 2003;348:
antihyperuricemic therapy on the clinical and radiographic progression of gout. 610-7. [PMID: 12584370]
Arthritis Rheum. 1991;34:1489-94. [PMID: 1747133] 164. Kasper DL, Braunwald E, Fauci AS, Hauser SL, Longo DL, Jameson JL.
161. Liu R, Liote F, Rose DM, Merz D, Terkeltaub R. Proline-rich tyrosine Harrison’s Principles of Internal Medicine. 16th ed. New York: McGraw-Hill;
kinase 2 and Src kinase signaling transduce monosodium urate crystal-induced 2004.
nitric oxide production and matrix metalloproteinase 3 expression in chondro- 165. Tokahashi S, Moriwaki Y, Yamamoto T, Tsutsumi Z, Ka T, Fukuchi M.
cytes. Arthritis Rheum. 2004;50:247-58. [PMID: 14730623] Effects of combination treatment using anti-hyperuricaemic agents with fenofi-
162. Bouchard L, de Medicis R, Lussier A, Naccache PH, Poubelle PE. Inflam- brate and/or losartan on uric acid metabolism. Ann Rheum Dis. 2003;62:572-5.
matory microcrystals alter the functional phenotype of human osteoblast-like cells [PMID: 12759298]

516 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 www.annals.org
Annals of Internal Medicine
Current Author Addresses: Dr. Choi: Division of Rheumatology, De- Dr. Mount: Brigham and Women’s Hospital, Renal Division, Room
partment of Medicine, University of British Columbia, Arthritis Re- 540, 4 Blackfan Circle, Boston, MA 02115.
search Centre of Canada, 895 West 10th Avenue, Vancouver, BC V5Z Dr. Reginato: Massachusetts General Hospital, 55 Fruit Street, Boston,
1L7. MA 02114.

www.annals.org 4 October 2005 Annals of Internal Medicine Volume 143 • Number 7 W-121

You might also like