+ All Categories
Home > Documents > Derrick MacFabe M.D.

Derrick MacFabe M.D.

Date post: 25-Feb-2022
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
51
Derrick MacFabe M.D. 1 b Fermentation Products of the Gut Microbiome: Bench to Bedside Implications for Autism Derrick F. MacFabe M.D. (Assist. Professor and Director) The Kilee Patchell-Evans Autism Research Group Depts. of Psychology and Psychiatry (Div. Develop. Disabilities) Schulich School of Medicine and Dentistry, Lawson Research Institute, University of Western Ontario, London, Canada ([email protected]) b KPEARG Website: http://psychology.uwo.ca/autism.htm Publications/webinars Nature of Things Autism Enigma: http://www.cbc.ca/natureofthings/episode/autism- enigma.html DISCLAIMER Research depicts studies using rodents with EEG electrodes/drug ports to measure brain activity and behaviour, with studies based on extensive biochemical and tissue culture research Similar to human patients undergoing workup for surgical treatment of epilepsies Research ethics in strict accordance with Canadian Council on Animal Care and University of Western Ontario Animal Use Committee This basic science research in no way is intended to support any treatment claims by any groups not medically sanctioned by experienced physicians practicing evidenced based medicine in autism spectrum and related disorders. NO CONFLICTS DECLARED OVERVIEW Clinical Presentation of Autism Spectrum Disorders- a whole body disorder? A Family of Disorders- Multiple Causes- Final Common Pathway Genetics of Autism Environmental Risk factors- infection Neuropathology of Autism (altered brain development and neuroinflammation) Autism as a Metabolic Disorder-A Brain in an Energy Shortage The microbiome- the “Inner Rainforest” Antibiotics and the Western Diet- Clear Cutting the “Inner Rainforest” Can infectious agents control behaviour? Are the microbes in charge? Kilee Patchell-Evans Autism Research Group-multi-disciplinary- gut link to autism Using Animal Models to Study ASD’s -repetitive/ impaired social behaviour -brain electrical activity -neuropathology -immune/metabolic/gene activation Autism as the “Diabetes of gut fatty acid metabolism” Special populations- ASD with GI symptoms/Somali expatriates Future Directions in Autism Research and Treatment Autism – A Brain Disorder of Repetitive Movement, Restricted Interests, Sensory Sensitivity and Impaired Socialization Originally 1:10,000 (1950’s) Now 1 in 90 persons (males>females) Abnormal Social Interaction- object fixation Speech and Language Difficulties Repetitive Stereotyped Movements Self-Injurious Impulsive Behavior Sensitivity to Sensory Input Savant Syndrome (rare)/restricted interests Regression/ variable course in some patients “picky eating” carb craving Comorbidities: Seizure disorder Gastrointestinal dysfunction Immune/metabol. abnormalities CNS/GI* Mitochondrial disorder/dysfunction Genetic< 5% Genetic/environmental interactions? Enlarged Brain Size Increased Neuronal Density Altered Cell Migration Seizure Disorder White Matter Disorder Glial/microglial Changes Neuroinflammation (Impaired Neurodevelopment and Cortico-cortical processing) Systemic Changes Immune System Gastrointestinal System Metabolic Disorder Detoxification Systems (glutathione) Genetic Factors Neurotransmitter Growth Factors Cell-cell Interaction Sex Linked (Fragile X) Metabolism (carnitine synthesis) Environment Metals Hydrocarbons Infectious Drug (valproate) Diet- Wheat Casein Allergy Carbohydrate? Hormonal Sex Hormones Oxytocin Vasopressin AUTISM
Transcript
Page 1: Derrick MacFabe M.D.

Derrick MacFabe M.D.

1

bFermentation Products of the Gut Microbiome:

Bench to Bedside Implications for Autism

Derrick F. MacFabe M.D. (Assist. Professor and Director)

The Kilee Patchell-Evans Autism Research Group

Depts. of Psychology and Psychiatry (Div. Develop. Disabilities)

Schulich School of Medicine and Dentistry,

Lawson Research Institute, University of Western Ontario, London, Canada ([email protected])

b

KPEARG Website:

http://psychology.uwo.ca/autism.htm

Publications/webinars

Nature of Things Autism Enigma:http://www.cbc.ca/natureofthings/episode/autism-

enigma.html

DISCLAIMER

Research depicts studies using rodents with EEG electrodes/drug ports to measure brain activity and behaviour, with studies based on extensive biochemical and tissue culture research

Similar to human patients undergoing workup for surgical treatment of epilepsies

Research ethics in strict accordance with Canadian Council on Animal Care and University of Western Ontario Animal Use Committee

This basic science research in no way is intended to support any treatment claims by any groups not medically sanctioned by experienced physicians practicing evidenced based medicine in autism spectrum and related disorders.

NO CONFLICTS DECLARED

OVERVIEW

Clinical Presentation of Autism Spectrum Disorders- a whole body disorder?

A Family of Disorders- Multiple Causes- Final Common PathwayGenetics of AutismEnvironmental Risk factors- infectionNeuropathology of Autism (altered brain development and neuroinflammation)Autism as a Metabolic Disorder-A Brain in an Energy ShortageThe microbiome- the “Inner Rainforest”Antibiotics and the Western Diet- Clear Cutting the “Inner Rainforest”Can infectious agents control behaviour? Are the microbes in charge?

Kilee Patchell-Evans Autism Research Group-multi-disciplinary- gut link to autismUsing Animal Models to Study ASD’s -repetitive/ impaired social behaviour

-brain electrical activity-neuropathology-immune/metabolic/gene activation

Autism as the “Diabetes of gut fatty acid metabolism”Special populations- ASD with GI symptoms/Somali expatriatesFuture Directions in Autism Research and Treatment

Autism – A Brain Disorder of Repetitive Movement, Restricted Interests, Sensory Sensitivity and Impaired Socialization

Originally 1:10,000 (1950’s)Now 1 in 90 persons (males>females)

Abnormal Social Interaction- object fixationSpeech and Language DifficultiesRepetitive Stereotyped MovementsSelf-Injurious Impulsive BehaviorSensitivity to Sensory InputSavant Syndrome (rare)/restricted interestsRegression/ variable course in some patients“picky eating” carb cravingComorbidities:Seizure disorder Gastrointestinal dysfunctionImmune/metabol. abnormalities CNS/GI*Mitochondrial disorder/dysfunctionGenetic< 5% Genetic/environmental interactions?

Enlarged Brain SizeIncreased Neuronal DensityAltered Cell MigrationSeizure Disorder

White Matter DisorderGlial/microglial ChangesNeuroinflammation(Impaired Neurodevelopment and Cortico-cortical processing)

Systemic ChangesImmune SystemGastrointestinal SystemMetabolic DisorderDetoxification Systems

(glutathione)

Genetic FactorsNeurotransmitterGrowth FactorsCell-cell InteractionSex Linked (Fragile X)Metabolism (carnitine synthesis)

EnvironmentMetalsHydrocarbonsInfectiousDrug (valproate)Diet- WheatCasein AllergyCarbohydrate?

HormonalSex HormonesOxytocinVasopressin

AUTISM

Page 2: Derrick MacFabe M.D.

Derrick MacFabe M.D.

2

The Genetics of Autism- Identical twin studies- 50-80% concordance - genetics and environment- Multiple Chromosomes- 2,3,7, 15, 16, 17 X- chromosome mapping-Multiple genes- brain development, neurotransmitters, language centres-Intercellular connections- Neurexins-Disorder of Gene expression (methylation, acetylation of histones)-Met Receptor Tyrosine Kinase &-Protein Kinase C beta 1 (brain, gut, immune)- at present 85-97% NO DEFINED GENETIC CAUSE-Oversimplistic to say one specific genetic cause-Epigenetics- interaction with genes/environment-“Spontaneous (?)” genetic mutation- Other genetic disorders where autism is associated

Fragile X**Angelman/Prader Willi Syndrome**Epilepsy- “tuberous sclerosis”* Rett Syndrome

Mitochondrial genetic disorder? / carnitine synthesis/absorbtion

Genetics is why you look like your father…

And if you don’t why you should!

Many environmental/infectious factors mimicGenetic transmissibility

(i.e. Tuberculosis, twins with same/different placenta)Genetic sensitivity to infection (similar pattern in ASDs!)

Environmental agents as gene “switches”

Neurodevelopment- “Lets Build a Brain”- Complex development

timing important- Many neurons die

Genetic (instruction)-EnvironmentInsults:Infection (virus)/inflammatory (IL-6)toxins (alcohol)/metals/drugs(valproate)Oxidative stress-Redox change- cell fate(germ cell-fetus-neonate)Cell to Cell Communication is Important in the organization of thedeveloping nervous system (programmed cell death and ordered cell migration)Reelin, neurexins , Gap junctions, see later……..environmental factors may alter neurodevelopment

Neuropathology of Human Autism– Neuroinflammation/mild cell lossOngoing chronic process throughout life of Patient (Pardo)

MG

MG RG

RG

NRL AUT

RG MG

Autism- A Disorder of Energy Utilization and Toxin Elimination

Oxidative Stress (Chauhan, James):Inflammation, impaired metabolismProcess similar to memory!!!Antioxidants- glutathioneFacilitators of mitochondrial function-carnitine, methylation-Methyl B12 (accessibility to CNS?)A mitochondrial disorder?(Mitochondrial DNA mutations- risk) Rossignol and Frye 2011

Page 3: Derrick MacFabe M.D.

Derrick MacFabe M.D.

3

Mitochondrial disease /dysfunction (Frye) energy dysfunction?Heterogeneous tissues affected/families/ complex inheritance

Inherited/Acquired mutations/Environmental Worsening

Autism- The Blind Men and the Elephant

Some common underlying cause involving behaviour, brain changes,GI/dietary symptoms, immunology, genetics, oxidative stress, mitochondrial disorder, environment, increase?????

ASDs

Kilee Patchell-Evans Autism Research Group 2011

Examining Animal Behaviour to Study Autism

Decreased/altered socializationfixation on objectssensitivity to sensory inputrepetitive behaviour/ seizure/dystoniaAggression , variable courseother factors normal/ improved?

Animal autism modelsPre/post natal factors

Examine brainDevelopmentElectrical ActivityNeuropathologyMetabolic markersfor subtleabnormalities

“GRAIFs” Gut Related Autism Inducing FactorsMicrobiome (100x host cells)

Bacterial metabolites- symbiosis/dysbiosis

Opportunistic Infections- key risk factori.e clostridia, yeast (chronic antibiotics)

Cell wall- LPS, beta glucan- innate immunity“priming immune system & fat metabolismFermentation products of dietary carbohydrate- Short chain fatty acids*

Barriers, variable metabolism

Acquired/genetic (met receptor tyrosine kinase)

Timing of exposure

Page 4: Derrick MacFabe M.D.

Derrick MacFabe M.D.

4

The Human Microbiome Gut Microbiome- Complex Ecosystem- Alteration with Antibiotics

Small bowel microbial overgrowth- bugs in the wrong place

Fluffy little bunnies are cute,but not in Australia!

Consequences of Overfeeding

Yangtze River Delta Pollutionone third of population China

++ fertilizer“overfeeding the ocean”

Ecosystem crashJelly fish bloom

“the clostridia of the ocean?”

Colonization of Newborn with Hospital Acquired Antibiotic resistant florae, feeding or long term antibiotic treatment- Altered Microbiome Development of Agriculture and Animal Husbandry

Page 5: Derrick MacFabe M.D.

Derrick MacFabe M.D.

5

Development of Cereal Diet, Animal Domestication & Urban CultureCo-Evolution of Endogenous Florae? (J. Diamond/M. Pollan)Cultural Taboos with cattle/dairy

“Urbanization of Cattle”- Crowding, Corn and Chronic Antibiotics(clostridia, propionibacteria)

Obese vs. lean rats- switch gut flora at birth- switch body type!Long term antibiotics in rats change gut bacteria/fat metabolism

“Its not my lifestyle, its my gut bugs!”

State of Modern day Human Microbiome Research

Digestive Tract Pathology in Autism- Lymphoid Nodular HyperplasiaImpaired Carbohydrate Digestion, Dysbiosis( Horvath, Williams)

Intestinal pathology on a subset of autistic patientsAssociated with regressive onset and GI symptomsModerate inflammatory process (nonspecific?)NOT CAUSED BY VACCINES!!!! Cause????

Psssst…Dr. MacFabeFeed Us and Spread Us Around!!

Can Enteric Bacteria Affect Brain Development/Behaviour?

Clinical- Food Craving/Symptom Worsening/ GI symptomsGut changes (gluten/casein) poorly studied (antigenic mimicry)Early gut colonizers- alteration with antibiotics (increased incidence)Unique bacterial species (clostridials, desulfovibrio, bacteriodetes)“Leaky” or malabsorbtive digestive tract (impairment of barriers)Production of bacterial metabolites (fuel for brain)Effect on Brain development, physiology, behaviour, immune function

Page 6: Derrick MacFabe M.D.

Derrick MacFabe M.D.

6

Pathogen “Control” of Host Nervous System for Propagation

Cordyceps Fungus RabiesClimbing (insects) biting (mammals)

Borna Disease

Brain specificMammals, birds

Nasal transmissionMovement disorder

Oral movements“food in mouth”

Human infection?(mood disorderSchizophrenia)

Pediatric Autoimmune Neuropsychiatric Disorder Associated with Streptococcal Infection (PANDAS)- (Swedo)

Associated with Group A Beta Strep infections”Clingy- OCD/Tick symptoms, relapsing remitting

Autoantibody to basal gangliaSimilar behaviour in family (sensitive population)

GM+, spore,Toxin A (enterotoxin),B (cytotoxin) binary?,biofilm (“hiding”)Severe- pseudomembranous

Colitis? Mild infections/carrier state?age of infectionFinegold- regressive ASDUnique bacterial populations

C. difficile

Carbohydrate Craving, Unique

OpportunisticBacteria

Diarrhea and Fecal Smearing in

AutismBehaviour facilitates growth and spread of autism implicated gut

pathogens (clostridials)?

Pathogen affecting host behaviour

Photo- M. Herbert

Clostridium Difficile – Epidemic

Genetics ARE important- Genetic mutations of infectious processes too!

Page 7: Derrick MacFabe M.D.

Derrick MacFabe M.D.

7

Autism in Somali Diaspora in North America

3% of general population, 35% of autism in some regionsAll conceived in Receiving Country- NOT Somalia

Large exposure to antibiotic/++ gastrointestinal infections

Propionic Acid- Neuroactive properties

Weak organic acid: lipid/water soluableUptake passive active (monocarboxylate receptors)G protein coupled receptorsIntracellular concentration (intracellular acidification)Unique CNS/GI immunological properties

Short Chain Fatty Acids – Propionic Acid (PPA)

Propionate:

Byproduct of bacterial metabolismClostridium, Propionibacteria (gut/acne)Desulfovibrio (Finegold)(butyrate, acetate)- short chain fatty acids

Common preservative of wheat and dairy products ( alone or Nabitor)Increased by ethanol, B12/biotin deficiency

Variable metabolism of propionate in population – Multiple mechanisms and multiple clinical presentationshares similarities with autism- underreported???

Role of diet, gut bacteria/barriers and “sickness” in propionate levels(other short chain fatty acids and metabolites)

•A Review of Propionic Acidemia:

• Part of a family of metabolic disorders (methylmalonic acidemia• propionyl CoA carboxylase, multiple carboxylase, biotinidase deficiency• Considerable polymorphisms (chromosome 3 and 13) – underreported• May be 10x more common (Asian, Mediterranean)• Elevated in other organic acidemia, biotin/B12 deficiency, alcohol• Developmental delay, seizure, movement disorder, GI disturbances.•Autism like behaviour• Acidosis/ propionate excretion may or may not be present•NB- difficult to measure PPA and metabolites

•A Review of Propionic Acidemia (cont):

•Mechanisms:

• Mitochondrial disorder leads to increased propionate/• propionyl CoA – mitochondrial toxin• Intracellular accumulation of short chain fatty acids leading to acidosis. • Increased nitric oxide, peroxide, impaired –SH, • NB-Carnitine depletion – mitochondrial uncoupling• Glutamate/dopamine 5HT release •lipoperoxidation (membrane damage)• Gene expression (Tyrosine OHase, enkephalins)• histone deacetylase inhibitor (gene expression)• “Sensitivity to metabolic stress”

Propionate leads to reduced antioxidant defences/PropCoA toxicityBiotin deficiency common in pregnancy (gut bacterial source)

Page 8: Derrick MacFabe M.D.

Derrick MacFabe M.D.

8

Rodent Model of Autism- Effects of Propionic Acid Administration(MacFabe et al, Behavioural Brain Research 2007),MacFabe, Microbial Ecology 2012 for Reviews

PPA

Autism Model –Propionic Acid (PPA)- behaviour/EEG

Effect immediate, transient (45min) but some permanent

PBS

Propionate Autism Model

-Injected into cerebral ventricles-NB buffered to pH 7.5- Reversible repetitive behaviour-Fixation on objects-Seizure +/behaviour cortex-Subcortical spiking

Automated Behavioural Monitoring (Ethovision):- Computerized long term quantification of movement- Combined with drug administration, brain electrical activity- Repetitive behaviour

Vehicle

Propionic Acid Autism Model

Total Distance Infusion Day 1 Low Dose

Time (min)

0 10 20 30 40 50 60 70 80 90 100 110 120

Dis

tanc

e To

tal (

cm)

0

200

400

600

800

1000

1200

1400

1600

PBS PPA

Intraventricular PPA- “ritual”

Hippocampal EEG- Repetitive motor loopNormal EEG

PPA induced kindling in Cortex,Hippocampus, CaudateNo effects seen with control treatments (PBS, propanol)Brain “remembers” exposure

Propionic Acid induces long convulsion- Acetate short (metabolized)

Page 9: Derrick MacFabe M.D.

Derrick MacFabe M.D.

9

Subcortical Spiking with PPA

Caudate spikes lead to limb dystoniaNormal Hipp, frontal spikingCaudate spike 500ms preceed CTXOnly PPA , no other groupsProduce Caudate eventNo kindling, immediateRole in movement disorder

encephalopathyautism??

Would be missed by traditionalSurface EEG leads in pts

0

2

4

6

8

10

Mea

n R

etro

puls

ion

Bou

ts Baseline Treatment

Retropulsion

Low Propionate

High Propionate

Propanol Sodium Acetate

PBS

*

0

3

6

9

12

15

Mea

n N

umbe

r of S

nake

Pos

ture

s

Snake Posture

Low Propionate

High Propionate

Propanol Sodium Acetate

PBS

*+

0

10

20

30

40

50

Mea

n Tu

rnin

g

Turning

Low Propionate

High Propionate

Propanol Sodium Acetate

PBS

##

Legend* = Significantly different from all control groups.# = Significantly different from low PA, propanol, PBS.+ = Significantly different from propanol and PBS

Propionic acid causes movementdisorder with caudate spiking

Caudate Spiking and Limb Dystonia Only Caused by PPA

Movement disorder effects of PPAMost sensitive

Social Behaviour (Ignoring/Mean Distance Apart)(Shultz et al. Neuropharmacology, 2008)

vehicle PPA

Effect apparent after one dose, reversible post metabolismReduced play behaviour

Social “Ignoring” of Normal Rat

PBS PPA

MacFabe et al; Behavioural Brain Research (2010)

PPA Rats Prefer “Favourite Objects” to other Rodents (MacFabe et al, 2010 BBR)

1 2 3 1 2 3Object:

40

30

20

10

0

Entri

es in

to z

one

Object Zone EntriesPPA PBS

1 2 3 1 2 3Object:

100

80

60

40

20

0

Dur

atio

n in

zon

e (s

ec)

Duration in Object Zone

1 2 3 1 2 3Object:

30

25

20

15

10

5

0

Sniff

Bou

ts

Object Sniff Bouts

A

B

C

B

C

PPAPBS

250

200

0

Dur

atio

n (s

ec)

Duration in Close Proximity

150

100

50

Proximity to rat Proximity to object

0

Perc

ent t

ime

Time Approaching Rat or Object40

30

20

10

Approaching rat Approaching object

A PPA PBS

NR NR

NO NO

Page 10: Derrick MacFabe M.D.

Derrick MacFabe M.D.

10

Long term effects- Ethovision- stereotypies/ object fixation

PBS-first trial PPA-first trial-nrl acquisition

PBS- reversal PPA-reversal- perseveration

Morris Water Maze- perseveration- Shultz et al, 2009

platform

platform

Neuropathology of PPA in Rodent Model:

Similarities to metabolic/autism spectrum disordersInnate neuroinflammation, oxidative stress, BBBAltered gene expressionAltered lipid metabolism

Hippocampal formation: GFAP (neuroplastic marker)reactive astogliosis

PBS High PPA

microinjections of propionate - very brief exposureAstrogliosis - prominent, hippocampus, cingulum, white matter Neuroinflammation (TNF alpha)Toxic or compensatory (neuroplastic response)

Results – CD68 Microglia – 14 day

Control (PBS) PPA

PPA increases activated microglia (neuroinflammation)Nitric oxide, cytokinesEndovascular involvement (microcirculation/ BBB)(c/f human autism!)

PhosphoCREB- (CAMP, Calcium- gene induction)

Control (PBS) PPA

PPA can induce multiple genes implicated in learning, memoryaddiction, neurodevelopmentEnvironment influencing genetic expression!(overexpression- Anxiety, obsession?

Page 11: Derrick MacFabe M.D.

Derrick MacFabe M.D.

11

Control (PBS) PPA

Monocarboxylate I Transporter (fatty acid uptake)

Transporter in brain/ gut endothelium- developmentImportant in weak acid transport (SCFAs, lactate) and ketonesAlterations in fatty acid transport?(critical in neurodevelopment and cell signaling)

Is Propionate Cytotoxic? - Activated Caspase 3’(apoptosis)

cont High PA

PPA is not grossly neurotoxic in hippocampusmay be a neuroprotectant!- histone deacetylase inhibitionNeuroinflammation with little neurotoxicity

PPA Digital analysis

Interleukin 6PBS PPA

WM

Hipp

PBS Vehicle High Dose PPA

PPA causes increase anti Nitro-tyrosine immunoreactivity in hippocampal formation increases “oxidative stress”

Anti Nitrotyrosine Immunoreactivity- oxidative stress

Oxidative Stress Markers

μM

of M

DA&

HN

E/m

g pr

otei

n

0.0

0.5

1.0

1.5

2.0

2.5

0.0

0.5

1.0

1.5

2.0

2.5

PBSPPAHigh

nM

car

bony

l der

ivat

ives

/mg

prot

ein

*

**

lipid peroxidation marker protein oxidation marker

Glutathione System

Spec

ific

Act

ivity

nmol

/min

/mg

of p

rote

in

0

20

40

80

100

μM o

f tot

al G

SH

/mg

of p

rote

in

2

4

6

8

10

12

PBSPPAHigh

**

GPx GST GR GSH

***

Increased Oxidative Stress in PPA Autism Model(MacFabe et al Am.J. Biochem.Biotech.2008)

PPA increases oxidative stress markers and impairs Glutathione metabolism (sequestration?)-brain “sensitive” to broad spectrum of environmental agents(ie metals, xenobiotics)-similarity to evidence of metabolic dysfunction in ASD patients-broad effects- metabolic encephalopathy

O

OC

HO

CH3

HO

OH

OH

O

OC

OCO

H3C

CH3

O

O CH3

OCHO

O CH3

CO OH

OHOH3C

Cl OHH3C

OC OH

O

OH

OH

OC

OH3C

CHO

O

O CH3

Fatty acids (FA) in Autism

• Building block of lipids (low FFA)• Characteristic COOH functional group• Saturated (animals) and unsaturated ( plants)

(mono vs poly), (trans vs cis)• Odd or even numbered (C2 - C26)• Modified (hydroxylated, methylated,

dicarboxylated, etc)

Page 12: Derrick MacFabe M.D.

Derrick MacFabe M.D.

12

Where do fatty acids (lipids) come from?

• Body synthesis

• Diet:– Essential and non essential fatty

acids– Products of enteric bacteria

• Essential fatty acids– Body cannot make– omega 3’ and omega 6– Balance between ω3 and ω6

affect functions

Omega 3 (ω3) FA

Omega 6 (ω6) FA

Functions of Fatty Acids• Main functions:

– Energy storage– Structural components of cell

membranes– Act as signal molecules in many

metabolic processes– Abnormal fatty acid composition in

Autism– Relative carnitine deficiency– Mitochondrial disorder?– (acquired?)

e.g. neuronal cell membrane

PE

Satu

rate

s

Mon

o-un

sat

Poly

-uns

at

Om

ega-

6:3

Plas

mal

ogen

s

Om

ega

3

Om

ega

6

Perc

ent (

%)

0

102030405060708090

100

PBS BUT PPA

Compound classes

I.E PhosphatidylethanolamineIncrease saturates

Decrease:monosaturates

omega 6/3Plasmologens (antioxidant)

Same trend in

PhosphatidylcholinePhospatidylserine/inositol

Sphingomyelin (White matter)Cardiolipin (mitochondria)

SCFA alter membrane fluidity, Signallng, Antioxidant,mitochondrial function

Thomas et al, J.Neurochem, 2010 Carnitine and Acylcarnitines• Quaternary ammonium

compound

• Synthesize from lysine and methionine (methylation step!)

• Essential for fatty acid transport from the cytosol into the mitochondria matrix for generation of energy (ATP)

• Important in brain and gut bioenergetics

Long term antibiotics (beta lactams)- deplete carnitine transportInherited defects in carnitine synthesis(TMLEHE) and transport (OCTN2)

Long and short chain acylcarnitines

Long chain acylcarnitines

TreatmentsPBS BUT PPA

nmol

/mg

brai

n tis

sue

0.00

0.05

0.10

0.15

0.20

0.25

0.30

a

b b

Short chain acylcarnitines

TreatmentsPBS BUT PPA

(nm

ol/m

g br

ain

tissu

e)

0

5

10

15

20

25

30

a

b

b

PPA causes increase in CNS Acylcarnitines- same as ASD patientsBinds to carnitine and CoEnzyme A- mitochondrial dysfunction

Ratio of bound to free carnitineR a tio o f f re e to b o u n d c a rn it in e

T re a tm e n tsP B S B U T P P A

Rat

io

0

1 0 0

2 0 0

3 0 0

4 0 0

5 0 0

a

bb

PPA causes increase in CNS bound carnitine- as in ASD patients

Bound to free carnitine

Page 13: Derrick MacFabe M.D.

Derrick MacFabe M.D.

13

PPA Rodent Model ASD Patients217 patients-17%Similar short and long chain acylcarnitinesAlso decreased glutathione, Redox changes NO genetic cause!

Frye , Melnyk & MacFabe 2012, Trans. Psych. 2013Metabolic markers in ASD Patients consistent with PPA

“Backing Up” TCA cycle as Rodent Model- Acquired Mito Disorder

CarnitineSequestration

By PPA

Common Infections, Chronic Antibiotics, Clostridia and CarnitineCollapse Leads to Constipation, Carbohydrate Malabsorbtion,

Convulsions and Compulsions!

Carnitine- Shuttle for mitochondrial fatty acid beta oxidationRoutine pre- peri or post natal infections-

Long term antibiotics (beta lactams)- deplete carnitine transport“Barren Gut”->Growth of clostridials- increased SCFA production

Further sequestration of carnitineImpaired fatty acid metabolism- mitochondrial encephalopathy

Carnitine

CAUSES CONSEQUENCES OF SCFAs

Long term antibiotics for routine infection (maternal /infant) Treatment of maternal β hemolytic strep

Gut dysmotility/inflammation/ carbohydrate malabsorbtion/altered gut permeability (tight junction impairment)

Hospitalisation (colonization of nosocomial bacteria) i.e. C-section, neonatal distress

Active uptake of SCFA to CNS (monocarboxylate transporters)

Prenatal drugs (valproate, ethanol)

pH dependent intracellular concentration of SCFA

Opportunistic infection (Clostridium spp., Desulfovibrio spp.)

Neurotransmitter synthesis and release (catecholamines, enkephalins) CNS/sympathetic nervous system

Maternal/Infant gut dysbiosis

Receptor activity (+NMDA, -GABA) SCFA G protein coupled receptors/Ca++ influx

Organic acidemias (propionic/methylmalonic, biotinidase/ holocarboxylase deficiency)

Gap junction closure, altered neurodevelopment, neuroinflammation

(B12/biotin deficiency)

Impaired mitochondrial function/ increased oxidative stress

Genetic/acquired impaired carnitine synthesis/absorption (TMLHE/OCTN2 genes, β- lactam antibiotics)

Reduced glutathione/increased sensitivity to xenobiotics (i.e. acetaminophen)

Mitochondrial disorder/dysfunction (inherited, acquired)

Decreased carnitine/altered lipid metabolism/membrane fluidity

Colitis (impaired barrier/SCFA metabolism), i.e. celiac disease, Met-receptor tyrosine kinase mutation

Altered gene expression (CREB activation, histone deacetylase inhibition)

Increased refined carbohydrate consumption – substrate for bacterial fermentation

Antisocial/perseverative/anxiety-like behavior, seizure/movement disorder, Restrictive food interests/carbohydrate craving

Causes Consequences of SCFAs

Diabetes Autism

Type 1 Type II

Can’t metabolize glucose Can’t metabolize SCFAs?Multi- system involvement

Multiple Causes (Genes/diet/environment)Present with Metabolic Crisis (i.e infection)

Treatment-Carbohydrate restriction (direct/indirect)Treatment-Insulin/glyburide Carnitine/bacterial

eradication/probiotics/MB12? Multi- system approach (metabolite measurement diagnostics)

Baseline T1 T2 T3 T4 T5 T6 T7 T8 --0

500

1000

1500

2000

2500

3000

3500

4000

4500

Mea

n To

tal D

ista

nce

Mov

ed (c

m)

Treatment Days

PBS PPA Buturate

TOTAL DISTANCE

* *

^

+

**

***

^^

^^

^^

*

Intraventricular infusion PPA and Butyric Acid induce increased locomotor activity in rats

(gut SCFA metabolites can alter behavior)

Butyric AcidPPA

Page 14: Derrick MacFabe M.D.

Derrick MacFabe M.D.

14

Average Startle

Trial type115 dB 73 dB 76 dB 82 dB no pulse

Ave

rage

sta

rtle

(mV

)

0

20

40

60

80

100PBS n = 30PPA n = 31

* **

Adolescent Behavioural Changes in Response to Early Exposure to PPA:Acoustic Startle

During first few days of life, rat pups are injected with sub cut PPA or PBS andBehaviourally tested as adolescents.

The amount that an animal is startled (“jumps”) in response to an acoustic stimulus is measured. PPA animals are more sensitive to stimuli – jump more – than PBS animals.- Reduced inhibition (i.e GABAergic dysfunction).

SCFA activate the transcription of TH gene

PPA and valproate induce tyrosine hydroxylase expression in PC12 cells(valproate modified SCFA- autism risk factor)

Via a CREB dependant mechanism (same in brain homogenate)Epigenetic control of catecholamine synthesis(Dopamine- movement, aggression, addiction)

Time-of-flight secondary ion mass spectrometry (ToF-SIMS)

Multiple analysis of metabolic intermediates/metals in CNSOxidative stress- nitrosylation,

altered lipid profiles (decreased cholesterol) and edema (Na and Cl) in White Matter- consistent with ASD

Gap Junctions (GJs)-intercellular pores

GJs in Central Nervous System

Neural electrical coupling

Glial couplingNeurodevelopment-cell migrationEEGGlial spatial bufferingBlood brain BarrierGating by pH, Dopamine, NO,cytokinesmetals, hydrocarbons, PPA!

Gap Junctional Coupling in Neural Systems:

» Cells connected- functional coupling

GJ

Page 15: Derrick MacFabe M.D.

Derrick MacFabe M.D.

15

Sooo…Is our PPA rodent model like human autism?• Hyperactivity/OCD Movement/ Object fixation/perseveration• Intermittent seizure/ Subcortical spiking with movement • Kindling/ Neuroplasticity/neonate-adult• Social Impairment• No gross neurotoxicity• Astrogliosis/microglia/ Neuroinflammation• White matter damage (lipoperoxidation, edema, altered membrane fluidity)• Oxidative stress/ impaired glutathione (broad spectrum detoxifier)• Same lipid/acylcarnitine profiles as ASD patients• induce catecholamine/CREB expression- HDAI (epigenetics)

•Propionate is known to cause:• Neutrophil/monocyte migration (specific SCFA receptors)• Mitochondrial uncoupling (fatty acids), increases in odd chain FAs, low chol.• Neuronal structural changes (cytoskeleton)/gene expression)• Intracellular acidification - Dopamine/glutamate/5HT release – gene induction• Impairment in cell-cell signal transduction (gap junctions, cytokines)

PPA

Active uptake to CNS

Repetitive/antisocial behaviour/Seizure

SCFA G protein receptorsNeurotransmitter Synthesis and release

Increased intracellular Calcium

Neuroinflammation/neurodevelopment

Gut motility and inflammationMalabsorbtion

Mitochondrial function/oxidative stressAltered lipid/membrane metabolism

Altered Gene Expression

Short Chain Fatty Acid BacterialFermentation Products

Gap Junction Closure

Evolutionary Psychiatry- population vs individual

Some direct/ indirect advantage to behavioural trait

Ziggy Freud Chucky Darwin

Are the Microbes in Charge?

“Where have we come from, where are we now, where are we going?”Gaugin

Development of Agriculture and Animal HusbandryCivilization/Culture

Modern Urbanization/Medicine/AntibioticsHuman Migration/Factory Farming

Emergent Diseases of Industrialization

Cautious Optimism- need for further rational study!Short Chain Fatty Acids not “Good” or “Bad”

Timing/Amount/Genetic Sensitivity Critical A MODEL IS NOT IDENTICAL TO THE HUMAN CONDITION

Page 16: Derrick MacFabe M.D.

Derrick MacFabe M.D.

16

A MODEL IS NOT IDENTICAL TO THE HUMAN CONDITION

Somali families Autism or no autism (small vs large groups)History

Stool- bacterial culture- isolation, animal models, screeningGenetics metabolics

Compare with other Somali populations (US, Sweden), Saudi (urban/rural) and general population

Assistance ASDCARC, U. of Guelph, King Saud and Karolinska

Bacterial Ecology of Regressive Autism/ “Repoopulate”

Microbial ecology ofregressive autism

Reconstitution of “scaffolding”of key gut

microbial communitiesC. diff/ulcerative colitis

Tore Midtvedt- Karolinska- Grandfather of the Microbiomefollowing babies for health or disease- germ free animals

Nobel Symposium-Autism and Gut-Karolinska(Stockholm)-Midvedt,Nicholson, Forssberg, Reichelt, Bolte, Bienenstock, Falk-Ytter

Summary

Autism is a complex problem needing a multi-disciplinary approach

with modern brain research and microbiology techniques, much is available to rationally examine autism as a defined brain disorder

detailed study of behaviour in animal models necessary to determine cause and develop rational clinical studies and treatment

Factors in brain development- neural migration, embryonic cell deathtoxic environmental compounds (dietary and enteric fatty acids)role of diet, metabolism and gut microbiome (antibiotic exposure)

Gut metabolites can alter brain electrical activity, behaviour, Pathology, gene induction and cellular metabolism (mitochondria)Bacteria modulating host behavior! Western environment changing!!Variable exposure/breakdown in humans/microflorae

Page 17: Derrick MacFabe M.D.

Derrick MacFabe M.D.

17

Acknowledgements- Kilee Patchell- Evans Autism Research Group

Klaus-Peter Ossenkopp, Ph.D., Donald Peter Cain, Ph.D., Martin Kavaliers, Ph.D., (Department of Psychology (Neuroscience)- University of Western Ontario)

Fred Possmayer PhD. (Biochemistry, Obstetrics/Gynecology)Lisa Tichenoff, Kelly Foley MSc., Roy Taylor, Francis Boon, Soelaha Shams ,Melissa

Meeking, Andrew Franklin MSc., Jennifer Hoffman MSc., Jennifer Mepham , Raymond Thomas, PhD., Karina Rodriguez-Capote, PhD, Yalda Mohammad-Asef,

Jessica Benzaquin, Sandy Shultz MSc., Stacey Holebrook

Heng Yong-Yi Ph.D and Leo Lau Ph.D (Surface Science Western- The University of Western Ontario)

Jeanette Holden, Ph.D (Queens University)

Martha Herbert, M.D. Ph.D, (Harvard University)

Edmund F. La Gamma, M.D. , Bistra Nankova, Ph.D(Westchester Medical Center - New York Medical College)

Sidney Finegold, M.D. (UCLA) and Emma Allen Vercoe (U. of Guelph)

Funding:GoodLife Children’s Charities

Brickenden FoundationAutism Research Institute

Natural Sciences and Engineering Research Council of Canada

AUTISM- HOPE FOR THE FUTURE

Active uptake to CNS

Repetitive/antisocial behaviour/Seizure

SCFA G protein receptorsNeurotransmitter Synthesis and release

Increased intracellular Calcium

Neuroinflammation/neurodevelopment

Gut motility and inflammationMalabsorbtion

Mitochondrial function/oxidative stressAltered lipid/membrane metabolism

Altered Gene Expression

Short Chain Fatty Acid BacterialFermentation Products

Gap Junction Closure

Page 18: Derrick MacFabe M.D.

Short-chain fatty acid fermentationproducts of the gut microbiome:implications in autism spectrumdisorders

Derrick F. MacFabe, MD*

Director: The Kilee Patchell-Evans Autism Research Group, Departments of Psychology (Neuroscience) andPsychiatry, Division of Developmental Disabilities, Lawson Research Institute, University of Western Ontario,London, ON, Canada, N6A 5C2

Recent evidence suggests potential, but unproven, links between dietary, metabolic, infective, and gastro-

intestinal factors and the behavioral exacerbations and remissions of autism spectrum disorders (ASDs).

Propionic acid (PPA) and its related short-chain fatty acids (SCFAs) are fermentation products of ASD-

associated bacteria (Clostridia, Bacteriodetes, Desulfovibrio). SCFAs represent a group of compounds derived

from the host microbiome that are plausibly linked to ASDs and can induce widespread effects on gut, brain,

and behavior. Intraventricular administration of PPA and SCFAs in rats induces abnormal motor movements,

repetitive interests, electrographic changes, cognitive deficits, perseveration, and impaired social interactions.

The brain tissue of PPA-treated rats shows a number of ASD-linked neurochemical changes, including innate

neuroinflammation, increased oxidative stress, glutathione depletion, and altered phospholipid/acylcarnitine

profiles. These directly or indirectly contribute to acquired mitochondrial dysfunction via impairment in

carnitine-dependent pathways, consistent with findings in patients with ASDs. Of note, common antibiotics

may impair carnitine-dependent processes by altering gut flora favoring PPA-producing bacteria and by

directly inhibiting carnitine transport across the gut. Human populations that are partial metabolizers of PPA

are more common than previously thought. PPA has further bioactive effects on neurotransmitter systems,

intracellular acidification/calcium release, fatty acid metabolism, gap junction gating, immune function,

and alteration of gene expression that warrant further exploration. These findings are consistent with the

symptoms and proposed underlying mechanisms of ASDs and support the use of PPA infusions in rats

as a valid animal model of the condition. Collectively, this offers further support that gut-derived factors,

such as dietary or enteric bacterially produced SCFAs, may be plausible environmental agents that can trigger

ASDs or ASD-related behaviors and deserve further exploration in basic science, agriculture, and clinical

medicine.

Keywords: autism; mitochondria; Clostridia; Desulfovibrio; propionic acid; butyric acid; carnitine; neuroinflammation;

oxidative stress; glutathione; gap junctions; microbiome; PUFA; epigenetics

Autism spectrum disorders (ASDs) are a family of

neurodevelopmental disorders of rapidly increas-

ing incidence (1) that are characterized by

impairments in communication and social interaction

along with restrictive and repetitive behaviors.

The brain tissue of patients with autism shows subtle

developmental abnormalities, specifically in those areas

concerned with language, facial expression, movement,

and social behavior (2). Individuals with autism may

show enlarged brain size in the first few years of life, with

altered migration of cortical, amygdalar, and cranial

nerve motor neurons, as well as cerebellar neurons (3).

Cell counts have shown that compared with controls,

brain samples of patients with autism contain smaller

neurons with increased cell density in cortical, limbic, and

cerebellar regions (4), possibly because of altered neuro-

genesis (5), apoptosis (6), neural cytoarchitecture (7, 8),

or a combination of these factors. There is a growing

interest in examining ASDs as a disorder of glial cell

function. Glial cells are of great importance in both the

developing and mature nervous system, particularly with

respect to cell�cell interactions during neural migration,

(page number not for citation purpose)

�THEMATIC CLUSTER: FOCUS ON AUTISM SPECTRUM DISORDERS

Microbial Ecology in Health & Disease 2012. # 2012 Derrick F. MacFabe. This is an Open Access article distributed under the terms of the Creative Commons Attribution-

Noncommercial 3.0 Unported License (http://creativecommons.org/licenses/by-nc/3.0/), permitting all non-commercial use, distribution, and reproduction in any medium,

provided the original work is properly cited.

1

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 19: Derrick MacFabe M.D.

and synaptic plasticity (9). Glia form a functional

syncytium necessary for the maintenance of a stable

neural microenvironment, especially during periods of

increased metabolic stress (10). Glial abnormalities may

manifest as an overall increase in white matter thickness

(11), increased thickness of the external capsule and

increased water content in the white matter (12). These

observed abnormalities in brain morphology are accom-

panied by increased CNS immune activity, including

increases in reactive astrocytes and activated microglia in

brains of patients with ASD, as well as the elevation of

proinflammatory cytokines in the cerebral spinal fluid.

These findings have been demonstrated in both young

and older patients, suggesting that an inflammatory pro-

cess may be present throughout the life span of indivi-

duals with autism (13). Similar findings of heightened

immune activity (i.e. increased Th2 cytokine levels) have

also been demonstrated in peripheral blood monocytes

from patients with ASDs (14).

To date, the majority of research has focused on genetic

causes of ASD, mainly on developmental abnormalities

in synaptic organization, neuromigration, and neuro-

transmission (15). However, the findings that genetic

syndromes appear to account for only 6�15% of ASD

cases, coupled with the observation of discordance

of severity among monozygotic twins (16), indicate a

solely genetic cause to be unlikely. An alternative ap-

proach is to examine ASDs as a whole body condition,

with many comorbidities involving immune, metabolic,

and gastrointestinal abnormalities, that may have a

variable symptomatic course (17�22). Recent studies

show widespread dysfunction in immune regulation,

detoxification, environmental exposures, redox regula-

tion/oxidative stress, and energy generation (19�22) that

affect many organ systems, including the brain. Further-

more, the recent findings of genetic mutations, such as

the MET receptor tyrosine kinase and protein kinase

Cb genes, involved in brain development, immune func-

tion, and ability to recover from gastrointestinal

insults (23, 24), and the X-linked 6-N-trimethyllysine

dioxygenase (TMLHE) gene, involved in carnitine/fatty

acid metabolism (25), raise the strong possibility that

genetic sensitivities to environmental factors may under-

lie the pathophysiology of the disorder in at least a subset

of patients.

Is autism an acquired disorder of mitochondrialdysfunction?Disorders of mitochondrial function, and their hetero-

geneous expression in different tissues or within families,

fulfill most of the criteria for widespread multiorgan

dysfunction and sensitivity to many diseases (26, 27).

Classic mitochondrial disease is overrepresented in pa-

tients with ASDs, encompassing 5% of the ASD popula-

tion (28). More intriguingly, an extensive meta-analysis

by Rossignol and Frye also found that about 30% of

children in the general ASD population exhibit biomar-

kers consistent with mitochondrial disease (22), including

a relative carnitine deficiency and altered lactate/pyruvate

ratios. Furthermore, a recent study reported that 80%

of the children with ASDs showed altered electron trans-

port chain in lymphocytes compared with neurotypic

controls (29). Mitochondria are central to this theme

as polymorphisms in mitochondrial genes, which can

be inherited or acquired from early prenatal environ-

mental insults, can result in altered susceptibility to many

diseases (26, 27). Similarly, abnormalities associated with

ASDs such as glutathione deficiency (30), increased

oxidative stress, elevated concentrations of TNF-a(31�33), and methylation abnormalities could further

impair mitochondrial function. The fact that classic

mitochondrial disorder is overrepresented but does not

account for the high occurrence of mitochondrial bio-

marker alterations in this population may indicate that

mitochondrial dysfunction in ASDs may be at least partly

environmentally acquired. This is important in light of a

rapidly growing incidence of ASDs (1) and the growing

consensus that the systemic abnormalities seen in ASDs

may arise from environmental triggers (34) in genetically

sensitive subpopulations (35, 36).

Mitochondrial dysfunction can result from a broad

assortment of environmental exposures that have been

implicated in the development of ASDs. Possible agents,

which are not mutually exclusive, include heavy metals

(37�40), chemicals (41), polychlorinated biphenyls (42),

pesticides (43�45), and finally enteric metabolic products

of ASD-associated intestinal bacteria (46�53).

The ‘inner’ outer environment, the entericmicrobiome as a source of environmentaltriggers of ASDsThe human digestive tract is host to a complex array of

intestinal bacterial florae, coined the microbiome, which

outnumber host cells at least 10 to 1. The microbiome

produces an array of bioactive metabolic products

capable of entering systemic circulation. It is important

to note that the enteric microbiome and its metabolic

products are not static and can be altered throughout the

life cycle of the individual, particularly throughout the

first 18 months of life (54). The metabolic products from

the gut microbiome can have profound and dynamic

effects on host metabolism, immune function, and gene

expression in many organ systems, including the CNS

(46, 55�60). In addition, it is also important to consider

the recent impact of the alteration of the human

microbiome and its metabolites through the introduction

of the high calorie Western diet, coupled with the high

exposure of antibiotics and disinfectants to human

beings, animals, and plants, as a possible source of

environmental triggers of many diseases of increasing

Derrick F. MacFabe

2(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 20: Derrick MacFabe M.D.

incidence (59), including ASDs. This is particularly

evident from human populations migrating to Western

societies, such as the Somali diaspora, who appear to

have a much higher incidence of ASDs than in their

country of origin (61).

Given the high number of reports of antibiotic

exposure, hospitalization, and gastrointestinal distur-

bances (62�64) in many patients with ASDs, our labora-

tory has been examining the neurobiological effects of

microbiota-produced short-chain fatty acids (SCFAs),

such as propionic acid (PPA) (46�53). PPA is a fermenta-

tion product of many bacteria, including some enteric

species (i.e. Clostridia, Desulfovibrio, and Bacteroidetes)

overexpressed in stool samples from patients with ASDs

(65, 66). PPA levels are also elevated in the stool samples

of patients with ASDs (67). This has led us to propose

PPA as a possible environmental trigger of ASDs. We

have found that when administered to rodents, brief

intracerebroventricular (ICV) infusions of PPA and its

related enteric SCFAs produce many behavioral, electro-

graphic, neuropathological, and biochemical changes,

consistent with an animal model of ASDs.

Neurobiological effects of enteric SCFAsPPA is found in the gut, along with other SCFAs, such as

acetate and butyrate, each of which are major metabolic

products of enteric bacteria, following fermentation of

dietary carbohydrates and some amino acids (58, 68, 69).

PPA is also produced by Propionibacteria present on

the skin, known to cause acne (70), and many bacteria

present in the oral mucosa, responsible for gingival

inflammation (71, 72). PPA is well known in animal

husbandry, where it is the main metabolic product in

ruminant cellulose digestion (73), may be a sex pher-

omone (74), and is involved in lactation (75). It is also

naturally present in a variety of foodstuffs (i.e. cheese)

(73) and is commonly used as a preservative (antifungal)

in many processed foods, particularly in refined wheat

and dairy products (76). However, the majority of PPA is

produced in the gut lumen by intestinal bacteria. PPA,

being a weak organic acid, exists in ionized and non-

ionized forms at physiological pH, thus allowing it to

readily cross the gut�blood barrier, and is principally

metabolized in liver. It also crosses the blood�brain

barrier and enters the CNS (77). In addition, PPA gains

access to the CNS via monocarboxylate transporter

uptake in the gut lumen and cerebrovascular endothe-

lium, which actively transport many carboxylic acids,

particularly PPA and ketones. PPA is also a specific

ligand of many G-coupled SCFA receptors (GPR41, 43)

(78�80). PPA and other SCFAs are taken up by glia and,

to a lesser extent, neurons once they enter the CNS (81,

82), where they are thought to comprise a major energy

source in cellular metabolism, particularly during early

brain development (81�83). PPA and other SCFAs

(i.e. butyrate) affect diverse physiological processes such

as cell signaling (84), neurotransmitter synthesis and

release (85), free radical production, mitochondrial

function, (86) lipid metabolism (87), immune function

(88), gap junction gating, intracellular pH maintenance

(89), and modulation of gene expression through phos-

phorylation and histone acetylation (90).

Although PPA may be beneficial at appropriate levels,

such as improving insulin sensitivity, lowering choles-

terol, and reducing food intake (69), excessive PPA may

have many negative effects on health and behavior. For

example, a number of inherited and acquired conditions,

such as propionic/methylmalonic acidemia, biotinidase/

holocarboxylase deficiency, ethanol/valproate exposure,

and, as discussed, mitochondrial disorders, are all known

to result from elevations of PPA and other SCFAs, partly

through the formation of propionyl coenzyme A (CoA)

and sequestration of carnitine (20, 46, 51, 91). Collec-

tively, these conditions present at varying ages with

developmental delay and regression, seizure/movement

disorder, metabolic acidosis, and gastrointestinal symp-

toms, which also change in severity with fluctuating PPA

levels and markers of mitochondrial dysfunction that are

somewhat reminiscent of ASDs (86, 92, 93).

Propionic acidemia is caused by deficient activity of

either one of two non-identical subunits of the biotin-

dependent enzyme propionyl CoA carboxylase (94). This

mitochondrial enzyme is responsible for the breakdown

of PPA and other SCFAs, as well as a number of amino

acids. The disorder may be 10 times more common than

previously reported, and there are multiple mutations and

variable metabolizers in many populations (95, 96).

Patients with propionic acidemia clinically present with

life-threatening illness during the neonatal period, char-

acterized by vomiting, severe metabolic acidosis, and

hyperammonemia. Neurological symptoms include devel-

opmental delay, seizure, choreathetoid movements, and

dystonia. Interestingly, some other patients, including

identical twins of more severely affected siblings, may

present later in life with varying severities of the disorder,

often without measurable increases in blood or urine PPA

or metabolites. Treatment includes reduction of carbohy-

drate and protein contents in the diet, eradication of

PPA-producing bacteria, and carnitine supplementation

to improve PPA clearance, which have some benefit

(46, 86, 92, 93).

Thus, PPA and its related SCFAs have broad effects

on cellular systems, providing a potential mechanism

where metabolic end products of the enteric microbiome

can alter host physiology and behavior (46, 60). This

and other evidence led us to propose that increased

PPA exposure at key neurodevelopmental periods is a

major environmental trigger of the brain and behavioral

changes observed in ASDs.

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 3(page number not for citation purpose)

Page 21: Derrick MacFabe M.D.

How are dietary and gastrointestinal factorsrelated to autism? � microbiome production ofenteric fatty acids as environmental triggersThere is a growing interest suggesting a link between

dietary and/or gastrointestinal factors and the worsening

and, in some cases, improvement of ASD symptoms, but

the mechanisms responsible for this remain elusive. As

indicated, SCFAs, such as PPA, are produced by many

gut bacteria by the breakdown of dietary carbohydrates

and amino acids, particularly from wheat products (97).

Of particular relevance are the Clostridia and Desul-

fovibrio, which have been proposed as infectious causes

of ASDs (64, 66). Clostridial species, a family of hetero-

geneous anaerobic, spore-forming, gram-positive rods

(98), are major gut colonizers in early life and are

producers of PPA.

Clostridia as spore formers are particularly resistant

to most antibiotics used for routine perinatal and early

childhood infections, and some species are a cause of

major hospital- and, recently, community-acquired com-

municable disease (i.e. C. difficile-induced colitis) (98).

Interestingly, spore-forming anaerobes and microerophilic

bacteria, particularly from clostridial species, have been

shown to be elevated in patients with ASDs (66, 99, 100).

Recently, Finegold isolated distinct species of Desul-

fovibrio, a gram-negative, aerotolerant, non-spore former,

from the stool of patients with ASDs, and, to a lesser

extent, non-affected siblings. Interestingly, in addition to

producing PPA following fermentation of peptones,

Desulfovibrio is resistant to most common antibiotics

and produces the gasotransmitter and potential mito-

chondrial toxin, hydrogen sulfide (101�103). He has

suggested eradication of these organisms with oritavancin

and aztreonam as a possible treatment of ASDs (64).

Furthermore, ASDs may show comorbidity with a variety

of gastrointestinal disorders, such as alterations in gut

motility, intestinal lesions and increased intestinal perme-

ability, bacterial dysbiosis, impaired carbohydrate diges-

tion/absorption, reflux esophagitis, and ileal hyperplasia

(46, 104, 105). An association between long-term anti-

biotic use, hospitalization, abdominal discomfort and the

onset of ASD symptoms after normal or near-normal

development has also been reported (62, 63, 100, 106).

These findings raise the possibility that gut-born factors

secondary to alteration of the gut microbiome by anti-

biotics or diet may influence brain function and sympto-

mology in patients with ASDs. Moreover, a compromised

gut�blood barrier (i.e. an acquired colitis) or impaired

colonocyte energy metabolism (102), which use SCFAs as

an energy substrate and act as a metabolic ‘sink’, may

allow for greater systemic and CNS access for such

compounds (59, 107).

PPA is also known to have a number of direct effects

on gastrointestinal physiology. As reviewed in the study

by MacFabe et al. (46), PPA increases contraction of

colonic smooth muscle, dilates colonic arteries, activates

mast cells, increases the release of serotonin from gut

enterochromaffin cells, and reduces gastric motility and

increases the frequency of contractions, reminiscent of

the clinical observations of gut dysmotility in patients

with ASD. PPA levels increase in infant colonic contents

postnatally and following formula feeding opposed to

breast feeding (54). Interestingly, intracolonic infusions

of PPA, mimicking bacterial overgrowth, can induce an

experimental colitis in infant rats but not later in life

(108). However, ascertaining PPA production in vivo is

difficult, as secondary to both passive and active diffu-

sion of SCFAs from the gut lumen, and intracellular

concentration levels in stool are often a poor indicator

of SCFA production by gut bacteria and may change

rapidly with diet (58).

Taken together, PPA appears to circumstantially

possess the necessary properties to interfere with gastro-

intestinal activity in a manner similar to the abnormal-

ities observed in ASDs. In support of this, reports from

parents of children with ASDs suggest that behavioral

and gastrointestinal symptoms increase when their

children ingest refined food products that provide high

carbohydrates for bacterial fermentation to produce PPA

or also contain PPA as a preservative. Furthermore,

behaviors and gut symptoms improve following the

elimination of these products from the diet (17, 106) or

the eradication of PPA-producing bacteria by broad-

spectrum antibiotics (109). Moreover, symptom exacer-

bation associated with propionic acidemia and its related

conditions bears some resemblance to that reported in

ASDs (86, 92, 93). Stool (67) and serum studies from

patients with ASDs provide further evidence linking PPA

to the condition, as patients with ASDs have metabolic

dysfunction, including impairments in B12, glutathione,

or carnitine metabolism (110, 111) and mitochondrial

disorder/dysfunction (20), which are consistent with the

effects of PPA on cellular metabolism (46, 51, 53).

The PPA rat model of ASDs � behavioral andelectrographic findingsAnimal models allow for the examination of factors

involved in human disorders, such as ASDs, using

experiments that cannot be conducted in humans. The

development of such models is essential and permits the

experimental examination of the effects of suspected

environmental agents on the pathophysiology and core

symptoms and comorbidities of ASDs. There have been a

number of valid animal models of ASDs, concentrating on

genetic knockouts of key neurodevelopmental processes

(112, 113) and also on exposures to environmental factors

such as metals, drugs (valproic acid, terbutaline) (114,

115), viruses (116) and inflammatory agents (i.e. lipopo-

lysaccharide, interleukin-6) (117). Each of these models

causes some, but not all, behavioral or pathological

Derrick F. MacFabe

4(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 22: Derrick MacFabe M.D.

changes reminiscent of ASDs. Therefore, to investigate the

hypothesis that that elevated levels of PPA can induce

bouts of behavioral, electrophysiological, neuropatholo-

gical, and biochemical effects similar to those observed in

ASDs, our initial research was concentrated on exposure

of adult rats to brief pulsed ICV infusions of PPA and its

related enteric SCFAs (i.e. acetate, butyrate), through

chronic indwelling brain cannula. CNS exposures were

initially done to test for a central, and not peripheral,

action of these compounds, as it has been argued that the

behaviors, dietary/gut symptoms, and metabolic findings

in ASDs are solely due to responses to gastrointestinal

pain, poor diet, or obsessional eating behavior and not a

direct effect of enteric compounds on the function of CNS.

Impairments in social behavior, including abnormal

play behaviors and other forms of social contact, are

among the most prominent symptoms of ASDs (49, 118).

Likewise, individuals with ASDs commonly suffer from

various forms and degrees of cognitive impairment,

including learning disabilities, restricted interests favoring

objects vs. social interactions, and ‘rigid’ perseverative

behavior, and insistence on ‘sameness’, and rituals

(18, 119). Patients with ASDs often experience motor-

related symptoms, such as hyperactivity, gait disturbances,

and stereotyped movements, possibly grouping ASDs with

the movement disorders (120, 121). There is increased

incidence of seizure disorder in ASDs and other condi-

tions associated with elevated levels of SCFAs (92, 122,

123). We thus tested the validity of the PPA model based

on the above behavioral and electrographic criteria.

Pulsed ICV infusions of PPA (4 ml of 0.052�0.26 M

solution at pH 7.5 over a 1 min period) or control

compounds (i.e. isomolar, acetate, butryate and propanol -

the non-acidic alcohol analogue of PPA) were performed

over a number of time courses (once weekly�5 weeks,

once a day�5 day, twice daily for 7�14 days) into the

cerebrospinal fluid of adult rats via chronic brain

indwelling canullae (Fig. 1).

PPA-infused rats showed bouts of increased repetitive

locomotor activity, turning, retropulsion, tics, social

impairment, perseveration, and restrictive preference for

objects versus novel rats (46�53) (Fig. 2). These behaviors

were rapidly induced within 1�2 min after single infu-

sions, but were transient, lasting approximately 20�30

min, consistent with levels in bloods of propionic

acidemia patients, and the half-life of PPA (124).

Interestingly, it was found that during the acquisition

session in the Morris water maze and T maze, tests for

visuospatial learning, the PPA-treated rats displayed mild

or no impairment in maze acquisition but marked

impairments following maze reversal (49, 50). In the case

of the Morris water maze, when the location of the target

platform was moved for the reversal session, the PPA-

treated rats displayed marked cognitive impairments,

often swimming toward the former location of the plat-

form, consistent with ‘rigid’ perseverative behavior, which

could be considered a deficit in ‘un-learning’.

Using rats with stereotactically implanted indwelling

cortical, hippocampal, and subcortical brain electrodes,

repeated PPA exposures (once weekly�5 weeks) pro-

duced a kindling response in hippocampal/neocortical

leads and complex-partial-seizure-like behavior, while

tics, retropulsion, and dystonia were accompanied by

sharp spiking in the basal ganglia. Infusions of butyrate

and acetate produced similar, but less pronounced effects,

while infusions of propanol were without effect (46)

(Fig. 3). Current studies have found similar behaviors at

PPA doses 1/20th of initial investigations (unpublished

observations).

Thus, ICV infusions of PPA and, to a lesser extent,

other SCFAs rapidly produce a number of striking

behaviors that resemble those seen in patients with

ASDs. In addition, many of these behaviors appear to

reverse, consistent with the metabolic breakdown of PPA.

Potential underlying mechanisms of PPA andtheir relation to autism � neurotransmittersPPA and its related SCFAs are capable of gaining access

to the brain and inducing widespread effects on CNS

function (77), including neurotransmitter synthesis and

release, calcium influx, intracellular pH maintenance,

lipid metabolism, mitochondrial function, gap-junction-

dependent intercellular gating, immune activation, and

gene expression, which we have proposed may contribute

to the behaviors and biochemical findings observed in the

PPA animal model and ASDs (46).

A � PPA repetitive behavior

B � control rat

C � PPA social

D � control pair social

E � Ethovision pair

F � PPA object fixation

Fig. 1. Behavioral videos of propionic acid infusions in rats

(click headings to view videos). Single intracerebroventricular

(ICV) infusions (4 ml of 0.26 M solution over 4 min) of

propionic acid (PPA), a metabolic end product of autism-

associated enteric bacteria, produce bouts of reversible hyper-

active and repetitive behavior (A) in adult rats, compared with

phosphate-buffered saline (PBS) vehicle infused control rat (B).

Rat pairs infused with PPA show markedly reduced social

interaction and play behavior (C), compared with pairs of rats

infused with PBS vehicle (D), which show typical social

behavior. Ethovision behavioral tracking of control and PPA-

treated rat pairs (E), showing further evidence of PPA-induced

hyperactive, repetitive and antisocial behavior. PPA-treated rat

displays fixation on objects (F) and a specific object preferences

(i.e. block vs. sphere). PPA-infused rats also show turning, tics,

dystonia, and retropulsion and electrographic evidence of

complex partial seizures and basal ganglial spiking, consistent

with findings in patients with autism spectrum disorders.

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 5(page number not for citation purpose)

Page 23: Derrick MacFabe M.D.

Of interest, PPA is capable of altering dopamine,

serotonin, and glutamate systems in a manner similar to

that observed in ASDs (125�130), partly via potentiating

intracellular calcium release (77, 131, 132). Furthermore,

and of importance to the symptoms observed in both

ASDs and the PPA model, similar alterations to the

serotonin and dopamine systems have been implicated in

abnormal social and motor behaviors (127�129). SCFAs,

including PPA, may increase synthesis of dopamine

and its related catecholamines through induction of

tyrosine hydroxylase, a key enzyme in the synthesis of

catecholamines (133). This may also occur peripherally in

the adrenals and sympathetic ganglia, where they may

contribute to the enhanced anxiety-like behavior and

increased sympathetic tone and cardiovascular instability

found in ASDs (134).

PPA and other SCFAs are also known to potentiate glu-

tamatergic transmission, inhibit GABAergic transmission

Fig. 2. (A) Intracebroventricular (ICV) infusions of propionic acid (PPA) in adult rats increase repetitive locomotor activity (Versamax

automated behavioral assay). Group mean values (9SEM) total distance, number of movements, and movement time in rats given 4 ml

ICV infusions of phosphate-buffered saline (PBS) vehicle or PPA (0.26 M) twice daily for 7 days. BL�baseline session with no infusion;

T1�T7 consecutive treatment days. *pB0.05. (B) Single ICV infusions (4 ml of 0.26 M solution over min) of PPA and isomolar acetic

acid (SA) in adult rat pairs reduce social behavior (Ethovision automated behavioral assay). Behavior is measured as mean distance

apart (cm) and time spent (%) in 5 cm proximity during dark hours. Data points represent group means of data collected during 10 min

periods. Animals were placed in a large open field in same-drug pairs. Both PPA and SA pairs displayed a significantly greater mean

distance apart and spent significantly less time in close proximity to each other, consistent with impaired social behavior found in

autism, compared with phosphate-buffered saline (PBS) vehicle controls. Propanol the non-acidic alcohol analogue of PPA was without

effect. *�different from PBS control group at pB0.05 or better. (C) Tracks representing object and socially related behavioral

movement of adolescent rats receiving ICV infusions of PPA (4 ml of 0.26 M solution over 4 min) or PBS. More locomotion near the

caged novel rat (NR) occurred in PBS rats than by PPA rats. Graphic representation of duration of time (s) spent in close proximity

(within 18 cm) of the novel rat or novel object (NO). PPA rats showed less approach behavior and remain close to the novel rat less

than PBS rats, indicative of object preference over social behavior, consistent with findings in patients with autism. Graphic

representation of group mean (9SEM) percent correct turns in the T-maze task during acquisition (Days 1 and 2) and reversal (Days 3

and 4). ��significant difference from Day 2 performance; *�significant different from PBS control group. Figures modified with

permission from MacFabe et al. (2008) and (2011), and Shultz et al. (2008). Figure 2A reproduced with permission from Science

(American Association for the Advancement of Science). Figures 2B and 2C reproduced with permission from Elsevier Ltd.

Derrick F. MacFabe

6(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 24: Derrick MacFabe M.D.

and increase the production of enkephalin (46, 130, 135),

supportive of a potential mechanism for the enhanced

excitation/reduced inhibition theory of ASDs (136). PPA

is also known to modulate neurofilament phosphoryla-

tion/dephosporylation, important in neurodevelopment

and neuroplasticity (137, 138).

Fig. 3. (I) Intracerebroventricular (ICV) propionic acid (PPA) infusions in rats induce electrographic and behavioral elicitation

of kindled seizures and movement disorder. Kindled seizure manifestation in response to repeated weekly 4 ml ICV infusions of low

(0.052 M) or high (0.26 M) PPA, isomolar acetic acid (SA), propanol, or phosphate-buffered saline vehicle (PBS) in adult rats: group

mean (9SEM) data summed across the five initial testing sessions. Latency to epileptiform spiking was measured from the end of the

ICV infusion to the start of seizure. Maximum convulsion stage was rated using the Racine kindling scale. Duration of the longest

convulsion was of the longest continuous convulsion during a session. PPA and, to a lesser extent, SA induce a kindling response, while

propanol, the non-acidic analogue of PPA, was without effect. *�different from propanol and PBS controls; #�different from low

PPA, propanol control, and PBS control; ��different from propanol control; **�different from all other groups. (II) Abnormal

behaviors in response to ICV infusions: group mean frequency (9SEM) of behaviors per baseline or initial test session. Either one or

both doses of PPA increased the abnormal behaviors relative to control treatments. Only PPA produced dystonic (snake like),

retropulsive behaviors. With the exception of sodium acetate, which increased turning, no control treatments increased abnormal

behavior. Black bars indicate treated animals; white bars indicate PBS (vehicle)-treated animals. *�pB0.05 or better vs. all control

groups; #�pB0.05 or better vs. low PPA, propanol, and PBS control. (III) Representative electrographic seizure records from rat in

the high PPA group. (A) Session 2, short bout of epileptiform spiking accompanied by contralateral hindlimb dystonia coincident with

spiking (event marker). Note spiking in frontal cortex and caudate but not dorsal hippocampus. (B) Session 3, single epileptiform spikes

in caudate and frontal cortex but not hippocampus. Only the caudate spikes were accompanied by brief contralateral hindlimb dystonia

(event markers), which led the frontal cortex spikes by approximately 500 ms. A prominent frontal cortex spike (arrow) is not

accompanied by a spike in the caudate or by limb dystonia. (C) Session 5, bout of spiking in dorsal hippocampus not accompanied by

corresponding bouts of spiking in frontal cortex or caudate or by limb dystonia. (D) Session 5, single epileptiform spikes occur first in

the caudate and lead spikes in other traces. These are followed by a short bout of epileptiform spiking that is accompanied by brief

retropulsion, followed immediately by contralateral hindlimb dystonia, which ends coincident with the end of the bout of spiking (event

marker). (E) Session 5 at approximately 1 min after the records in D, single epileptiform spikes in all three traces, with the caudate

spikes leading the spikes in the other traces. Each spike is accompanied by brief contralateral hindlimb adduction and immediate

dystonia (event markers). (F) Session 5 at approximately 6 min after the records in E, short bout of epileptiform spiking with hindlimb

dystonia beginning coincident with caudate spiking (event marker), with frontal cortex and hippocampal spiking beginning after onset

of caudate spiking and hindlimb dystonia. This is followed 2 s later by the beginning of the first sustained kindled seizure displayed by

this rat (duration�35s), which was accompanied by the first conventional kindled convulsion (Stage 2), which did not include limb

dystonia. C�caudate; FC�frontal cortex; H�dorsal hippocampus; HD�hindlimb dystonia. Amplitude calibration�50 mv; time

marker�5 s. Figures modified with permission from MacFabe et al. (2007). Reproduced with permission from Elsevier Ltd.

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 7(page number not for citation purpose)

Page 25: Derrick MacFabe M.D.

PPA-induced intracellular acidification,increased oxidative stress, impaired antioxidantcapacity, and gap junction closureThe findings that PPA and its related fatty acids often

induce similar behavioral effects, whereas isomolar infu-

sions of propanol do not (46, 49, 50), make it tempting to

speculate that the carboxylate functional group and/or

acidic properties of PPA may be involved in its effects

on behavior. Once low-molecular-weight organic acids

are protonated in acidic conditions, they become more

lipid soluble and, thus, gain access to the CNS, either by

passive diffusion or active transport (77, 78, 80). Once

they enter the CNS, they may concentrate within cells

and induce intracellular acidification. Previous studies

have found that PPA- and acetate-related acidosis in rats

can alter social behavior (139). Furthermore, clinical

metabolic acidosis from varied etiologies in humans often

involves bouts of confusion and movement disorder

similar in some respects to those found in ASDs (140).

The effects of varying intracellular pH on cellular

physiology are broad. Notably, PPA-induced intracellular

acidification can inhibit mitochondrial function, which is

vital for normal metabolism of other fatty acids (141).

Intracellular accumulation of organic acids creates a

buildup of acyl-coenzyme, which interrupts metabolism

(141). PPA sequesters carnitine function and inhibits

CoA function and, thus, impairs mitochondrial metabo-

lism and energy production, leading to further decrease in

cytoplasmic pH via the accumulation of other organic

acids (131, 141). This is consistent with the impaired

mitochondrial metabolism in ASDs, including carnitine

deficiency, mitochondrial dysfunction, and systemic ele-

vations of nitric oxide metabolites (20, 46, 142). In

addition, valproate, an antiepileptic drug, structurally

related to PPA, and a known prenatal risk factor for

ASDs (143), similarly alters mitochondrial metabolism

and causes the depletion of carnitine stores and ence-

phalopathy (144, 145). Nitropropionic acid, another PPA

derivative, is a potent inhibitor of mitochondrial function

and produces a model of Huntington chorea when

administered to rodents (146). The basal ganglia may

be particularly sensitive to PPA and its derivatives,

because of the region’s high metabolic demand and

damage in many movement disorders, including organic

acidemias and mitochondrial disorders, thus offering

an explanation of the tics, dystonias, and repetitive be-

haviors in these conditions and the PPA rodent model

(46, 86).

An impaired mitochondrial metabolic process similar to

the one described could lead to a range of negative effects

on the CNS. Of relevance, similar encephalopathic pro-

cesses associated with increased oxidative stress, such as

those found in organic acidemias, are known to produce

symptoms consistent with findings in the PPA model and

ASDs (86, 93, 147). Consistent with this hypothesis,

biochemical analyses of homogenates of the brain sample

from PPA-treated rats (4 ml of 0.26 M solution BID�7

days, ICV) demonstrated an increase in oxidative stress

markers (protein carbonylation and lipoperoxidation), as

well as abnormalities in glutathione-associated pathways,

particularly in brain regions implicated in ASDs (46, 47,

148) (Fig. 4).

Glutathione participates in both antioxidant defense

and xenobiotic detoxification over a broad range of

environmental organic compounds and metals (149,

150), including those implicated in ASDs. Impairments

in glutathione-associated pathways suggest reduced cel-

lular defense and are considered markers of increased

oxidative stress (149). Of particular interest is the evi-

dence of genetic and acquired impairments in glutathione-

associated pathways in patients with ASDs (150�152),

suggesting a plausible mechanism for altered sensitivity to

a wide range of environmental agents (metals, pesticides,

drugs) proposed in ASDs (46, 47). Of note, acetamino-

phen, when given for common pediatric illnesses, may

overwhelm glutathione metabolism and has been pro-

posed as a possible trigger for ASDs (153�155). Further-

more, even brief exposures to agents that alter redox

levels in cells early in development may change cellular

developmental trajectory and ultimate cell fate, which

may provide a plausible mechanism for neurodevelop-

mental alterations in ASDs (156). Overall, PPA-induced

metabolic abnormalities and oxidative stress are con-

sistent with findings from ASDs and ASD-related

disorders.

Further to its capability to decrease intracellular pH,

both directly via intracellular concentration and indir-

ectly via inhibition of mitochondrial functions, we have

proposed that a major effect of PPA is through the

closure of gap junctions via intracellular acidification

(46, 50, 157). As reviewed in MacFabe et al. (46), gap

junctions are intercellular channels composed of con-

nexin proteins that are gated by a number of factors

influenced upon by PPA, including dopamine, calcium,

and cytokines. Gap junctions play a major role in cel-

lular differentiation, and in particular, peripheral nerve,

cardiac, uterine, and gastrointestinal function. How-

ever, in the CNS, gap junction coupling is vital for the

synchronization of neural electrical activity within dis-

crete functional cell groups and is more extensive during

early brain development and neuronal migration. Astro-

cytes are electrotonically connected by gap junctions,

forming a syncytium to spatially buffer calcium, gluta-

mate, and potassium (158), and apoptotic factors are

capable of passing through these glial gap junctions (159,

160). Thus, closed glial gap junctions may render neurons

hyperexcitable to rising extracellular potassium and

glutamate (160), while closed neuronal gap junctions

would be neuroprotective (159). In turn, this decrease in

gap junction coupling may lead to inhibited cortical

Derrick F. MacFabe

8(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 26: Derrick MacFabe M.D.

Fig. 4. (A) Propionic acid (PPA) infusions in rats induces a significant increase in lipid and protein oxidation, supportive of increased

oxidative stress via increased oxidant production and decreased antioxidant defenses in homogenates of discrete regions of rodent brain

consistent with findings in patients with autism. Interestingly, brain stem appears relatively unaffected. (B) PPA-treated rats showed

significantly decreased total GSH, a decreased activity of GPx, but an increase in the activity of GST. Conversely, the activity of GR was

unchanged. These findings suggest that GSH may be involved in the metabolic clearance of PPA, or alternatively, the synthesis of GSH

was impaired by PPA. The observed reduction of the activity of GPx may be due to the decreased availability of GSH, a cofactor of

GPx or may reflect an overall reduction in antioxidant defenses induced by PPA. Black bars indicate PPA-treated animals; white bars

indicate PBS (vehicle)-treated animals. Figures modified with permission from MacFabe et al. (2008). Reproduced with permission from

Science (American Association for the Advancement of Science).

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 9(page number not for citation purpose)

Page 27: Derrick MacFabe M.D.

Fig. 5. (Continued)

Derrick F. MacFabe

10(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 28: Derrick MacFabe M.D.

pruning in development, consistent with the larger brain

size found in ASDs (46). Gap junction communication is

involved in neurotransmission in areas that are impli-

cated in seizure and movement disorders, such as the

basal ganglia, prefrontal cortex, nucleus accumbens, and

hippocampus. Intrastriatal injections of gap junction

blockers produce stereotypical movements, hyperlocomo-

tion, and disruption of motor sequencing in rodents (127,

161). Furthermore, gap junction knockout mice show

abnormal brain development, exaggerated responses to

neurotoxic insults, seizure disorder, and abnormal beha-

viors (162). Interestingly, gap junction blockers also

inhibit tight junctions in many cellular systems (163),

thereby possibly contributing to altered barrier function

in vascular endothelium and gut in ASDs (107). Given

these findings, it seems possible that PPA-induced

alterations to gap junction function may contribute

to electrophysiologic and motor impairments observed

in both the PPA model and ASDs, but also in neural

development, as well as systemic effects (i.e. gastrointest-

inal motility). These potential mechanisms are the subject

of further study in our group.

However it is unclear at this stage whether these effects

on pH and oxidative stress are causal to behavioral

induction or the results of PPA dependent mechanisms

such as neuroinflammation or altered lipid metabolism.

Quite possibly they are not necessarily exclusive and may

be mutually reinforcing, leading to a vicious cycle of these

processes.

Neuropathological effects of PPA infusions oninflammation/CREB induction/epigenetics/lipidtransportWe have performed a number of studies examining the

neurohistological effects of PPA and its related fatty acids

on innate neuroinflammatory processes (46�50). Immu-

nohistochemical examinations of the brain tissue from

PPA-treated rats (1�14 days), demonstrated by DAB

immunostaining and semiquantitative image densitome-

try, show a neuroinflammatory response, characterized

by increased activated microglia and reactive astrog-

liosis in some brain areas, including the hippocampus,

cingulate, neocortex, and white matter. Interestingly

these effects occur in the absence of apoptotic neuronal

loss, as measured by the staining of cleaved caspase 3

(Fig. 5).

These findings are consistent with those from autopsy

cases from patients with ASDs, regardless of age of death,

indicating that an ongoing inflammatory process may be

present throughout the life of the individual (13, 164).

Microglia are a heterogeneous family of cells of myeloid

lineage, which migrate from the bone marrow to the CNS

both pre- and postnatally, and play a currently under-

appreciated role in neurodevelopmental disorders (165).

When activated, they produce high levels of proinflam-

matory cytokines and reactive oxygen species such as

hydrogen peroxide and nitric oxide. Interestingly, micro-

glia also may be involved in neuroplastic responses, such

as synaptic reorganization, and cortical pruning (166).

Such effects are conceivable in both ASDs and the PPA

model, as specific SCFA receptors exist on immune cells

(88), and increased levels of cytokines such as tumor

necrosis factor and macrophage chemoattractant protein

are found in ASDs (13).

In addition, immunohistochemical analysis of PPA-

treated brain sections showed evidence of increased

immunoreactivity of the activated, phosphorylated form

of cyclic-AMP responsive element binding protein

(pCREB). Likewise, the results of in vitro studies show

that SCFAs can induce pCREB in PC12 cells, leading to

increased catecholamine synthesis (46, 133). CREB is

interesting as this important neuroregulatory protein

known to play a key role in the epigenetic expression

of a number of genes implicated in neuroplasticity,

addiction, movement and mood disorders, and memory

acquisition (167). Thus, it is possible that increased

activation of CREB-dependent epigenetic modulation

of memory- or movement-related pathways following

PPA administration could result in normal memory

acquisition with perseverative behavior, repetitive beha-

vior, or seizure similar to those observed in the PPA

model and ASDs (46). Further to these findings, the fact

that some other SCFAs and valproate, a compound with

some structural similarity to PPA and a risk factor for

autism, also are capable of further gene regulation via

their histone deacetylase inhibition activity (90, 133, 168),

makes the epigenetic effects of these compounds on

ASD-related genes a subject of further study in our

laboratory (169).

Fig. 5. Neuropathology (avidin�biotin complex immunohistochemistry) and semiquantitative image densitometry of coronal brain

sections of dorsal hippocampus (CA2) and external capsule of adult rats with 14 day BID ICV infusions of Propionic acid (PPA) or

phosphate-buffered saline (PBS). PPA induced significant reactive astrogliosis (anti-GFAP) and microglial activation (anti-CD68),

without apoptotic neuronal cell loss (anit-cleaved caspase 3) in rat hippocampus, similar to finding in autopsy brain from patients

with autism. Nuclear translocation of anti-CREB and an increase of anti phosphoCREB immunoreactivity are observed in neural,

glial, and endovascular epithelium by PPA treatment, suggestive of gene induction. PPA increases monocarboxylate transporter

1 immunoreactivity, primarily in white matter external capsule, suggestive of alterations in brain shot-chain fatty acid transport/

metabolism. Black bars indicate PPA-treated animals; white bars indicate PBS (vehicle)-treated animals. Horizontal measurement

bar�100m. Figures modified with permission from MacFabe et al. (2007); see publication for details of immunostaining procedures.

Reproduced with permission from Elsevier Ltd.

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 11(page number not for citation purpose)

Page 29: Derrick MacFabe M.D.

Fig. 6. (Continued)

Derrick F. MacFabe

12(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 30: Derrick MacFabe M.D.

Preliminary results following immunostaining of brain

sections with anti-monocarboxylate transporter 1 anti-

body (1:1000 dilution, Chemicon AB1286), a major

transporter of PPA, related SCFA and ketones (79, 80,

82, 83), reveals significant increases in external capsule

white matter immunoreactivity following repeated PPA

infusions, suggestive of altered brain fatty acid transport.

It is important to note that these neuropathological

changes, although consistent with findings in ASD

autopsy cases, occur hours or days following PPA

administration, which is considerably later than the

transient behavioral effects, which are induced within

minutes and last approximate 20�30 min. Nonetheless,

they may be important in the neuroplastic kindling

response and perseverative learning behavior from

repeated exposures of PPA and merit further study.

Effects of PPA on lipid metabolism andmitochondrial functionSeveral reports have indicated that abnormal lipid

metabolism may occur in ASDs and ASD-related dis-

orders (170�175), including fatty acids, phospholipases

A2, and membrane phospholipids. Clinical studies sug-

gest improvements in core symptoms of patients with

ASDs following supplementation with polyunsaturated

fatty acids (PUFAs) (173, 176, 177) or cholesterol (178).

PUFAs, the major lipid constituents of neuronal mem-

branes, are absolutely essential for normal brain devel-

opment and function, modulating membrane fluidity,

gene expression, cell signaling, neuronal excitability, and

oxidant protection, and provide a source of energy for the

cells, particularly during early neurodevelopment (110,

179). Lipid composition in cellular membranes is parti-

cularly important in the formation of lipid rafts, which

alter membrane fluidity and the function of membrane-

bound proteins, including cell receptors and gap and tight

junctions (51, 53, 107, 180).

Carnitine is a quaternary ammonium compound

synthesized from the amino acids lysine and methionine

principally by liver and kidney, and it is also obtained

from diet. It is critical for the transport of fatty acids into

the inner mitochondrial membrane for b-oxidation and

energy production (181). Clinical studies have reported a

relative carnitine deficiency and abnormal acylcarnitine

profile (elevated short and long chain) in erythrocytes

of ASDs of uncertain etiology (142, 182, 183) and a

common X-linked genetic defect in carnitine synthesis in

a subset of patients with ASDs (25), suggesting carni-

tine supplementation as a possible treatment for the

disorder (20).

We thus used our PPA rodent model to examine

whether there is any evidence for alterations in brain

phospholipids and acylcarnitines following intraventricu-

lar infusions with PPA or butyrate (4 ml of 0.26 M

solution BID�7 days) to adult rats (51, 53).

Brain lipid analysis was performed via GC mass

spectroscopy/electrospray ionization mass spectrometry

(see Thomas et al. (51, 53) for technical details). Altered

lipid profiles were observed in rat brain phospholipids

following infusion with both PPA and, to a lesser extent,

butyrate (Fig. 6).

PPA infusion resulted in decreased levels of total

monounsaturates and total v6 fatty acids and elevated

levels of total saturates in all of the studied phospholi-

pids. In addition, a decline in total plasmalogen phos-

phatidylethanolamine and the ratio of v6:v3 was also

present. Elevated levels of saturated fatty acids have been

reported in red blood cells (170, 176) and plasma (184) of

several autistic patients. These findings were accompa-

nied by a concomitant decrease in total monounsaturates,

particularly, 18:1n9 fatty acid. A decline in the level of

this fatty acid along with several other monounsaturates

(20:1n9, 22:1n9, 16:1n7, and PE 24:1n9) has also been

observed in bloods drawn from patients with autism (174,

176, 185).

Carnitine and acylcarnitine analysis showed a non-

significant trend toward lower free carnitine in butyrate-

and PPA-treated animals; however, there was a consistent

significant (p�0.02) increase in total acylcarnitines, total

long-chain (C12 to C24) acylcarnitines, total short-chain

(C2 to C9) acylcarnitines, and the ratio of free to bound

carnitine following infusions with PPA and butyrate.

Increases in the accumulation of short- and long-chain

acylcarnitines, but not medium-chain acylcarnitines,

Fig. 6. (A) Changes in phosphatidylethanolamine and phosphatidylcholine in rat brain tissue following repeated 4 ml intracerebroven-

tricular infusions with phosphate-buffered buffer (PBS), butyric acid (BUT), and propionic acid (PPA) 0.26 M twice daily for 7 days are

consistent with findings in patients with autism. Values represent means9SE. Arrows indicate significant difference between treatments

[PPA and BUT compared with the control (PBS)], and direction of the changes (increase or decrease) at LSD�0.05, N�6 per

treatment. Monounsat�monosaturates; polyunsat�polyunsaturates. (B) Changes in acylcarnitine and ratio of bound to free carnitine

in rat brain tissue following intracerebroventricular infusion with PBS, BUT, and PPA are consistent with findings in patients with

autism. Values represent means9SE. Means accompanied by different superscript letters (e.g. a, b, and c) indicate significant difference

between treatments at LSD�0.05, N�6 per treatment. Long-chain acylcarnitine�C12�C24, short-chain acylcarnitine�C2�C9. (C)

PPA and, to a lesser extent, BUT infusions increase short- and long-chain acylcarnitines, but not medium-chain acylcarnitines, similar

to findings in patients with autism. Values (percent by weight) represent means9SE. Means in the same row accompanied by different

superscript letters (e.g. a, b, and c) are significantly different between treatments at LSD�0.05, N�9. Numbers preceding C represents

carbon number of the fatty acid moiety attach to carnitine. Long-chain acylcarnitine�C12�C24, short-chain acylcarnitine�C2�C9.

Figures modified with permission from Thomas et al. (2010). Reproduced with permission from Wiley Blackwell.

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 13(page number not for citation purpose)

Page 31: Derrick MacFabe M.D.

were observed. Specifically, when these acylcarnitines

were grouped according to chain length, acetyl carni-

tine (C2) was the major contributor to the increases ob-

served in short-chain acylcarnitines, while C16:0, C18:0,

C18:1, C22:0, and C22:1 were the major contributors

to the increases observed in long-chain acylcarnitines

(51, 53).

These unique acylcarnitine profiles are similar to those

obtained from blood samples of patients with ASDs (182,

183) which observed elevations in the levels of some long-

chain acylcarnitine species (14:1 and 14:2). Elevations in

14:1 and other long-chain acylcarnitines (16:0, 18:0, 18:1,

22:0, and 22:1) consistent with findings in our model.

PPA is thought to affect mitochondrial fatty acid

metabolism by binding to propionyl CoA and by

sequestering carnitine (86, 186, 187). Elevation in the

levels of these acylcarnitines indicates a potential meta-

bolic disturbance by PPA infusions to affect mitochon-

drial metabolism in manners consistent with a unique

acylcarnitine profile and a proposed environmental

induction of mitochondrial dysfunction and increased

oxidative stress in ASDs (51, 188).

Of particular interest, PPA is a known inhibitor of

mitochondrial function, through sequestration of carni-

tine and the production of propionyl CoA, a potential

cytotoxin (189, 190). PPA also inhibits the incorporation

of acetate during lipid synthesis, an effect which is more

pronounced in males (191, 192). This is interesting in

light of low cholesterol levels being central to many forms

of ASDs (178). Since PPA is metabolized through the

mitochondrial metabolic pathways and affects mitochon-

drial cardiolipin levels and membrane stability (51, 53),

we propose that excess exogenous PPA or its related

SCFAs overwhelm mitochondrial metabolism, thereby

causing mitochondrial dysfunction. Such mitochondrial

dysfunction is consistent with an increase in oxidative

stress. Evidence for oxidative stress has been demon-

strated in the rodent PPA model of ASDs by the increase

in nitric oxide-/hydrogen peroxide-producing activated

microglia, coupled with the elevation of oxidized proteins

and lipids, reduction of glutathione, and alteration of

phospholipid/acylcarnitine profiles in brain homoge-

nates, particularly in brain regions affected in ASDs

(46, 51, 53). Such findings are all consistent with those

obtained from patients with ASDs (110, 151, 152,

193, 194).

Impairment of carnitine metabolism from avariety of causes may be central to ASDpathogenesis and regressionAlthough the basis of the reduced blood carnitine levels

in ASDs remains unclear, we have made the observation

that there are diverse clinical conditions circumstantially

linked to ASDs and gastrointestinal dysfunction that

show disruptions in carnitine metabolism as a common

observation. There has been a growing interest in the role

of carnitine in brain physiology and disease (181),

particularly in brain GABAergic and astrocyte metabo-

lism (195). As carnitine is endogenously produced from

lysine and methionine, persons with defects in methyla-

tion pathways, common in individuals with ASDs (196),

would have impairments in endogenous carnitine synth-

esis. A particularly vulnerable subgroup would be the

recently discovered X-linked defect in the TMLHE

enzyme, responsible for the first step in carnitine bio-

synthesis, which is a risk factor for non-dysmorphic

autism in males (25). Collectively, these persons would

rely more on dietary sources of carnitine, critical during

periods of rapid development, and thus may be more

sensitive to factors that impair carnitine uptake from

the gut. It is known that carnitine is transported across

the gut�blood and blood�brain barriers via the Na�-

dependent organic cation/carnitine transporter 2

(OCNT2) (197). Carnitine transport deficits have been

implicated in colitis (198) and also may impair blood�brain barrier integrity, allowing non-neurotropic influ-

enza A virus to enter the CNS and inducing a neonatal

encephalopathy (199). Interestingly, long-term adminis-

tration of common antibiotics (i.e. b-lactams) for routine

pediatric infections, in addition to altering gut flora

favoring PPA-producing species, has been shown to

directly inhibit the OCNT2 transporter, affecting carni-

tine transport across gut�blood and blood�brain barriers

(197), and, thus, may contribute to a relative systemic

carnitine deficiency. Additionally, antibiotics given as

forms of pivalyl esters will cause an increased urinary loss

of carnitine. Such interferences could be significant

considering the reported high incidence of antecedent

long-term antibiotic use in some patients with ASDs (65,

66, 200, 201) and the finding of unique enteric PPA-

producing bacteria and gut carbohydrate malabsorbtion

in regressive ASDs (65, 66, 105). This offers a potential

explanation for autistic regression and also temporary

behavioral improvements in some patients following

vancomycin or metronidazole treatment, which transi-

ently eradicates these bacteria (65, 66, 109, 201). How-

ever, given that Finegold has proposed eradication of

ASD-causing bacteria with oritavancin and aztreonam,

antibiotics which could further depress carnitine absorp-

tion, as a possible treatment of ASDs, warrants the

necessity of following patient carnitine levels and provid-

ing possible carnitine supplementation in such a study

(64). Furthermore, removal of refined carbohydrates

from the diet, which has been suggested as an empiric

treatment to improve the behavioral fluctuations and

gastrointestinal symptoms in ASDs, may act by reducing

substrate for these bacteria to produce PPA (46). Feeding

of a high-carbohydrate diet in rats is known to increase

SCFA levels and produce anxiety and aggressive behavior

(139). Interestingly, preeclamptic mothers, who have an

Derrick F. MacFabe

14(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 32: Derrick MacFabe M.D.

increased risk of having offspring affected by ASDs

(202), have similar short- and long-chain acylcarnitine

profiles (203) as found in the patients with ASDs and the

PPA rodent model. Although the overall relationships

remain unproven, it appears that taken collectively,

we have proposed the above observations link the

decreased carnitine levels in some patients with ASDs

with several genetic and environmental factors consistent

with regression, gastrointestinal symptomatology, micro-

biology, and lipid biomarkers and with experimental

findings obtained using the PPA model. Furthermore,

oral carnitine and its derivative acetyl-l-carnitine

have both neuroprotective (181, 204, 205) and colopro-

tective properties (206) and deserve further investigation

as therapeutic agents in developmental disorders, includ-

ing ASDs (22, 46, 51, 207). These findings would also

warrant more rigorous and possibly repeated carnitine/

acylcarnitine screening of ‘patients at risk’, such as

infants with clinical evidence of developmental delay

and gastrointestinal dysfunction, particularly in the

presence of maternal/infant hospital-acquired infection

or long-term antibiotic use.

Conclusions/limitations/future directionsIt is important to note an animal model is unlikely to

completely replicate a human disease. The usefulness of

models relates to the various types of validity that can be

shown to exist for specific conditions. Face validity, or

the degree to which the model is able to capture

the phenomenology of the disorder, is usually an essential

first step in developing a model (208, 209), but addi-

tional evidence for construct validity, that is, a probable

theoretical rationale for the model, possibly based

on etiology is equally as important. Crawley (120)

has provided a list of symptoms of ASDs that would

aid in the establishment of face validity for a rodent

model, along with suggested behavioral tests for such

symptoms.

In the initial development of our model of ASDs, we

focused on four broad aspects, such as behavioral, brain

electrographic and neuropathological characteristics and

biochemical markers, which would provide some face, as

well as, construct validity for the model. At a behavioral

level, we were interested in determining whether PPA

would induce hyperactivity, stereotypies and repetitive

behaviors, object preference, perseveration, and social

impairment consistent with ASDs. Electroencephalo-

graphic recordings allowed us to monitor for cortical

and subcortical epileptiform activity and develop-

ment of seizures along with behavioral assessment of

seizure and movement disorder. Finally, we also looked

for possible neuropathological effects that might be

consistent with those observed in humans, such as

innate neuroinflammatory, astroglial and microglial

changes, and neuroplastic (i.e. CREB activation, mono-

carboxylate transport) changes, as well as biochemical

markers suggestive of increased oxidative stress, altered

phospholipid/acylcarnitine profiles, and mitochondrial

dysfunction.

These initial studies provide support for the face validity

of the intraventricular PPA administration in adult rat

model of autism, as the behavioral and electrographic

changes observed resemble those seen in the human

condition. The rapid induction and transient nature of

these peculiar behavioral and electrographic effects, their

potentiation with repeated exposures, and the absence

of such behaviors in rats receiving control compounds

(1-propanol) suggest the involvement of diverse PPA-

activated neural mechanisms and present the first attempt

in the field to model the possible fluctuating, as opposed to

static behavioral course in ASDs.

Evidence from human studies suggests that ASD is

a condition that represents an ongoing neuroinflamma-

tory or neurometabolic disorder possibly resulting from

an increased sensitivity to oxidative stress or acquired

mitochondrial dysfunction from a variety of environmen-

tal risk factors. The neuropathological and biochemical

findings of our model support this hypothesis. Interest-

ingly, the observed impairments in the glutathione system,

increased oxidative stress, and altered phospholipids/

acylcarnitine profiles, suggestive of mitochondrial dys-

function in our model, are consistent with the findings in

human autism cases and would provide a plausible

mechanism for increased environmental sensitivity to a

variety of agents. The similarities in neuropathological

and biochemical changes between the animal model and

human ASD cases, coupled with the presence of PPA-

producing bacteria, could represent similar underlying

etiological processes subsequent to central nervous system

insult with PPA. However, much additional work needs to

be done before these similarities can be taken as support

for construct validity of the animal model.

It is important to note, other than reports of increased

PPA in stool samples, studies are lacking (210) that have

systematically examined PPA and its related metabolites

in ASDs. The short half-life and rapid metabolism via boxidation, incorporation into acylcarnitines and other

lipids, and intracellular concentration of PPA via specific

transporters and pH-dependant mechanisms also make

definitive measurement and interpretation difficult. How-

ever, there is some indirect evidence in patients with

ASDs suggestive of increased PPA, including a relative

carnitine deficiency, unique phospholipid and short- and

long-chain acylcarnitine profiles, elevations of nitric

oxide metabolites, and oxidative stress markers, known

to be elevated by PPA in experimental systems, including

our rodent model. The underappreciated incidence of

complex gene polymorphisms in propionic acidemia in

some populations, a potential population of partial

metabolizers of PPA, and documented ASD symptoms

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 15(page number not for citation purpose)

Page 33: Derrick MacFabe M.D.

in some of these patients, is interesting. The antiepileptic

drug valproate, a prenatal risk factor for autism, alters

both fatty acid and biotin metabolism and causes the

depletion of carnitine stores each of which could

theoretically increase PPA levels. Prenatal exposures to

ethanol, known to produce developmental delay, also

increases PPA levels to millimolar levels putatively by

depleting intracellular carnitine stores.

Despite the lack of studies on PPA and its related fatty

acids in ASDs, abnormalities of lipid metabolism and

mitochondrial dysfunction have been postulated as a

potential cause of the condition. Patient studies have

noted impairments in mitochondrial b-oxidation of fatty

acids, reductions in essential PUFAs levels, and anecdotal

reports of improvement in some patients with essential

fatty acid and carnitine supplementation.

Likewise, the intriguing finding that long-term admin-

istration of common antibiotics for routine infections not

only may favor PPA-producing bacteria but also may

impair colonocyte function and directly inhibits gut

carnitine transport, leading to an acquired mitochondrial

encephalopathy, should certainly be investigated as a key

factor in ASD regression. Notably, screening for unique

enteric bacterial populations, plasma phospholipids/acyl-

carnitines, inherited carnitine transport defects, ‘partial

metabolizers’ of PPA (i.e. organic acidemias, mitochon-

drial disorder), and gut dysfunction deserves further

examination. This may be useful not only in patients

with ASDs but also in ‘persons at risk’, including

individuals in neonatal units, those with immune defi-

ciencies and patients on long-term antibiotic usage.

Likewise, to follow these biomarkers in response to

proposed treatments (eradication of PPA-producing

bacteria, probiotics, dietary carbohydrate restriction,

carnitine supplementation) in animal models and ulti-

mately patients would prove valuable. Importantly, the

CAUSES

Gut dysmotility/inflammation/carbohydrate malabsorption/altered gut permeability (tight junction impairment)

Hospitalisation (colonization of nosocomial bacteria) i.e. C-section, neonatal distress

Active uptake of SCFA to gut and CNS(monocarboxylate transporters)

Prenatal drugs (valproate, ethanol) pH dependent intracellular concentrationof SCFA

Opportunistic infection (Clostridium spp.,Desulfovibrio spp.)

Neurotransmitter synthesis and release(catecholamines, enkephalins)CNS/sympathetic nervous system

Maternal/Infant gut dysbiosis, decreased “beneficial” bacteria

Receptor activity (+NMDA, -GABA)SCFA G protein coupled receptors/Ca2+

influx

Organic acidemias (propionic/methylmalonic, biotinidase holocarboxylase deficiency)

Gap junction closure, alteredneurodevelopment, neuroinflammation

vitamin deficiencies (B12/biotin deficiency) Impaired mitochondrial function/increased oxidative stress

Genetic/acquired impaired carnitine synthesis/absorption (TMLHE/OCTN2 genes, β-lactam antibiotics)

Reduced glutathione/increasedsensitivity to xenobiotics (i.e.acetaminophen)

Mitochondrial disorder/dysfunction (inherited, acquired) (i.e. H2S, pesticides, metals)

Decreased carnitine/altered lipidmetabolism/membrane fluidity

Colitis (impaired barrier/SCFA metabolism), i.e. celiac disease, OCTN2/Met-receptor tyrosine kinase mutation

Altered gene expression (CREBactivation, histone deacetylaseinhibition)

Increased refined carbohydrate consumption – substrate for bacterial fermentation

Antisocial/perseverative/anxiety-like behavior, seizure/movement disorder,Restrictive food interests/carbohydratecraving

CONSEQUENCES OF SCFAs

Long term antibiotics for routine infection(maternal/infant) Treatment of maternal β hemolytic strep

Fig. 7. Potential causes and consequences of increased enteric short-chain fatty acid production and/or decreased breakdown and their

relation to autism spectrum disorder. These findings, which are not mutually exclusive, may contribute to the pathophysiology,

behavioral symptoms, and comorbidities of autism.

Derrick F. MacFabe

16(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 34: Derrick MacFabe M.D.

longitudinal examination of the infant gut microbiome,

characterizing not only the presence, but also the absence,

of specific gut bacterial population in infants who

progress neurotypically or subsequently go on to develop

ASDs will be critical. Novel bacterial species, such as

Desulfovibrio, and their ability to produce H2S, a possible

mitochondrial toxin, possessing both colonotoxic and

neurotoxic effects synergistic to PPA, and their inter-

action with Clostridia and other members of the gut

microbiome deserve further investigation. Lastly the

neurobiological effects of other short chain fatty acids,

their derivatives and the array of other metabolites of the

microbiome remain largely unexplored.

In a broader context, it is intriguing to postulate that

microbiota, through natural selection, have evolved to

use their metabolites to modulate the physiology and

ultimately behavior of the host, to promote survival.

This has been documented in behavioral neurobiology,

with such examples as cordyceps fungus producing

climbing behavior in ants, and borna and rabies viruses

eliciting salivary transmission and biting behavior in

mammals (211, 212). It has recently been established that

transplantation of fecal material in germ-free mice can

alter brain gene expression and phenotypic behavior of

the host (213). In light of this, the observation of

restrictive eating of carbohydrates, diarrhea and fecal

smearing in patients with ASDs which could theoreti-

cally promote organism growth and spread, is intriguing.

It is also worth noting that many of the effects of lower

doses of SCFAs on gut physiology and immune function

are indeed beneficial to host and ultimately bacterial

survival. Finally, it is important to note the ability of

SCFAs to elicit anxiety-like, perseverative, repetitive,

ritualistic, and antisocial behaviors that are common

to many other neuropsychiatric conditions such as

obsessive compulsive, anxiety, attention deficit/hyperac-

tive, mood, and eating disorders, irritable bowel syn-

drome, pediatric autoimmune neuropsychiatric disorder

associated with streptococcal infections (PANDAS),

and schizophrenia, where infectious agents have been

proposed (214�216).

The growing incidence of ASDs and ASD-related

conditions, coupled with the observed alterations in the

human microbiome secondary to dietary, medical, and

agricultural factors, and their potential effect on human

and animal behavior, should be further examined. The

impact of human migration and urbanization, domes-

tication of plant and animals, and resultant human

diseases shaping cultures is not trivial and has been

discussed elsewhere (217).

In conclusion, these present studies examined the

effects of intracerebroventricular PPA in adult rats and

their relation to the pathophysiology of ASDs. Given

the intriguing multiple effects of PPA on many neurolo-

gical, gastroenterological, metabolic, and immunological

processes, together with our initial findings on behavior,

electrophysiology, neuropathology, and biochemistry, the

rat intraventricular PPA model maybe a useful paradigm

for the examination of the disparate symptoms and

pathophysiology of ASDs (Fig. 7).

However, notwithstanding of the remarkable effects of

these compounds on an ‘intact’ adult brain, autism is,

of course, a neurodevelopmental disorder, with evidence

of altered nervous system development, involving white

matter abnormalities and disorders of neuronal micro-

circuitry. Treatments of rodents with PPA, other SCFAs

and appropriate control compounds, and ultimately

infection of ASD-associated bacteria during critical times

of pre- and postnatal development are essential steps in

extending the validity of this animal model. Such studies

using systemic and dietary exposure at these time periods

are ongoing in our laboratory and our collaborators (52,

130, 218).

Although compelling, further research examining PPA

and its related SCFAs is still needed at both the basic and

the clinical levels to better understand the underlying

mechanisms of how these and other metabolic products

of the host microbiota may modulate host physiology

throughout the lifecycle and whether they are directly

involved in the disparate brain and behaviors of ASDs

and ASD-related conditions. Conceptually, it is the

author’s opinion that the pathophysiology of ASDs may

be more completely understood as being similar to

conditions such as ethanol intoxication, or diabetes,

and the resultant complex interactions between diet,

genetics, metabolism, host microbiome, and behavior,

that are well known to exist in these treatable disorders

throughout the life cycle. Considering the marked in-

crease, considerable morbidity, and social burden of

ASDs, collaboration in the fields of microbiology, clinical

and basic neuroscience, immunology, biochemistry, gas-

troenterology, obstetrics, genetics, and epidemiology is

warranted and should be encouraged.

Acknowledgments

This manuscript was prepared as a summary of two lectures given in

1) honor of the 100th birthday of Dr. Philip Trexler and in memory

of the late Dr. Morris Pollard (Notre Dame University, South

Bend, IN) and 2) as a Nobel Symposium lecture in ‘The Gut and the

Brain � with Focus on Autism Spectrum Disorder’ (Karolinska

Institute, Stockholm, Sweden). The author would like to thank

all the investigators, collaborators, students, and technologists

of the Kilee Patchell-Evans Autism Research Group, University of

Western Ontario, Canada. Particular thanks go to Drs. Klaus-Peter

Ossenkopp, Martin Kavaliers, Donald Peter Cain (behavior), and

Fred Possmayer (biochemistry); graduate students/postdoctoral

fellows Kelly Foley, Jennifer Hoffman, Melissa Meeking, Jennifer

Mepham; and Drs. Sandy Shultz (behavior), Karina Rodriguez-

Capote (oxidative stress), and Raymond Thomas (lipid biochemis-

try). Expert technical assistance was provided by Roy Taylor

(pathology), Francis Boon (behavior, surgery, EEG), and Lisa

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 17(page number not for citation purpose)

Page 35: Derrick MacFabe M.D.

Tichenoff (Project Manager and generous assistance with manu-

script preparation). We would also like to express our utmost thanks

to David Patchell-Evans, for his tireless devotion to persons with

autism, and his daughter, Kilee Patchell-Evans. Our heartfelt thanks

go out to countless parents and caregivers of persons with autism

who have shared their stories. This research was supported by

contributions from GoodLife Children’s Charities, Autism Canada,

and Autism Research Institute to DFM.

Conflict of interest and fundingThe author has not received any funding or benefits from

industry or elsewhere to conduct this study.

References

1. Autism and Developmental Disabilities Monitoring Network

Surveillance Year 2008 Principal Investigators. Prevalence of

autism spectrum disorders � autism and developmental dis-

abilities monitoring network, 14 sites, United States, 2008.

MMWR Surveill Summ 2012; 61: 1�19.

2. Cody H, Pelphrey K, Piven J. Structural and functional

magnetic resonance imaging of autism. Int J Dev Neurosci

2002; 20: 421�38.

3. Courchesne E, Pierce K. Brain overgrowth in autism during a

critical time in development: implications for frontal pyramidal

neuron and interneuron development and connectivity. Int J

Dev Neurosci 2005; 23: 153�70.

4. Bauman ML, Kemper TL. Neuroanatomic observations of

the brain in autism: a review and future directions. Int J Dev

Neurosci 2005; 23: 183�7.

5. Connolly AM, Chez M, Streif EM, Keeling RM, Golumbek

PT, Kwon JM, et al. Brain-derived neurotrophic factor and

autoantibodies to neural antigens in sera of children with

autistic spectrum disorders, Landau-Kleffner syndrome, and

epilepsy. Biol Psychiatry 2005; 59: 354�63.

6. Araghi-Niknam M, Fatemi SH. Levels of Bcl-2 and P53 are

altered in superior frontal and cerebellar cortices of autistic

subjects. Cell Mol Neurobiol 2003; 23: 945�52.

7. Palomo T, Beninger RJ, Kostrzewa RM, Archer T. Brain sites

of movement disorder: genetic and environmental agents

in neurodevelopmental perturbations. Neurotox Res 2003; 5:

1�26.

8. Raymond GV, Bauman ML, Kemper TL. Hippocampus in

autism: a Golgi analysis. Acta Neuropathol (Berl) 1996; 91:

117�9.

9. Bittman K, Becker DL, Cicirata F, Parnavelas JG. Connexin

expression in homotypic and heterotypic cell coupling in

the developing cerebral cortex. J Comp Neurol 2002; 443:

201�12.

10. Kirchhoff F, Dringen R, Giaume C. Pathways of neuron-

astrocyte interactions and their possible role in neuroprotec-

tion. Eur Arch Psychiatry Clin Neurosci 2001; 251: 159�69.

11. Carper RA, Moses P, Tigue ZD, Courchesne E. Cerebral

lobes in autism: early hyperplasia and abnormal age effects.

Neuroimage 2002; 16: 1038�51.

12. Hendry J, Devito T, Gelman N, Densmore M, Rajakumar N,

Pavlosky W, et al. White matter abnormalities in autism

detected through transverse relaxation time imaging. Neuro-

image 2006; 29: 1049�57.

13. Vargas DL, Nascimbene C, Krishnan C, Zimmerman AW,

Pardo CA. Neuroglial activation and neuroinflammation in the

brain of patients with autism. Ann Neurol 2005; 57: 67�81.

14. Molloy CA, Morrow AL, Meinzen-Derr J, Schleifer K,

Dienger K, Manning-Court P, et al. Elevated cytokine levels

in children with autism spectrum disorder. J Neuroimmunol

2005; 172: 198�205.

15. Ameis SH, Szatmari P. Imaging-genetics in autism spectrum

disorder: advances, translational impact, and future directions.

Front Psychiatry 2012; 3: 46.

16. Hu VW, Frank BC, Heine S, Lee NH, Quackenbush J. Gene

expression profiling of lymphoblastoid cell lines from mono-

zygotic twins discordant in severity of autism reveals differ-

ential regulation of neurologically relevant genes. BMC

Genomics 2006; 7: 118.

17. Jyonouchi H, Sun S, Itokazu N. Innate immunity associated

with inflammatory responses and cytokine production against

common dietary proteins in patients with autism spectrum

disorder. Neuropsychobiology 2002; 46: 76�84.

18. Arndt TL, Stodgell CJ, Rodier PM. The teratology of autism.

Int J Dev Neurosci 2005; 23: 189�99.

19. Herbert MR. Contributions of the environment and environ-

mentally vulnerable physiology to autism spectrum disorders.

Curr Opin Neurol 2010; 23: 103�10.

20. Rossignol DA, Frye RE. A review of research trends in

physiological abnormalities in autism spectrum disorders:

immune dysregulation, inflammation, oxidative stress, mito-

chondrial dysfunction and environmental toxicant exposures.

Mol Psychiatry 2012; 17: 389�401.

21. Frye RE, Rossignol DA. Mitochondrial dysfunction can

connect the diverse medical symptoms associated with autism

spectrum disorders. Pediatr Res 2011; 69: 41R�7R.

22. Rossignol DA, Frye RE. Mitochondrial dysfunction in autism

spectrum disorders: a systematic review and meta-analysis.

Mol Psychiatry 2011; 17: 290�314.

23. Campbell DB, Buie TM, Winter H, Bauman M, Sutcliffe JS,

Perrin JM, et al. Distinct genetic risk based on association of

MET in families with co-occurring autism and gastrointestinal

conditions. Pediatrics 2009; 123: 1018�24.

24. Lintas C, Sacco R, Garbett K, Mirnics K, Militerni R,

Bravaccio C, et al. Involvement of the PRKCB1 gene in

autistic disorder: significant genetic association and reduced

neocortical gene expression. Mol Psychiatry 2008; 14: 705�18.

25. Celestino-Soper PB, Violante S, Crawford EL, Luo R, Lionel

AC, Delaby E, et al. A common X-linked inborn error of

carnitine biosynthesis may be a risk factor for nondysmorphic

autism. Proc Natl Acad Sci USA 2012; 109: 7974�81.

26. Bayona-Bafaluy MP, Muller S, Moraes CT. Fast adaptive

coevolution of nuclear and mitochondrial subunits of ATP

synthetase in orangutan. Mol Biol Evol 2005; 22: 716�24.

27. Nishigaki Y, Fuku N, Tanaka M. Mitochondrial haplogroups

associated with lifestyle-related diseases and longevity in the

Japanese population. Geriatr Gerontol Int 2010; 10(Suppl 1):

S221�35.

28. Weissman JR, Kelley RI, Bauman ML, Cohen BH, Murray

KF, Mitchell RL, et al. Mitochondrial disease in autism

spectrum disorder patients: a cohort analysis. PLoS One

2008; 3: e3815.

29. Giulivi C, Zhang YF, Omanska-Klusek A, Ross-Inta C, Wong

S, Hertz-Picciotto I, et al. Mitochondrial dysfunction in

autism. JAMA 2010; 304: 2389�96.

30. Vali S, Mythri RB, Jagatha B, Padiadpu J, Ramanujan KS,

Andersen JK, et al. Integrating glutathione metabolism and

mitochondrial dysfunction with implications for Parkinson’s

disease: a dynamic model. Neuroscience 2007; 149: 917�30.

31. Samavati L, Lee I, Mathes I, Lottspeich F, Huttemann M.

Tumor necrosis factor alpha inhibits oxidative phosphorylation

through tyrosine phosphorylation at subunit I of cytochrome

c oxidase. J Biol Chem 2008; 283: 21134�44.

32. Vempati UD, Diaz F, Barrientos A, Narisawa S, Mian AM,

Millan JL, et al. Role of cytochrome C in apoptosis: increased

Derrick F. MacFabe

18(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 36: Derrick MacFabe M.D.

sensitivity to tumor necrosis factor alpha is associated with

respiratory defects but not with lack of cytochrome C release.

Mol Cell Biol 2007; 27: 1771�83.

33. Suematsu N, Tsutsui H, Wen J, Kang D, Ikeuchi M, Ide T,

et al. Oxidative stress mediates tumor necrosis factor-alpha-

induced mitochondrial DNA damage and dysfunction in

cardiac myocytes. Circulation 2003; 107: 1418�23.

34. Hallmayer J, Cleveland S, Torres A, Phillips J, Cohen B,

Torigoe T, et al. Genetic heritability and shared environmental

factors among twin pairs with autism. Arch Gen Psychiatry

2011; 68: 1095�102.

35. Herbert MR, Russo JP, Yang S, Roohi J, Blaxill M, Kahler SG,

et al. Autism and environmental genomics. Neurotoxicology

2006; 27: 671�84.

36. Ashwood P, Anthony A, Torrente F, Wakefield AJ. Sponta-

neous mucosal lymphocyte cytokine profiles in children with

autism and gastrointestinal symptoms: mucosal immune acti-

vation and reduced counter regulatory interleukin-10. J Clin

Immunol 2004; 24: 664�73.

37. Fowler BA, Woods JS. Ultrastructural and biochemical

changes in renal mitochondria during chronic oral methyl

mercury exposure: the relationship to renal function. Exp Mol

Pathol 1977; 27: 403�12.

38. Shenker BJ, Guo TL, O I, Shapiro IM. Induction of apoptosis

in human T-cells by methyl mercury: temporal relationship

between mitochondrial dysfunction and loss of reductive

reserve. Toxicol Appl Pharmacol 1999; 157: 23�35.

39. Goyer RA. Toxic and essential metal interactions. Annu Rev

Nutr 1997; 17: 37�50.

40. Pourahmad J, Mihajlovic A, O’Brien PJ. Hepatocyte lysis

induced by environmental metal toxins may involve apoptotic

death signals initiated by mitochondrial injury. Adv Exp Med

Biol 2001; 500: 249�52.

41. Hiura TS, Li N, Kaplan R, Horwitz M, Seagrave JC, Nel AE.

The role of a mitochondrial pathway in the induction of

apoptosis by chemicals extracted from diesel exhaust particles.

J Immunol 2000; 165: 2703�11.

42. Wong PW, Garcia EF, Pessah IN. Ortho-substituted PCB95

alters intracellular calcium signaling and causes cellular

acidification in PC12 cells by an immunophilin-dependent

mechanism. J Neurochem 2001; 76: 450�63.

43. Sherer TB, Richardson JR, Testa CM, Seo BB, Panov AV, Yagi

T, et al. Mechanism of toxicity of pesticides acting at complex

I: relevance to environmental etiologies of Parkinson’s disease.

J Neurochem 2007; 100: 1469�79.

44. Yamano T, Morita S. Effects of pesticides on isolated rat

hepatocytes, mitochondria, and microsomes II. Arch Environ

Contam Toxicol 1995; 28: 1�7.

45. Astiz M, de Alaniz MJ, Marra CA. Effect of pesticides on cell

survival in liver and brain rat tissues. Ecotoxicol Environ Saf

2009; 72: 2025�32.

46. MacFabe DF, Cain DP, Rodriguez-Capote K, Franklin AE,

Hoffman JE, Boon F, et al. Neurobiological effects of

intraventricular propionic acid in rats: possible role of short

chain fatty acids on the pathogenesis and characteristics of

autism spectrum disorders. Behav Brain Res 2007; 176: 149�69.

47. MacFabe DF, Rodriguez-Capote K, Hoffman JE, Franklin

AE, Mohammad-Asef Y, Taylor A, et al. A novel rodent model

of autism: intraventricular infusions of propionic acid increase

locomotor activity and induce neuroinflammation and oxida-

tive stress in discrete regions of adult rat brain. Am J Biochem

& Biotech 2008; 4: 146�66.

48. MacFabe DF, Cain NE, Boon F, Ossenkopp KP, Cain DP.

Effects of the enteric bacterial metabolic product propionic

acid on object-directed behavior, social behavior, cognition,

and neuroinflammation in adolescent rats: relevance to autism

spectrum disorder. Behav Brain Res 2011; 217: 47�54.

49. Shultz SR, MacFabe DF, Ossenkopp KP, Scratch S, Whelan J,

Taylor R, et al. Intracerebroventricular injection of propionic

acid, an enteric bacterial metabolic end-product, impairs social

behavior in the rat: implications for an animal model of

autism. Neuropharmacology 2008; 54: 901�11.

50. Shultz SR, MacFabe DF, Martin S, Jackson J, Taylor R, Boon

F, et al. Intracerebroventricular injections of the enteric

bacterial metabolic product propionic acid impair cognition

and sensorimotor ability in the Long-Evans rat: further

development of a rodent model of autism. Behav Brain Res

2009; 200: 33�41.

51. Thomas RH, Foley KA, Mepham JR, Tichenoff LJ, Possmayer

F, MacFabe DF. Altered brain phospholipid and acylcarnitine

profiles in propionic acid infused rodents: further development

of a potential model of autism spectrum disorders. J Neuro-

chem 2010; 113: 515�29.

52. Ossenkopp KP, Foley KA, Gibson J, Fudge MA, Kavaliers M,

Cain DP, et al. Systemic treatment with the enteric bacterial

fermentation product, propionic acid, produces both condi-

tioned taste avoidance and conditioned place avoidance in rats.

Behav Brain Res 2012; 227: 134�41.

53. Thomas RH, Meeking MM, Mepham JR, Tichenoff L,

Possmayer F, Liu S, et al. The enteric bacterial metabolite

propionic acid alters brain and plasma phospholipid molecular

species: further development of a rodent model of autism

spectrum disorders. J Neuroinflammation 2012; 9: 153.

54. Midtvedt AC, Midtvedt T. Production of short chain fatty

acids by the intestinal microflora during the first 2 years of

human life. J Pediatr Gastroenterol Nutr 1992; 15: 395�403.

55. Patterson PH. Maternal infection and autism. Brain Behav

Immun 2012; 26: 393.

56. Yap IK, Angley M, Veselkov KA, Holmes E, Lindon JC,

Nicholson JK. Urinary metabolic phenotyping differentiates

children with autism from their unaffected siblings and age-

matched controls. J Proteome Res 2010; 14: 705�18.

57. Forsythe P, Sudo N, Dinan T, Taylor VH, Bienenstock J. Mood

and gut feelings. Brain Behav Immun 2010; 24: 9�16.

58. Roy CC, Kien CL, Bouthillier L, Levy E. Short-chain fatty

acids: ready for prime time? Nutr Clin Pract 2006; 21: 351�66.

59. Nicholson JK, Holmes E, Kinross J, Burcelin R, Gibson G, Jia

W, et al. Host-gut microbiota metabolic interactions. Science

2012; 336: 1262�7.

60. Human Microbiome Project Consortium. Structure, function

and diversity of the healthy human microbiome. Nature 2012;

486: 207�14.

61. Barnevik-Olsson M, Gillberg C, Fernell E. Prevalence of

autism in children of Somali origin living in Stockholm: brief

report of an at-risk population. Dev Med Child Neurol 2010;

52: 1167�8.

62. Atladottir HO, Thorsen P, Ostergaard L, Schendel DE,

Lemcke S, Abdallah M, et al. Maternal infection requiring

hospitalization during pregnancy and autism spectrum dis-

orders. J Autism Dev Disord 2010; 40: 1423�30.

63. Atladottir HO, Thorsen P, Schendel DE, Ostergaard L,

Lemcke S, Parner ET. Association of hospitalization for

infection in childhood with diagnosis of autism spectrum

disorders: a Danish cohort study. Arch Pediatr Adolesc Med

2010; 164: 470�7.

64. Finegold SM. Desulfovibrio species are potentially important

in regressive autism. Med Hypotheses 2011; 77: 270�4.

65. Finegold SM, Molitoris D, Song Y, Liu C, Vaisanen ML,

Bolte E, et al. Gastrointestinal microflora studies in late-onset

autism. Clin Infect Dis 2002; 35: S6�16.

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 19(page number not for citation purpose)

Page 37: Derrick MacFabe M.D.

66. Finegold SM, Dowd SE, Gontcharova V, Liu C, Henley KE,

Wolcott RD, et al. Pyrosequencing study of fecal microflora of

autistic and control children. Anaerobe 2010; 16: 444�53.

67. Wang L, Christophersen CT, Sorich MJ, Gerber JP, Angley

MT, Conlon MA. Elevated fecal short chain fatty acid and

ammonia concentrations in children with autism spectrum

disorder. Dig Dis Sci 2012; 57: 2096�102.

68. Jan G, Belzacq AS, Haouzi D, Rouault A, Metivier D,

Kroemer G, et al. Propionibacteria induce apoptosis of color-

ectal carcinoma cells via short-chain fatty acids acting on

mitochondria. Cell Death Differ 2002; 9: 179�88.

69. Al-Lahham SH, Peppelenbosch MP, Roelofsen H, Vonk RJ,

Venema K. Biological effects of propionic acid in humans;

metabolism, potential applications and underlying mechan-

isms. Biochim Biophys Acta 2010; 1801: 1175�83.

70. Zouboulis CC, Eady A, Philpott M, Goldsmith LA, Orfanos

C, Cunliffe WC, et al. What is the pathogenesis of acne? Exp

Dermatol 2005; 14: 143�52.

71. Borgstrom MK, Edwardsson S, Svensater G, Twetman S. Acid

formation in sucrose-exposed dental plaque in relation to

caries incidence in schoolchildren. Clin Oral Investig 2000; 4:

9�12.

72. Niederman R, Zhang J, Kashket S. Short-chain carboxylic-

acid-stimulated, PMN-mediated gingival inflammation. Crit

Rev Oral Biol Med 1997; 8: 269�90.

73. Zarate G, Gonzalez S, Chaia AP. Assessing survival of

dairy propionibacteria in gastrointestinal conditions and

adherence to intestinal epithelia. Methods Mol Biol 2004;

268: 423�32.

74. Brennan PA, Kendrick KM. Mammalian social odours:

attraction and individual recognition. Philos Trans R Soc

Lond B Biol Sci 2006; 361: 2061�78.

75. Yonezawa T, Haga S, Kobayashi Y, Katoh K, Obara Y. Short-

chain fatty acid signaling pathways in bovine mammary

epithelial cells. Regul Pept 2009; 153: 30�6.

76. Brock M, Buckel W. On the mechanism of action of the

antifungal agent propionate. Eur J Biochem 2004; 271: 3227�41.

77. Karuri AR, Dobrowsky E, Tannock IF. Selective cellular

acidification and toxicity of weak organic acids in an acidic

microenvironment. Br J Cancer 1993; 68: 1080�7.

78. Tamai I, Takanaga H, Maeda H, Sai Y, Ogihara T, Higashida

H, et al. Participation of a proton-cotransporter, MCT1, in the

intestinal transport of monocarboxylic acids. Biochem Biophys

Res Commun 1995; 214: 482�9.

79. Bergersen L, Rafiki A, Ottersen OP. Immunogold cytochem-

istry identifies specialized membrane domains for monocar-

boxylate transport in the central nervous system. Neurochem

Res 2002; 27: 89�96.

80. Conn AR, Fell DI, Steele RD. Characterization of alpha-keto

acid transport across blood-brain barrier in rats. Am J Physiol

1983; 245: E253�60.

81. Peinado A, Yuste R, Katz LC. Extensive dye coupling between

rat neocortical neurons during the period of circuit formation.

Neuron 1993; 10: 103�14.

82. Maurer MH, Canis M, Kuschinsky W, Duelli R. Correlation

between local monocarboxylate transporter 1 (MCT1) and

glucose transporter 1 (GLUT1) densities in the adult rat brain.

Neurosci Lett 2004; 355: 105�8.

83. Rafiki A, Boulland JL, Halestrap AP, Ottersen OP, Bergersen

L. Highly differential expression of the monocarboxylate

transporters MCT2 and MCT4 in the developing rat brain.

Neuroscience 2003; 122: 677�88.

84. Nakao S, Moriya Y, Furuyama S, Niederman R, Sugiya H.

Propionic acid stimulates superoxide generation in human

neutrophils. Cell Biol Int 1998; 22: 331�7.

85. DeCastro M, Nankova BB, Shah P, Patel P, Mally PV, Mishra

R, et al. Short chain fatty acids regulate tyrosine hydroxylase

gene expression through a cAMP-dependent signaling path-

way. Brain Res Mol Brain Res 2005; 142: 28�38.

86. Wajner M, Latini A, Wyse AT, Dutra-Filho CS. The role of

oxidative damage in the neuropathology of organic acidurias:

insights from animal studies. J Inherit Metab Dis 2004; 27:

427�48.

87. Hara H, Haga S, Aoyama Y, Kiriyama S. Short-chain fatty

acids suppress cholesterol synthesis in rat liver and intestine.

J Nutr 1999; 129: 942�8.

88. Le Poul E, Loison C, Struyf S, Springael JY, Lannoy V,

Decobecq ME, et al. Functional characterization of human

receptors for short chain fatty acids and their role in

polymorphonuclear cell activation. J Biol Chem 2003; 278:

25481�9.

89. Rorig B, Klausa G, Sutor B. Intracellular acidification reduced

gap junction coupling between immature rat neocortical

pyramidal neurones. J Physiol 1996; 490(pt-1): 31�49.

90. Parab S, Nankova BB, La Gamma EF. Differential regulation

of the tyrosine hydroxylase and enkephalin neuropeptide

transmitter genes in rat PC12 cells by short chain fatty acids:

concentration-dependent effects on transcription and RNA

stability. Brain Res 2007; 1132: 42�50.

91. Schreiber J, Chapman KA, Summar ML, Ah MN, Sutton VR,

MacLeod E, et al. Neurologic considerations in propionic

acidemia. Mol Genet Metab 2012; 105: 10�5.

92. Nyhan WL, Bay C, Beyer EW, Mazi M. Neurologic nonmeta-

bolic presentation of propionic acidemia. Arch Neurol 1999;

56: 1143�7.

93. Al-Owain M, Kaya N, Al-Shamrani H, Al-Bakheet A, Qari A,

Al-Muaigl S, et al. Autism spectrum disorder in a child with

propionic acidemia. J Inherit Metab Dis 2012 (in press).

94. Perez B, Desviat LR, Rodriguez-Pombo P, Clavero S,

Navarrete R, Perez-Cerda C, et al. Propionic acidemia:

identification of twenty-four novel mutations in Europe and

North America. Mol Genet Metab 2003; 78: 59�67.

95. Yorifuji T, Kawai M, Muroi J, Mamada M, Kurokawa K,

Shigematsu Y, et al. Unexpectedly high prevalence of the mild

form of propionic acidemia in Japan: presence of a common

mutation and possible clinical implications. Hum Genet 2002;

111: 161�5.

96. Desviat LR, Perez B, Perez-Cerda C, Rodriguez-Pombo P,

Clavero S, Ugarte M. Propionic acidemia: mutation update

and functional and structural effects of the variant alleles. Mol

Genet Metab 2004; 83: 28�37.

97. Haska L, Andersson R, Nyman M. The effect of dietary fiber

from wheat processing streams on the formation of carboxylic

acids and microbiota in the hindgut of rats. J Agric Food Chem

2011; 59: 3406�13.

98. Stackebrandt E, Rainey FA. Phylogenetic relationships. In:

Rood JI, McClane BA, Songer JG, Titball RW, eds. The

clostridia, molecular biology and pathogenesis. New York, NY:

Academic Press; 1997. p. 3�19.

99. Song Y, Liu C, Finegold SM. Real-time PCR quantitation of

clostridia in feces of autistic children. Appl Environ Microbiol

2004; 70: 6459�65.

100. Finegold SM. State of the art; microbiology in health and

disease. Intestinal bacterial flora in autism. Anaerobe 2011; 17:

367�8.

101. Barton LL, Fauque GD. Biochemistry, physiology and bio-

technology of sulfate-reducing bacteria. Adv Appl Microbiol

2009; 68: 41�98.

102. Lagoutte E, Mimoun S, Andriamihaja M, Chaumontet C,

Blachier F, Bouillaud F. Oxidation of hydrogen sulfide remains

a priority in mammalian cells and causes reverse electron

Derrick F. MacFabe

20(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 38: Derrick MacFabe M.D.

transfer in colonocytes. Biochim Biophys Acta 2010; 1797:

1500�11.

103. Gadalla MM, Snyder SH. Hydrogen sulfide as a gasotrans-

mitter. J Neurochem 2010; 113: 14�26.

104. White JF. Intestinal pathophysiology in autism. Exp Biol Med

2003; 228: 639�49.

105. Williams BL, Hornig M, Buie T, Bauman ML, Cho PM,

Wick I, et al. Impaired carbohydrate digestion and transport

and mucosal dysbiosis in the intestines of children with

autism and gastrointestinal disturbances. PLoS One 2011; 6:

e24585.

106. Horvath K, Papadimitriou JC, Rabsztyn A, Drachenberg C,

Tildon JT. Gastrointestinal abnormalities in children with

autistic disorder. J Pediatr 1999; 135: 559�63.

107. Liu Z, Li N, Neu J. Tight junctions, leaky intestines, and

pediatric diseases. Acta Paediatr 2005; 94: 386�93.

108. Nafday SM, Chen W, Peng L, Babyatsky MW, Holzman IR,

Lin J. Short-chain fatty acids induce colonic mucosal injury

in rats with various postnatal ages. Pediatr Res 2005; 57:

201�4.

109. Mellon AF, Deshpande SA, Mathers JC, Bartlett K. Effect of

oral antibiotics on intestinal production of propionic acid.

Arch Dis Child 2000; 82: 169�72.

110. Chauhan A, Chauhan V. Oxidative stress in autism. Patho-

physiology 2006; 13: 171�81.

111. James SJ, Rose S, Melnyk S, Jernigan S, Blossom S, Pavliv O,

et al. Cellular and mitochondrial glutathione redox imbalance

in lymphoblastoid cells derived from children with autism.

FASEB J 2009; 23: 2374�83.

112. Moy SS, Nadler JJ, Magnuson TR, Crawley JN. Mouse models

of autism spectrum disorders: the challenge for behavioral

genetics. Am J Med Genet C Semin Med Genet 2006; 142:

40�51.

113. Crawley JN. Mouse behavioral assays relevant to the symp-

toms of autism. Brain Pathol 2007; 17: 448�59.

114. Hornig M, Chian D, Lipkin WI. Neurotoxic effects of

postnatal thimerosal are mouse strain dependent. Mol

Psychiatry 2004; 9: 833�45.

115. Zerrate MC, Pletnikov M, Connors SL, Vargas DL, Seidler FJ,

Zimmerman AW, et al. Neuroinflammation and behavioral

abnormalities after neonatal terbutaline treatment in rats:

implications for autism. J Pharmacol Exp Ther 2007; 322:

16�22.

116. Lancaster K, Dietz DM, Moran TH, Pletnikov MV. Abnormal

social behaviors in young and adult rats neonatally infected

with Borna disease virus. Behav Brain Res 2006; 176: 141�8.

117. Smith SE, Li J, Garbett K, Mirnics K, Patterson PH. Maternal

immune activation alters fetal brain development through

interleukin-6. J Neurosci 2007; 27: 10695�702.

118. Zwaigenbaum L, Bryson S, Rogers T, Roberts W, Brian J,

Szatmari P. Behavioral manifestations of autism in the first

year of life. Int J Dev Neurosci 2005; 23: 143�52.

119. DiCicco-Bloom E, Lord C, Zwaigenbaum L, Courchesne E,

Dager SR, Schmitz C, et al. The developmental neurobiology

of autism spectrum disorder. J Neurosci 2006; 26: 6897�906.

120. Crawley JN. Designing mouse behavioral tasks relevant to

autistic-like behaviors. Ment Retard Dev Disabil Res Rev 2004;

10: 248�58.

121. Rinehart NJ, Tonge BJ, Iansek R, McGinley J, Brereton AV,

Enticott PG, et al. Gait function in newly diagnosed children

with autism: cerebellar and basal ganglia related motor

disorder. Dev Med Child Neurol 2006; 48: 819�24.

122. Feliz B, Witt DR, Harris BT. Propionic acidemia: a neuro-

pathology case report and review of prior cases. Arch Pathol

Lab Med 2003; 127: e325�8.

123. Besag FM. Behavioral aspects of pediatric epilepsy syndromes.

Epilepsy Behav 2004; 5(Suppl 1): S3�13.

124. Brusque AM, Mello CF, Buchanan DN, Terracciano ST,

Rocha MP, Vargas CR, et al. Effect of chemically induced

propionic acidemia on neurobehavioral development of rats.

Pharmacol Biochem Behav 1999; 64: 529�34.

125. Cannizzaro C, Monastero R, Vacca M, Martire M. [3H]-DA

release evoked by low pH medium and internal H� accumula-

tion in rat hypothalamic synaptosomes: involvement of cal-

cium ions. Neurochem Int 2003; 43: 9�17.

126. Mitsui R, Ono S, Karaki S, Kuwahara A. Neural and non-

neural mediation of propionate-induced contractile responses

in the rat distal colon. Neurogastroenterol Motil 2005; 17:

585�94.

127. Moore H, Grace AA. A role for electrotonic coupling in the

striatum in the expression of dopamine receptor-mediated

stereotypies. Neuropsychopharmacology 2002; 27: 980�92.

128. Sziray N, Leveleki C, Levay G, Marko B, Harsing LG, Jr.,

Mikics E, et al. Mechanisms underlying the long-term beha-

vioral effects of traumatic experience in rats: the role of

serotonin/noradrenaline balance and NMDA receptors. Brain

Res Bull 2007; 71: 376�85.

129. Neuhaus E, Beauchaine TP, Bernier R. Neurobiological

correlates of social functioning in autism. Clin Psychol Rev

2010; 30: 733�48.

130. El-Ansary AK, Ben BA, Kotb M. Etiology of autistic features:

the persisting neurotoxic effects of propionic acid. J Neuroin-

flammation 2012; 9: 74.

131. Bonnet U, Bingmann D, Wiemann M. Intracellular pH

modulates spontaneous and epileptiform bioelectric activity

of hippocampal CA3-neurones. Eur Neuropsychopharmacol

2000; 10: 97�103.

132. Severson CA, Wang W, Pieribone VA, Dohle CI, Richerson

GB. Midbrain serotonergic neurons are central pH chemor-

eceptors. Nat Neurosci 2003; 6: 1139�40.

133. Shah P, Nankova BB, Parab S, La Gamma EF. Short chain

fatty acids induce TH gene expression via ERK-dependent

phosphorylation of CREB protein. Brain Res 2006; 1107:

13�23.

134. Ming X, Julu PO, Brimacombe M, Connor S, Daniels ML.

Reduced cardiac parasympathetic activity in children with

autism. Brain Dev 2005; 27: 509�16.

135. Rigo FK, Pasquetti L, Malfatti CR, Fighera MR, Coelho RC,

Petri CZ, et al. Propionic acid induces convulsions and protein

carbonylation in rats. Neurosci Lett 2006; 408: 151�4.

136. Gogolla N, Leblanc JJ, Quast KB, Sudhof T, Fagiolini M,

Hensch TK. Common circuit defect of excitatory-inhibitory

balance in mouse models of autism. J Neurodev Disord 2009;

1: 172�81.

137. de Almeida LM, Funchal C, Pelaez PL, Pessutto FD, Loureiro

SO, Vivian L, et al. Effect of propionic and methylmalonic

acids on the in vitro phosphorylation of intermediate filaments

from cerebral cortex of rats during development. Metab Brain

Dis 2003; 18: 207�19.

138. Clavero S, Perez B, Rincon A, Ugarte M, Desviat LR.

Qualitative and quantitative analysis of the effect of splicing

mutations in propionic acidemia underlying non-severe phe-

notypes. Hum Genet 2004; 115: 239�47.

139. Hanstock TL, Clayton EH, Li KM, Mallet PE. Anxiety and

aggression associated with the fermentation of carbohydrates

in the hindgut of rats. Physiol Behav 2004; 82: 357�68.

140. Puwanant M, Mo-Suwan L, Patrapinyokul S. Recurrent

D-lactic acidosis in a child with short bowel syndrome. Asia

Pac J Clin Nutr 2005; 14: 195�8.

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 21(page number not for citation purpose)

Page 39: Derrick MacFabe M.D.

141. Brass EP, Beyerinck RA. Effects of propionate and carnitine

on the hepatic oxidation of short- and medium-chain-length

fatty acids. J Biochem 1988; 250: 819�25.

142. Filipek PA, Juranek J, Nguyen MT, Cummings C, Gargus JJ.

Relative carnitine deficiency in autism. J Autism Dev Disord

2004; 34: 615�23.

143. Kim KC, Kim P, Go HS, Choi CS, Yang SI, Cheong JH, et al.

The critical period of valproate exposure to induce autistic

symptoms in Sprague-Dawley rats. Toxicol Lett 2011; 201:

137�42.

144. Coulter DL. Carnitine, valproate, and toxicity. J Child Neurol

1991; 6: 7�14.

145. Schulpis KH, Karikas GA, Tjamouranis J, Regoutas S,

Tsakiris S. Low serum biotinidase activity in children with

valproic acid monotherapy. Epilepsia 2001; 42: 1359�62.

146. Borlongan CV, Koutouzis TK, Sanberg PR. 3-Nitropropionic

acid animal model and Huntington’s disease. Neurosci Biobe-

hav Rev 1997; 21: 289�93.

147. Chauhan A, Chauhan V, Brown WT, Cohen I. Oxidative stress

in autism: increased lipid peroxidation and reduced serum

levels of ceruloplasmin and transferrin � the antioxidant

proteins. Life Sci 2004; 75: 2539�49.

148. Chauhan A, Audhya T, Chauhan V. Brain region-specific

glutathione redox imbalance in autism. Neurochem Res 2012;

37: 1681�9.

149. Monks TJ, Ghersi-Egea JF, Philbert M, Cooper AJ, Lock EA.

Symposium overview: the role of glutathione in neuroprotec-

tion and neurotoxicity. Toxicol Sci 1999; 51: 161�77.

150. James SJ, Cutler P, Melnyk S, Jernigan S, Janak L, Gaylor DW,

et al. Metabolic biomarkers of increased oxidative stress and

impaired methylation capacity in children with autism. Am J

Clin Nutr 2004; 80: 1611�7.

151. El-Ansary A, Al-Daihan S, Al-Dbass A, Al-Ayadhi L.

Measurement of selected ions related to oxidative stress and

energy metabolism in Saudi autistic children. Clin Biochem

2009; 43: 63�70.

152. Al-Gadani Y, El-Ansary A, Attas O, Al-Ayadhi L. Metabolic

biomarkers related to oxidative stress and antioxidant status in

Saudi autistic children. Clin Biochem 2009; 42: 1032�40.

153. Torres AR. Is fever suppression involved in the etiology of

autism and neurodevelopmental disorders? BMC Pediatr 2003;

3: 9.

154. Neustadt J, Pieczenik SR. Medication-induced mitochondrial

damage and disease. Mol Nutr Food Res 2008; 52: 780�8.

155. Schultz ST, Klonoff-Cohen HS, Wingard DL, Akshoomoff

NA, Macera CA, Ji M. Acetaminophen (paracetamol) use,

measles-mumps-rubella vaccination, and autistic disorder: the

results of a parent survey. Autism 2008; 12: 293�307.

156. Noble M, Proschel C, Mayer-Proschel M. Oxidative-

reductionist approaches to stem and progenitor cell function.

Cell Stem Cell 2011; 8: 1�2.

157. Rorig B, Sutor B. Serotonin regulates gap junction coupling in

the developing rat somatosensory cortex. Eur J Neurosci 1996;

8: 1685�95.

158. Anders JJ. Lactic acid inhibition of gap junctional intercellular

communication in in vitro astrocytes as measured by fluores-

cence recovery after laser photobleaching. GLIA 1988; 1:

371�9.

159. Frantseva MV, Kokarovtseva L, Naus CG, Carlen PL,

MacFabe D, Perez Velazquez JL. Specific gap junctions

enhance the neuronal vulnerability to brain traumatic injury.

J Neurosci 2002; 22: 644�53.

160. Perez-Velazquez JL, Frantseva MV, Naus CC. Gap junctions

and neuronal injury: protectants or executioners? Neuroscien-

tist 2003; 9: 5�9.

161. Juszczak GR, Swiergiel AH. Properties of gap junction

blockers and their behavioural, cognitive and electrophysiolo-

gical effects: animal and human studies. Prog Neuropsycho-

pharmacol Biol Psychiatry 2009; 33: 181�98.

162. Wiencken-Barger AE, Djukic B, Casper KB, McCarthy KD.

A role for Connexin43 during neurodevelopment. GLIA 2007;

55: 675�86.

163. Nagasawa K, Chiba H, Fujita H, Kojima T, Saito T, Endo T,

et al. Possible involvement of gap junctions in the barrier

function of tight junctions of brain and lung endothelial cells. J

Cell Physiol 2006; 208: 123�32.

164. Pardo CA, Eberhart CG. The neurobiology of autism. Brain

Pathol 2007; 17: 434�47.

165. Madhusudan A, Vogel P, Knuesel I. Impact of prenatal

immune system disturbances on brain development. J Neu-

roimmune Pharmacol, in press.

166. Dringen R. Oxidative and antioxidative potential of brain

microglial cells. Antioxid Redox Signal 2005; 7: 1223�33.

167. Carlezon WA, Jr., Duman RS, Nestler EJ. The many faces of

CREB. Trends Neurosci 2005; 28: 436�45.

168. Wang Z, Xu L, Zhu X, Cui W, Sun Y, Nishijo H, et al.

Demethylation of specific Wnt/beta-catenin pathway genes

and its upregulation in rat brain induced by prenatal valproate

exposure. Anat Rec (Hoboken) 2010; 293: 1947�53.

169. Nankova BB, La Gamma EF, Taylor AR, Tichenoff L,

MacFabe DF. Intraventricular enteric short chain fatty

acid infusions in rats induce behavioural, neuropathological,

lipid and epigenetic changes consistent with Autism. abstr.

International Meeting for Autism Research, May 17�19, 2012,

Toronto.

170. Bell JG, Sargent JR, Tocher DR, Dick JR. Red blood cell fatty

acid compositions in a patient with autistic spectrum disorder:

a characteristic abnormality in neurodevelopmental disorders?

Prostaglandins Leukot Essent Fatty Acids 2000; 63: 21�5.

171. Vancassel S, Durand G, Barthelemy C, Lejeune B, Martineau

J, Guilloteau D, et al. Plasma fatty acid levels in autistic

children. Prostaglandins Leukot Essent Fatty Acids 2001; 65:

1�7.

172. Richardson AJ. Clinical trials of fatty acid treatment in

ADHD, dyslexia, dyspraxia and the autistic spectrum. Pros-

taglandins Leukot Essent Fatty Acids 2004; 70: 383�90.

173. Amminger GP, Berger GE, Schafer MR, Klier C, Friedrich

MH, Feucht M. Omega-3 fatty acids supplementation in

children with autism: a double-blind randomized, placebo-

controlled pilot study. Biol Psychiatry 2007; 61: 551�3.

174. Wiest MM, German JB, Harvey DJ, Watkins SM, Hertz-

Picciotto I. Plasma fatty acid profiles in autism: a case-control

study. Prostaglandins Leukot Essent Fatty Acids 2009; 80:

221�7.

175. Tamiji J, Crawford DA. The neurobiology of lipid metabolism

in autism spectrum disorders. Neurosignals 2010; 18: 98�112.

176. Bell JG, MacKinlay EE, Dick JR, MacDonald DJ, Boyle RM,

Glen AC. Essential fatty acids and phospholipase A2 in

autistic spectrum disorders. Prostaglandins Leukot Essent

Fatty Acids 2004; 71: 201�4.

177. Meguid NA, Atta HM, Gouda AS, Khalil RO. Role of

polyunsaturated fatty acids in the management of Egyptian

children with autism. Clin Biochem 2008; 41: 1044�8.

178. Aneja A, Tierney E. Autism: the role of cholesterol in

treatment. Int Rev Psychiatry 2008; 20: 165�70.

179. Champeil-Potokar G, Chaumontet C, Guesnet P, Lavialle M,

Denis I. Docosahexaenoic acid (22:6n-3) enrichment of

membrane phospholipids increases gap junction coupling

capacity in cultured astrocytes. Eur J Neurosci 2006; 24:

3084�90.

Derrick F. MacFabe

22(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 40: Derrick MacFabe M.D.

180. Zhao S, Jia L, Gao P, Li Q, Lu X, Li J, et al. Study on the effect

of eicosapentaenoic acid on phospholipids composition in

membrane microdomains of tight junctions of epithelial cells

by liquid chromatography/electrospray mass spectrometry.

J Pharm Biomed Anal 2008; 47: 343�50.

181. Jones LL, McDonald DA, Borum PR. Acylcarnitines: role in

brain. Prog Lipid Res 2010; 49: 61�75.

182. Clarke JT, Clark-Taylor BE. Is autism a disorder of fatty

acid metabolism? Possible dysfunction of mitochondrial

beta oxidation by long chain acyl-CoA dehydrogenase. Med

Hypotheses 2004; 62: 970�5.

183. Frye RE. Biomarkers of abnormal energy metabolism in

children with autism spectrum disorder. N Am J Med Sci,

in press.

184. Pastural E, Ritchie S, Lu Y, Jin W, Kavianpour A, Khine Su-

Myat K, et al. Novel plasma phospholipid biomarkers of

autism: mitochondrial dysfunction as a putative causative

mechanism. Prostaglandins Leukot Essent Fatty Acids 2009;

81: 253�64.

185. Bu B, Ashwood P, Harvey D, King IB, Water JV, Jin LW. Fatty

acid compositions of red blood cell phospholipids in children

with autism. Prostaglandins Leukot Essent Fatty Acids 2006;

74: 215�21.

186. Brass EP, Fennessey PV, Miller LV. Inhibition of oxidative

metabolism by propionic acid and its reversal by carnitine in

isolated rat hepatocytes. Biochem J 1986; 236: 131�6.

187. Roe RC, Millington DS, Maltby DA, Bohan TP, Hoppel CL.

L-carnitine enhances excretion of propoinyl coenzyme A as

propionylcarnitine in propoinic acidemia. J Clin Invest 1984;

73: 1785�8.

188. Tonin AM, Grings M, Knebel LA, Zanatta A, Moura AP,

Ribeiro CA, et al. Disruption of redox homeostasis in cerebral

cortex of developing rats by acylcarnitines accumulating in

medium-chain acyl-CoA dehydrogenase deficiency. Int J Dev

Neurosci 2012; 30: 383�90.

189. Schwab MA, Sauer SW, Okun JG, Nijtmans LG, Rodenburg

RJ, van den Heuvel LP, et al. Secondary mitochondrial

dysfunction in propionic aciduria: a pathogenic role for

endogenous mitochondrial toxins. Biochem J 2006; 398:

107�12.

190. Brass EP. Interaction of carnitine and propionate with

pyruvate oxidation by hepatocytes from clofibrate-treated

rats: importance of coenzyme A availability. J Nutr 1992;

122: 234�40.

191. Wolever TM, Fernandes J, Rao AV. Serum acetate: propionate

ratio is related to serum cholesterol in men but not women. J

Nutr 1996; 126: 2790�7.

192. Wolever TM, Spadafora PJ, Cunnane SC, Pencharz PB.

Propionate inhibits incorporation of colonic [1,2-13C]acetate

into plasma lipids in humans. Am J Clin Nutr 1995; 61:

1241�7.

193. James SJ, Melnyk S, Jernigan S, Cleves MA, Halsted CH,

Wong DH, et al. Metabolic endophenotype and related

genotypes are associated with oxidative stress in children

with autism. Am J Med Genet B Neuropsychiatr Genet

2006; 141: 947�56.

194. Wegiel J, Kuchna I, Nowicki K, Imaki H, Wegiel J, Marchi E,

et al. The neuropathology of autism: defects of neurogenesis

and neuronal migration, and dysplastic changes. Acta Neuro-

pathol 2010; 119: 755�70.

195. Scafidi S, Fiskum G, Lindauer SL, Bamford P, Shi D, Hopkins

I, et al. Metabolism of acetyl-L-carnitine for energy and

neurotransmitter synthesis in the immature rat brain.

J Neurochem 2010; 114: 820�31.

196. James SJ, Melnyk S, Fuchs G, Reid T, Jernigan S, Pavliv O,

et al. Efficacy of methylcobalamin and folinic acid treat-

ment on glutathione redox status in children with autism.

Am J Clin Nutr 2009; 89: 425�30.

197. Miecz D, Januszewicz E, Czeredys M, Hinton BT, Berezowski

V, Cecchelli R, et al. Localization of organic cation/carnitine

transporter (OCTN2) in cells forming the blood-brain barrier.

J Neurochem 2008; 104: 113�23.

198. Yamamoto-Furusho JK, Mendivil-Rangel EJ, Villeda-Ramirez

MA, Fonseca-Camarillo G, Barreto-Zuniga R. Gene expres-

sion of carnitine organic cation transporters 1 and 2 (OCTN) is

downregulated in patients with ulcerative colitis. Inflamm

Bowel Dis 2011; 17: 2205�6.

199. Yao D, Kuwajima M, Chen Y, Shiota M, Okumura Y, Yamada

H, et al. Impaired long-chain fatty acid metabolism in

mitochondria causes brain vascular invasion by a non-

neurotropic epidemic influenza A virus in the newborn/suckling

period: implications for influenza-associated encephalopathy.

Mol Cell Biochem 2007; 299: 85�92.

200. Fallon J. Could one of the most widely prescribed antibiotics

amoxicillin/clavulanate ‘‘augmentin’’ be a risk factor for

autism? Med Hypotheses 2005; 64: 312�5.

201. Sandler RH, Finegold SM, Bolte ER, Buchanan CP, Maxwell

AP, Vaisanen ML, et al. Short-term benefit from oral

vancomycin treatment of regressive-onset autism. J Child

Neurol 2000; 15: 429�35.

202. Mann JR, McDermott S, Bao H, Hardin J, Gregg A.

Pre-eclampsia, birth weight, and autism spectrum disorders.

J Autism Dev Disord 2010; 40: 548�54.

203. Thiele IG, Niezen-Koning KE, van Gennip AH, Aarnoudse

JG. Increased plasma carnitine concentrations in preeclampsia.

Obstet Gynecol 2004; 103: 876�80.

204. Scafidi S, Racz J, Hazelton J, McKenna MC, Fiskum G.

Neuroprotection by acetyl-L-carnitine after traumatic injury to

the immature rat brain. Dev Neurosci 2010; 32: 480�7.

205. Patel SP, Sullivan PG, Lyttle TS, Rabchevsky AG. Acetyl-

L-carnitine ameliorates mitochondrial dysfunction following

contusion spinal cord injury. J Neurochem 2010; 114: 291�301.

206. Fortin G, Yurchenko K, Collette C, Rubio M, Villani AC,

Bitton A, et al. L-carnitine, a diet component and organic

cation transporter OCTN ligand, displays immunosuppressive

properties and abrogates intestinal inflammation. Clin Exp

Immunol 2009; 156: 161�71.

207. Rossignol DA, Rossignol LW, Smith S, Schneider C,

Logerquist S, Usman A, et al. Hyperbaric treatment for

children with autism: a multicenter, randomized, double-blind,

controlled trial. BMC Pediatr 2009; 9: 21.

208. Willner P. The validity of animal models of depression.

Psychopharmacology (Berl) 1984; 83: 1�16.

209. Willner P. Animal models as simulations of depression. Trends

Pharmacol Sci 1991; 12: 131�6.

210. Bresolin N, Freddo L, Vergani L, Angelini C. Carnitine,

carnitine acyltransferases, and rat brain function. Exp Neurol

1982; 78: 285�92.

211. Kavaliers M, Choleris E, Agmo A, Pfaff DW. Olfactory-

mediated parasite recognition and avoidance: linking genes to

behavior. Horm Behav 2004; 46: 272�83.

212. Kaushik M, Lamberton PH, Webster JP. The role of parasites

and pathogens in influencing generalised anxiety and predation-

related fear in the mammalian central nervous system. Horm

Behav, in press.

213. Heijtz RD, Wang S, Anuar F, Qian Y, Bjorkholm B,

Samuelsson A, et al. Normal gut microbiota modulates brain

development and behavior. Proc Natl Acad Sci USA 2011; 108:

3047�52.

214. Suren P, Bakken IJ, Aase H, Chin R, Gunnes N, Lie KK, et al.

Autism spectrum disorder, ADHD, epilepsy, and cerebral palsy

in norwegian children. Pediatrics 2012; 130: e152�8.

Gut SCFA fermentation products in autism

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260 23(page number not for citation purpose)

Page 41: Derrick MacFabe M.D.

215. Sokol MS. Infection-triggered anorexia nervosa in children:

clinical description of four cases. J Child Adolesc Psychophar-

macol 2000; 10: 133�45.

216. Kerbeshian J, Burd L. Is anorexia nervosa a neuropsychiatric

developmental disorder? An illustrative case report. World J

Biol Psychiatry 2009; 10: 648�57.

217. Diamond J. Guns, germs, and steel. New York: W.W. Norton &

Company; 1997.

218. Foley KA, Kavaliers M, Ossenkopp K-P, MacFabe DF.

Prenatal exposure to propionic acid and lipopolysaccharides

produces developmental delay, anxiety-like behaviour, and

hyper-sensitivity to accoustic startle in adolescent rats. abstr.

International Meeting for Autism Research, May17�19, 2012,

Toronto.

*Derrick F. MacFabeDirector: The Kilee Patchell-Evans Autism Research GroupDepartments of Psychology (Neuroscience) and PsychiatryDivision of Developmental DisabilitiesLawson Research InstituteRoom 9244, Social Science CentreUniversity of Western OntarioLondon, Ontario, Canada, N6A 5C2Tel: 519 661 2111 ext 84703Email: [email protected]

Derrick F. MacFabe

24(page number not for citation purpose)

Citation: Microbial Ecology in Health & Disease 2012, 23: 19260 - http://dx.doi.org/10.3402/mehd.v23i0.19260

Page 42: Derrick MacFabe M.D.

Unique acyl-carnitine profiles are potential biomarkersfor acquired mitochondrial disease in autism spectrumdisorder

RE Frye1, S Melnyk1 and DF MacFabe2

Autism spectrum disorder (ASD) has been associated with mitochondrial disease (MD). Interestingly, most individuals with ASDand MD do not have a specific genetic mutation to explain the MD, raising the possibility of that MD may be acquired, at least in asubgroup of children with ASD. Acquired MD has been demonstrated in a rodent ASD model in which propionic acid (PPA), anenteric bacterial fermentation product of ASD-associated gut bacteria, is infused intracerebroventricularly. This animal modelshows validity as it demonstrates many behavioral, metabolic, neuropathologic and neurophysiologic abnormalities associatedwith ASD. This animal model also demonstrates a unique pattern of elevations in short-chain and long-chain acyl-carnitinessuggesting abnormalities in fatty-acid metabolism. To determine if the same pattern of biomarkers of abnormal fatty-acidmetabolism are present in children with ASD, the laboratory results from a large cohort of children with ASD (n¼ 213) whounderwent screening for metabolic disorders, including mitochondrial and fatty-acid oxidation disorders, in a medically basedautism clinic were reviewed. Acyl-carnitine panels were determined to be abnormal if three or more individual acyl-carnitinespecies were abnormal in the panel and these abnormalities were verified by repeated testing. Overall, 17% of individuals withASD demonstrated consistently abnormal acyl-carnitine panels. Next, it was determined if specific acyl-carnitine species wereconsistently elevated across the individuals with consistently abnormal acyl-carnitine panels. Significant elevations in short-chainand long-chain, but not medium-chain, acyl-carnitines were found in the ASD individuals with consistently abnormal acyl-carnitinepanels—a pattern consistent with the PPA rodent ASD model. Examination of electron transport chain function in muscle andfibroblast culture, histological and electron microscopy examination of muscle and other biomarkers of mitochondrial metabolismrevealed a pattern consistent with the notion that PPA could be interfering with mitochondrial metabolism at the level of thetricarboxylic-acid cycle (TCAC). The function of the fatty-acid oxidation pathway in fibroblast cultures and biomarkers forabnormalities in non-mitochondrial fatty-acid metabolism were not consistently abnormal across the subgroup of ASD children,consistent with the notion that the abnormalities in fatty-acid metabolism found in this subgroup of children with ASD weresecondary to TCAC abnormalities. Glutathione metabolism was abnormal in the subset of ASD individuals with consistent acyl-carnitine panel abnormalities in a pattern similar to glutathione abnormalities found in the PPA rodent model of ASD. These datasuggest that there are similar pathological processes between a subset of ASD children and an animal model of ASD with acquiredmitochondrial dysfunction. Future studies need to identify additional parallels between the PPA rodent model of ASD and thissubset of ASD individuals with this unique pattern of acyl-carnitine abnormalities. A better understanding of this animal model andsubset of children with ASD should lead to better insight in mechanisms behind environmentally induced ASD pathophysiologyand should provide guidance for developing preventive and symptomatic treatments.Translational Psychiatry (2013) 3, e220; doi:10.1038/tp.2012.143; published online 22 January 2013

Introduction

Autism spectrum disorders (ASD) are a heterogeneous groupof neurodevelopmental disorders that are characterized byimpairments in social interaction and communication alongwith restrictive and repetitive behaviors.1 Many of thecognitive and behavioral features of ASD are believed toarise from central nervous system dysfunction, but abnorm-alities in many non-central nervous system tissues have beenassociated with ASD.2,3 Recent studies have implicated

abnormalities in systemic physiology that transcend organ

specific dysfunction, at least in some children with ASD.2–4

Thus, it is possible that organs other than the brain and/or

systemic abnormalities could be the source of the primary

pathophysiological that manifest, in part, with secondary brain

dysfunction.A recent meta-analysis found that 5% of children with ASD

meet criteria for a classic mitochondrial disease (MD) and

suggest that this subgroup has distinct clinical characteristics

1Department of Pediatrics, Arkansas Children’s Hospital Research Institute, Little Rock, AR, USA and 2The Kilee Patchell-Evans Autism Research Group-Departmentsof Psychology (Neuroscience) and Psychiatry, Lawson Research Institute, Schulich School of Medicine and Dentistry, University of Western Ontario, London, ON,CanadaCorrespondence: Dr RE Frye, Department of Pediatrics, Arkansas Children’s Hospital Research Institute, Slot 512-41B, 13 Children0s Way, Little Rock, AR 72202, USA.E-mail [email protected] study has not been presented or published previously.

Received 2 July 2012; revised 27 October 2012; accepted 10 November 2012Keywords: acyl-carnitines; autism spectrum disorder; clostridia; microbiome; mitochondrial disease; propionic acid

Citation: Transl Psychiatry (2013) 3, e220; doi:10.1038/tp.2012.143& 2013 Macmillan Publishers Limited All rights reserved 2158-3188/13

www.nature.com/tp

Page 43: Derrick MacFabe M.D.

that distinguish it from the general ASD population.3,5 Thisstudy also found that about 30% of the general ASDpopulation exhibited biomarkers consistent with MD.3 Thehigh prevalence of abnormal mitochondrial biomarkers in ASDhas been suggested to be due to mitochondrial dysfunctionthat is more prevalent and distinct from classic MD. Such anotion is supported by a recent study that found that 80% ofthe children with ASD demonstrated below normal function ofthe electron transport chain (ETC) in lymphocytes.6

The reason for mitochondrial dysfunction in ASD isunknown, but the fact that only 23% of children with ASDand MD have a known mitochondrial deoxyribonucleic acid(mtDNA) abnormality suggests that MD may be acquiredrather than genetic in many ASD cases.3 Indeed, some havesuggested that the systemic abnormalities in ASD such asmitochondrial dysfunction may arise from environmentaltriggers7 in genetically sensitive subpopulations.8,9 Entericshort-chain fatty-acids, such as propionic acid (PPA),10–17

which are fermentation by-products of ASD-associatedenteric bacteria (that is, Clostridia, Desulfovibrio, Sutterellaand Bacteroidetes), have been suggested as a possibleenvironmental triggers in ASD.18,19 Interestingly, humans withimpairments in PPA metabolism20–22 exhibit neurodevelop-mental conditions with ASD features.23

Recently, a rodent model has been development in whichreversible (30 min) bouts of ASD-type (that is, stereotyped,perseverative and impaired social) behaviors are produced bybrief intracerebroventricular infusions of PPA (http://www.psy-chology.uwo.ca/autism/autism6.htm for behavioral video).This animal model demonstrates several characteristics thathave been reported in ASD such as tics, electrographicseizures, innate neuroinflammation and redox, lipid, phos-phatidylethanolamine, mitochondrial, acyl-carnitine and car-nitine abnormalities.10–12,14–17 This animal model provides anunderstanding of how exogenous agents, such as PPA, cancause reversible behavioral, metabolic, neuropathologicaland neurophysiological changes associated with ASD. Mostimportantly, this animal model has predictive value as itdemonstrates biomarkers of abnormal mitochondrial fatty-acid metabolism (that is, acyl-carnitine elevations) that couldbe used as routine biomarkers if found in children with ASD.

Several lines of evidence suggest that mitochondrial fatty-acid oxidation could be abnormal in a subset of children withASD. First, free carnitine, the cofactor used to transport long-chain and very-long-chain fatty-acids into the mitochondrialmatrix, has been shown to be depleted in children with ASD.24

Free carnitine can be depleted if it remains bound tounprocessed fatty-acids due to a reduction in mitochondrialfatty-acid beta-oxidation.25 Second, elevations in long-chainand very-long-chain fatty-acids have been reported in childrenwith ASD as compared with controls, suggesting excessunprocessed fatty-acid in the serum of children with ASD.26

Third, a case study and case series of patients with ASD havereported elevations in acyl-carnitines, the standard biomarkerfor mitochondrial fatty-acid oxidation deficits.27,28 Thus, thereis ample evidence to suggest that abnormalities in fatty-acidmetabolism (that is, acyl-carnitine elevations) may be found inchildren with ASD.

Figure 1 demonstrates the acyl-carnitine elevations in brainhomogenates found in rats exposed to intracerebroventricular

infusions of PPA as compared with those exposed tophosphate buffered saline vehicle control. These abnormal-ities included short-chain (2–5 carbon length) and long-chain(13–18 carbon length) acyl-carnitines but not medium-chain(6–12 carbon length) acyl-carnitines.13 We hypothesize that asubset of children with ASD manifest biomarkers of abnormalmitochondrial fatty-acid metabolism that are similar to thosereported in the PPA rodent model of ASD. Here we review thecharts of consecutive patients seen in a medically basedautism clinic who underwent a systematic workup for mito-chondrial disorders per recently published guidelines, whichincluded screening for fatty-acid metabolism disorders.3

Overall, 17% of children with ASD were found to demonstratea unique pattern of acyl-carnitine abnormalities that weresimilar to the acyl-carnitine abnormalities found in the rodentPPA model of ASD. The potential causes of these abnormalitiesand their possible relation to ASD pathogenesis is discussed.

Materials and methods

Subject population. Parents of patients seen from 2008–2011 in a medically based autism clinic were requested toconsent to allow their child’s medical information to beanonymously abstracted into a clinical database that con-tained medical history, physical examination findings and theresults of neurological and metabolic testing. Approximately98% of parents (326 total patients) signed the consent.

Figure 1 Acyl-carnitine elevations in the brain of rats treated intracerebroven-tricularly with propionic acid. Notice that the majority of fatty-acid elevations were inshort-chain (2–5 carbon length) and long-chain (13–18 carbon length) fatty-acids ascompared with the medium-chain (6–12 carbon length) fatty-acids. This is adaptedfrom Thomas et al.13 where it was presented as a table.

Acyl-carnitine profiles in autismRE Frye et al

2

Translational Psychiatry

Page 44: Derrick MacFabe M.D.

Metabolic evaluation. A standardized metabolic workup formitochondrial metabolism disorders was conducted on mostpatients.29,30 The algorithm for this evaluation is depicted inFigure 2. Initial testing included laboratory tests to identifyabnormalities in the respiratory chain, tricarboxylic-acid cycle(TCAC) and fatty-acid oxidation pathways. Abnormalitiesdetected in initial testing were confirmed with repeat testing.If abnormalities could not be replicated, laboratories testswere reconsidered during metabolic stress or illness if a highindex of suspicion remained for the patient.

An acyl-carnitine panel, which measures short-chain,medium-chain and long-chain acyl-carnitines, was used asthe primarily laboratory test to detect defects in the fatty-acidoxidation pathway. An acyl-carnitine panel was measuredat initial testing in 213 of the consented patients. Theacyl-carnitine panel was considered abnormal if three ormore acyl-carnitines were elevated in the panel. An abnormalacyl-carnitine panel was confirmed by repeat testing.

If acyl-carnitine abnormalities were confirmed, non-mito-chondrial disorders of fatty-acid metabolism were ruled-outbefore a MD workup was initiated. Disorders ruled-outincluded generalized hyperlipidemia, hypercholesterolemia,multiple carboxylase deficiencies (that is, biotinidase defi-ciency), zinc deficiency, abnormal copper metabolism andhypoglycemia. After such disorders were ruled-out, a MDworkup was pursued. The initial step in the MD workup wasexamination for mtDNA gene abnormalities by either atargeted analysis for common mutations and/or deletions

and/or sequencing of the entire mtDNA genome (BaylorMedical Genetics Laboratory, Houston, TX, USA).31,32

When a conclusive mtDNA abnormality could not beidentified, nuclear mitochondrial gene testing and/or amuscle and/or skin biopsy was recommended. Nuclearmitochondrial gene abnormalities were ruled-out using anoligonucleotide array with comparative genomic hybridizationanalysis that examines B180 nuclear genes involved inmitochondria function, including genes involved in fatty-acid oxidation, carnitine metabolism, mitochondrial biogen-esis, mtDNA maintenance, transcription and translation, andETC complex assembly (MitoMet, Baylor Medical GeneticsLaboratory).

In some patients, the quadricep muscle was biopsied andanalyzed with light and electron microscopy, as well as formtDNA content.33,34 In some patients, fibroblasts obtainedfrom a skin biopsy were cultured. ETC function was examinedon frozen muscle and cultured fibroblasts (Baylor MedicalGenetics Laboratory).35 Both uncorrected ETC function andETC function correcting for citrate synthase are presented.Fibroblasts were incubated with d3-palmitate and L-carnitinein duplicate for 72 h to determine function of the fatty-acidoxidation pathway (Baylor Institute of Metabolic Disease,Dallas, TX).36

Determination of acyl-carnitine abnormalities. To calcu-late the prevalence of having an abnormal acyl-carnitinepanel in the ASD sample, the prevalence of having an acyl-

Non-mitochondrialdisorder confirmed

Skin and muscle biopsy Histology (muscle) Electron microscopy (muscle)Electron Transport Chain (muscle and skin)Beta-oxidation studies (skin)mtDNA content (muscle)

Targeted genetic testingMitochondrial DNA sequencing

Empiric treatment:L-CarnitineCoenzyme Q10B vitaminsAntioxidantsCreatine Monohydrate

Abnormalitiessuggest respiratory

chain disorder

mtDNAtesting

inconclusive

Normal

Morning fastingLactic acidPyruvic acidCarnitine panelAcyl-carnitine panelAmino acidsUbiquinoneAmmoniaCreatine KinaseASTALTCO2

GlucoseUrine organic acids

If high index-of-suspicionrepeat labs under metabolic

stress or illness

Repeat specificabnormal lab test(s)

Normal

Abnormal

Abnormalitiesin fatty-acidmetabolism

Rule-out non-mitochondrialfatty acid metabolic disorderswith morning fastingInsulinCholesterol panelTriglyceridesRBC Zinc RBC CopperBiotinidase enzyme activityUrine acyl-glycine

Non-mitochondrialdisordered ruled out

Specific Treatment

NuclearMitochondrial

GeneTestingNuclear

genetesting

inconclusive

Figure 2 Algorithm for metabolic workup of autistic spectrum disease patients evaluated in the medically based autism clinic. Patients are screened with biomarkers ofabnormal mitochondrial function in the fasting state. Abnormalities are verified with repeat fasting biomarker testing. For patients with biomarkers for a fatty-acid oxidationdefect, other disorders of fatty-acid metabolism are ruled-out before further workup for a mitochondrial disorder. Patients with consistent biomarkers for mitochondrialdysfunction are first investigated for genetic causes of their mitochondrial disorder before considering a muscle and/or skin biopsy. mtDNA, mitochondrial deoxyribonucleicacid; RBC, red-blood cell.

Acyl-carnitine profiles in autismRE Frye et al

3

Translational Psychiatry

Page 45: Derrick MacFabe M.D.

carnitine panel with the first laboratory test was multiplied bythe percent of patients confirmed to have an abnormal acyl-carnitine panel on repeat laboratory testing. This was done toaccount for the fact that some patients did not repeat theacyl-carnitine panel even though it was abnormal. Todetermine the specific acyl-carnitine species that wereconsistently elevated across the subgroup of patients withconsistently abnormal acyl-carnitine panels, we examinedthe first two acyl-carnitine panels measured for each patient.Values for each individual acyl-carnitine species (for eachpatient) were transformed to a percent of the upper limit ofnormal for the specific acyl-carnitine species. The mean ands.e. were then calculated for each acyl-carnitine species tosummarize the group data. Statistical significance wascalculated as the significance of the difference betweenthe upper limit of normal and the group mean for eachacyl-carnitine specific using a z-distribution derived from thegroup mean and s.e.

Measurement of glutathione metabolism. Glutathionemetabolism was evaluated in four participants. Thesepatients were compared with normative values establishedin a previous study on redox metabolism.37 These controlsincluded 42 healthy children ranged from 2–7 years of agewith no history of developmental delay or neurologicalsymptoms. Independent sample t-tests were used forcomparison. Fasting blood samples were collected intoethylenediaminetetraacetic acid vacutainer tubes and wereimmediately chilled on ice before centrifuging at 4000 g for10 min at 4 1C. To prevent metabolite inter-conversion theice-cold samples were centrifuged within 15 min of the bloodcollection and the plasma stored at � 80 1C until analysiswithin 2 weeks. Details of the methodology for high-pressureliquid chromatography with electrochemical detection andmetabolite quantitation have been previously described.38

Total and free-reduced glutathione, oxidized glutathione(GSSG) and the total-reduced glutathione/GSSG and free-reduced glutathione/GSSG ratios were measured.

Results

Prevalence and patterns of abnormal acyl-carnitines.Seventy-four (35%) of the 213 patients tested demonstratedan increase in three or more acyl-carnitines when initiallymeasured. Forty-two (57%) of the 74 underwent repeat acyl-carnitine testing. Three or more acyl-carnitines were abnor-mal a second time in 20 (48%) of the 42 patients, resulting ina prevalence of 17% of ASD children who manifestedconsistent acyl-carnitine panel abnormalities. Figure 3demonstrates the mean values of each acyl-carnitine speciesrelative of the upper limit of normal. C4OH, C14 and C16:1were significantly elevated as compared with the upper limitof normal (z¼ 2.18, P¼ 0.01; z¼ 5.71, Po0.0001; z¼ 2.85,P¼ 0.02, respectively), and were 186%, 226% and 131% ofthe upper limit of normal, respectively.

Clinical characteristics. Clinical characteristics of the 20patients with consistent elevations in the acyl-carnitine panelare given in Supplementary Table 1. The average age was8.7 years (s.d. 2.25) with a male to female ratio of 3:1.

Autistic disorder was diagnosed in 70% of the participantswhile 25% had a diagnosis of pervasive developmentaldisorder-not otherwise specified and 5% were diagnosedwith Asperger syndrome. Developmental regression wasreported in 45% of patients.

Glutathione metabolism. The subset of children in whichglutathione metabolism was examined demonstrated signifi-cantly lower total-reduced glutathione (t¼ 12.75, Po0.0001)and free-reduced glutathione (t¼ 10.04, Po0.0001) valuesand total-reduced glutathione/GSSG (t¼ 9.07, Po0.001) andfree-reduced glutathione/GSSG (t¼ 4.69, Po0.0001) ratios aswell as higher GSSG (t¼ 2.61, P¼ 0.01) values as comparedwith typically developing controls (Figure 4), suggesting both areduction in the production of gluthathione and increase ingluthathione utilization by reactive oxygen species.

Genetic characteristics. Nuclear DNA examinations werenormal in the great majority of patient (94%) in which suchexaminations were conducted. mtDNA was normal in 85% ofthe patients in which it was examined. The two mtDNAabnormalities that were identified involved novel maternallyinherited homoplasmic cytochrome B gene mutations

Figure 3 Average acyl-carnitine values (with s.e. bars) from 20 patients withconsistent abnormal elevations in multiple acyl-carnitines. Acyl-carnitine values arerepresented as percent upper limit of normal for each acyl-carnitine species. Noticethat C4OH, C14 and C16:1 are significantly elevated as compared with themaximum upper limit of normal.

Acyl-carnitine profiles in autismRE Frye et al

4

Translational Psychiatry

Page 46: Derrick MacFabe M.D.

(15533A4G and 15404T4C), which altered evolutionaryconserved amino acids. mtDNA content in muscle from fourpatients ranged from 109–189% of normal with a mean of160.5% (s.d.±27%).

Neurological and biochemical testing. Neurological andbiochemical characteristics of the patients are given inSupplementary Table 2. No abnormalities were found inthe majority of patients that underwent an extended 23 hvideo electroencephalogram. Acyl-glycine panel, aminoacids, glucose, insulin, Co-Q10, biotin, cholesterol andtriglyceride levels were unremarkable in all patients in whichthey were tested. Urine organic acids were abnormal in themajority in which it was tested with elevations in TCACmetabolites, specifically elevations in citrate and/or isocitraterepresenting the majority of the abnormalities. Lactate waselevated in about half of the patients. Carnitine panelwas abnormal in half of the patients in whom it was measuredwith 33% of the patients having high-esterified carnitine and17% having low free carnitine. Creatine kinase and pyruvatewas elevated in a minority of patients in which it wasmeasured. Interestingly, red-blood cell zinc was borderlinelow in 70% of the patients in which it was measured and red-blood cell copper was slightly elevated in 35% of the patients

in which it was measured. The great majority of individuals(90%) in which red-blood cell zinc and red-blood cellcopper were both measured demonstrated an abnormalityin at least one.

Muscle histology. All five of the patients that underwentmuscle biopsy demonstrated abnormal histological andelectron microscopy findings. Four demonstrated fibertype 1 predominance with two also demonstrating fiber type2 atrophy. The fifth demonstrated myofiber size irregularity,increased sarcoplasmic lipid and scattered succinate dehy-drogenase hyper-reactive fibers. Electron microscopydemonstrated an increased number of mitochondria in thesubsarcolemmal region in all and also in the intermyofibrillarregion in two. Mitochondria were maloriented in two patientsand degeneration of membranous organelles was seen inthree cases.

ETC and fatty-acid oxidation function. ETC functionwas testing on all five muscles biopsies. Corrected anduncorrected ETC activity is shown in Figures 5a and b,respectively, and demonstrate a partial defect in complexesI/III and I/III rotenone sensitive (RS). ETC and fatty-acidoxidation testing was conducted on fibroblast cultures from

Figure 4 Gluthathione abnormalities in four children with consistent elevations in multiple acyl-carnitine species. Notice that the patients have lower total (tGSH, mM) andfree (fGSH, mM) reduced gluthathione, as well as lower tGSH/fGSSG (free-oxidized gluthathione, mM) and fGSH/fGSSG ratios and higher fGSSG as compared with typicallydeveloping controls, suggesting both a reduction in the production of gluthathione and increase in gluthathione utilization by reactive species.

Figure 5 Electron transport chain (ETC) function of muscle (a, b) and fibroblast culture (c, d), as well as function of the fatty-acid oxidation pathway in fibroblast cultures(e, f). Graph values represent percent of normal ETC function, uncorrected (a, c) or corrected for citrate synthase (b, d). Muscle ETC results suggest a partial defectin complexes I/III and I/III rotenone sensitive (RS) while fibroblast culture ETC function suggests a partial defect in complex II/III activity. In fibroblast culture ETC studiescomplexes I/III RS and IV demonstrate considerable variability due to overactivity (4200% of the mean) in complex I/III RS in three patients and complex IV in onecase. Fatty-acid oxidation values represent mean of specific acyl-carnitine species (higher is worse) uncorrected (e) and corrected for citrate synthase (f). Elevation in theshort-chain fatty-acid D3-C4 was due to three patients demonstrating high D4-C4 values. The one patient with a significantly elevated D4-C4 value was found not to have amutation in ;the short-chain acyl-CoA dehydrogenase gene suggesting that the abnormalities in fatty-acids in fibroblast culture were due to other mitochondrial metabolismabnormalities.

Acyl-carnitine profiles in autismRE Frye et al

5

Translational Psychiatry

Page 47: Derrick MacFabe M.D.

eight patients. Overall complex II/III activity was deficientacross patients, and complex I/III RS demonstrated variableand elevated (4200% of the mean) activity, which was alsoseen to a lesser extent in complex IV (Figures 5c and d).Functional fatty-acid oxidation testing demonstrated eleva-tions in the short-chain fatty-acid D3-C4 (Figures 5e and f).

This effect was due to three patients with high D4-C4. Onlyone patient had a high enough elevation for a short-chainacyl-CoA dehydrogenase defect to be considered. However,sequencing of exons 1–10 of the short-chain acyl-CoAdehydrogenase gene (GeneDx, Gaithersburg, MD, USA)39

for the patient was normal.

ElectronTransport

Chain

Fatty Acids

Fatty Acids

Pyruvate

Acetyl-CoA

Acetyl-CoA

Complex INADH

CitrateOxalacetate NADH

Complex II

Complex III

IsocitrateTricarboxylicAcid Cycle

TricarboxylicAcid Cycle

Malate

NADH

Fumarate α-ketoglutarate

α-ketoglutarate

Complex IV

Succinyl-CoA

Succinyl-CoA

SuccinateComplex V

FADH2

FADH2

PyruvateLactate

NADHNADH

CitrateOxalacetate

Distal Malate ProximalIsocitrate

Fumarate NADH

Succinate

Methylmalonyl-CoAPropionic Acid Propionyl-CoA

Lactate

a

b

Figure 6 The tricarboxylic-acid cycle during (a) typical metabolism and (b) with high levels of propionic acid. Propionic acid is metabolized to propionyl-CoA, which inhibitsthe proximal portion of the tricarboxylic-acid cycle and enhances the distal portion of the tricarboxylic-acid cycle (see discussion for details). FADH2, flavin adenine dinucleotide;NADH, nicotinamide adenine dinucleotide.

Acyl-carnitine profiles in autismRE Frye et al

6

Translational Psychiatry

Page 48: Derrick MacFabe M.D.

MD diagnosis criteria. Using the Morava et al.40 criteria forthe five patients that underwent muscle biopsy found thatthree patients were rated as having a definite MD and twopatients were rated as having probable MD.

Discussion

In this report, 17% of a large cohort of children with ASDdemonstrated consistent elevations in short-chain and long-chain, but not medium-chain, acyl-carnitines. This pattern ofacyl-carnitine abnormalities is similar to elevations in brainacyl-carnitines seen in the PPA rodent model of ASD.13 Othermetabolic abnormalities, specifically mitochondrial dysfunc-tion and glutathione abnormalities, were identified in thepatient cohort that are similar to the PPA rodent model of ASD.Such abnormalities are discussed below in detail.

Mitochondrial abnormalities is ASD patients are consis-tent with PPA toxicity. ETC function testing in muscledemonstrated a partial deficit in complexes I/III and I/III RSactivity. These ETC abnormalities, along with other biomar-kers of mitochondrial dysfunction, are consistent with PPAinterfering with mitochondrial metabolism, potentially throughthe TCAC. The TCAC utilizes two electron carriers,nicotinamide adenine dinucleotide (NADH) and flavin ade-nine dinucleotide (FADH2) to shuttle electrons to complexes Iand II of the ETC, respectively. Normally TCAC reactionsproduce 3 NADH and 1 FADH2, resulting in a 3:1 NADH toFADH2 ratio (Figure 6a).

PPA is metabolized to propionyl-CoA, which is furthermetabolized to produce methymalonic-CoA. Methymalonic-CoA enters the TCAC half way through the cycle as succinyl-CoA, thereby essentially ‘short circuiting’ the TCAC(Figure 6b). Elevated succinyl-CoA enhances the distal halfof the TCAC and inhibits the proximal half of the TCAC. As thedistal half of the TCAC produces 1 NADH and 1 FADH2, if theproximal half of the TCAC is inhibited, the NADH to FADH2

ratio will change from 3:1 to 1:1. As NADH is metabolized bycomplex I a reduction in the production in NADH will result in arelative deficit in complex I, consistent with the findings fromthe ETC muscle studies. In addition, inhibition of the proximalportion of the TCAC will also result in a build-up of the firstmetabolites in the TCAC, consistent with the elevations incitrate and isocitrate in our patients. Furthermore, the endproduct of the fatty-acid oxidation pathway, acetyl-CoA, is thefirst metabolite of the proximal half of the TCAC. Thus,inhibition of the proximal half of the TCAC inhibits the fatty-acid oxidation pathway. This is consistent with the fact thatexamination of the fatty-acid oxidation pathway in fibroblastculture did not reveal any abnormalities to explain theelevations in acyl-carnitines.

Examination of fibroblasts, which occurs after 6 or moreweeks of fibroblast growth in culture, demonstrated aboveaverage activity of complexes I/III RS and IV and a partialdeficit in complex II/III. Interestingly, overactivity in complexesI and IV have been reported in children with ASD/MD.41,42 Thedisparity in the ETC findings between muscle and fibroblastculture can be explained by the alternative use of citratesynthase, the first enzyme in the TCAC, for metabolizingpropionyl-CoA. Normally citrate synthase produces citrate

from acetyl-CoA and oxaloacetate. In the context of highlevels of propionyl-CoA, citrate synthase produces methylcitrate, a dead end metabolite, from propionyl-CoA. This willalso result in a competition for citrate synthase by bothpropionyl-CoA and acetyl-CoA, further blocking the metabo-lism of metabolic pathways that produce acetyl-CoA as an endproduct, such as the fatty-acid oxidation pathway. Thisoveruse of citrate synthase will most likely also result in anupregulation of citrate synthase over time.

If an agent, such as PPA, that suppressed mitochondrialfunction was present in vivo but not in vitro, mitochondrialfunction in the muscle, but not the fibroblast culture, would bemore compromised. If upregulation of citrate synthaseoccurred due to excess PPA in vivo and then PPA wasremoved in vitro, the high activity of citrate synthase in vitrowould overproduce citrate and enhance the proximal portion ofthe TCAC, which preferentially produces NADH, the electroncarrier metabolized by complex I. This is consistent with theobserved I/III RS overactivity seen in fibroblast culture.

One common theme of the observed ETC dysfunction inboth muscle and fibroblast culture was that complex dysfunc-tion primarily occurred when evaluating the function ofcomplex III with complex I or II, suggesting that it is theinteraction between complex III and complex I or II rather thanat complex I or II specifically. Interesting, the 15533A4Gcytochrome b mutation identified in one of the cases has beenshown to have a complicated effect on complex III function.Rather than causing a frank decrease in complex III function,this mutation appears to result in delayed assembly of theI,III,IV supercomplex,43 thus influencing the interaction ofcomplex III with other complexes rather than specificallyaffecting only complex III. Interestingly, PPA has been shownto have its detrimental effect on the ETC through inhibition ofcomplex III function.44 Furthermore, alterations in brainomega 3/6 cardiolipin profiles found in the PPA rodent modelcould change inner mitochondrial membrane fluidity, and,thus, could potentially affect mitochondrial ETC complexinteractions.13

Mitochondrial abnormalities are consistent withacquired MD. All patients that underwent a workup for MDdemonstrated probable or definite MD by standardizedcriteria.2,40 However, the majority of the patients in this studydid not have any identifiable genetic causes for their MD.This is not surprising as only 23% of children with ASD andMD have a known mtDNA abnormality.3 The effect of thecytochrome b mutation found in two boys is complicated.45,46

For example, delayed supercomplex assembly associatedwith the 15533A4G gene mutation in a child with aneurodevelopment disorder was restored when mutanttransmitochondrial cybrids were developed from the15533A4G case.43 This suggests that this mtDNA mutationis a risk factor that requires interactions with nuclearmutations, polymorphisms or epigenetics and/or environ-mental triggers or modulators in order for the diseasephenotype to be expressed.43 Thus, the characteristics ofthis series of patient are consistent with the notion that thesystemic abnormalities seen in this subgroup of ASD patientsmay arise from environmental triggers7 in geneticallysensitive subpopulations.8,17,47,48

Acyl-carnitine profiles in autismRE Frye et al

7

Translational Psychiatry

Page 49: Derrick MacFabe M.D.

Glutathione metabolism and oxidative stress abnormal-ities in ASD patients. Four patients in the series underwentmeasurements of glutathione metabolism. Overall, there wasa marked decreased in total and free-reduced glutathionewith a slight increase in GSSG, suggesting a primary deficitin the production of glutathione and increase in utilization.This finding parallels the PPA rodent animal model. Indeed,intracerebroventricular infusion of PPA in rats decreasestotal glutathione in brain homogenates.11 Further evidencefor increased oxidative stress has been demonstrated in therodent PPA model, including increased brain protein carbo-nylation and lipid peroxidation, altered phospholipid profilesand increased activated microglia.11,13,16 Such findings areall consistent with those from ASD patients.49–51

Abnormalities in zinc and copper in ASD patients. Inter-estingly, a large proportion of the patients in the seriesdemonstrated mild abnormalities in zinc and/or copperconcentrations. Such abnormalities have been reported inthe ASD population previously52–54 and zinc supplementa-tion (along with B6) has been shown to decrease copperlevels and improve function in ASD in an uncontrolledstudy.54 Some have hypothesized that abnormalities in zincand copper metabolism could result in poor metallothioneinfunction, leading to susceptibility to environmental toxicantsthrough increased oxidative stress or mitochondrial dysfunc-tion.52 Interestingly, low zinc levels have been associatedwith pediatric inflammatory bowel disease55 and increasedinflammation in animal models of colitis56 and zinc supple-mentation appears to be protective of bowel inflammation inclinical57 and animal studies.58 Although speculative, thismay occur via a decrease in the activity of coloprotectivemetallothionein59 in the intestinal mucosa, or impairment ofT- and B-cell interaction60 that may contribute to gutdysbiosis favoring ASD-associated bacteria.

Potential links to unique ASD microbial populations.Enteric bacterial populations found in increased numbers instool samples of ASD patients (Clostridia, Desulfovibrio) areknown to produce PPA from fermentation of dietarycarbohydrates.18,19,31 Impaired carbohydrate digestion andtransport in children with ASD can result in a higherconcentration of dietary carbohydrates for these bacterialpopulations to ferment.18,19,61 A recent study has shown thatstool from ASD patients have elevations in PPA and othershort-chain fatty-acids.62 In addition, Desulfovibrio is capableof producing PPA from fermentation of peptones and canproduce hydrogen sulfide, a potential mitochondrial toxin,which may act synergistically with PPA to promote mitochon-drial dysfunction.17,18,63,64

Interestingly, administration of common antibiotics (that is,beta lactams) for routine pediatric infections alters gut florafavoring PPA-producing species. This could be significantconsidering the reported high incidence of antibiotic use insome ASD patients.18,19,65,66 In addition, this offers a potentialexplanation for temporary behavioral improvements in somepatients following vancomycin or metronidazole treatment,which eradicates these bacteria18,19,66 and profoundlyreduces stool PPA.67 Furthermore, removal of refinedcarbohydrates from the diet, which has been suggested as

an empiric treatment to improve the behavioral fluctuations,gastrointestinal symptoms and dysbiosis in ASD,8 may act byreducing substrate for these bacteria to produce PPA.11

Although low concentrations of PPA may be beneficial,humans with impairments in PPA metabolism (that is,propionic or methymalonic acidemia, holocarboxylase, bioti-nidase or B12 deficiency, valproate or ethanol exposure)exhibit neurodevelopmental conditions with behavioral andbiochemical similarities to ASD.68,69 PPA and related short-chain fatty-acids (that is, butyrate and acetate) have broadeffects on cellular systems.13,20,26,27,48,70–72 They are activelytaken up into the brain73 and can affect diverse physiologicalprocesses such as cell signaling,71 neurotransmitter synth-esis and release,70 mitochondrial function,20 lipid metabo-lism,13,74 immune function,75 cell–cell interactions76 and geneexpression.72 Thus, there are many potential mechanismswhere metabolic end products of the enteric microbiome canalter host physiology.17

Of particular interest, PPA is a known inhibitor of mitochon-drial function, through sequestration of carnitine and theproduction of propionyl-CoA, a potential cytotoxin.44,77 Methy-malonic acid, a metabolite of PPA, results in abnormalmitochondrial morphology,78 tissue specific ETC dysfunction,79

inhibition of the complex I and II function when interacting withcomplex III in the brain80 and reductions in reduced glu-tathione,78 similar to the patient cohort presented and the PPAanimal model. As PPA is metabolized through the TCAC, wepropose that excess exogenous PPA or related short-chainfatty-acid metabolites interfere with mitochondrial metabolism,thereby causing acquired mitochondrial dysfunction.

Impaired carnitine metabolism can act synergisticallywith PPA-producing bacteria. The Naþ dependentorganic cation/carnitine transporter 2 transports carnitineacross the gut-blood and blood-brain barriers.81 Antibiotics(that is, beta lactams) commonly used to treat pediatric,infections directly inhibit the organic cation/carnitine trans-porter 2 transporter, thus directly impairing carnitine reab-sorption.81 This could be significant considering the highincidence of antibiotic use in ASD patients, which can alsopromote gut dysbiosis favoring ASD-associated gut bacterialpopulations that produce PPA.18,19,65,66 Given that bothcarnitine deficiency and PPA can be detrimental to mito-chondrial metabolism, it is possible that antibiotic overusecan cause these two effects to act synergistically to cause anacquired mitochondrial disorder, especially in geneticallysusceptible individuals.

Interestingly, children with ASD, as a group, have beenfound to have reduced blood carnitine3 and a X-linked inbornerror of carnitine biosynthesis has been shown to be a riskfactor for ASD.82 Furthermore, oral carnitine, and its derivativeacetyl-L-carnitine, have both neuroprotective83,84 and colo-protective properties.85 Given that carnitine supplementationimproves function in children with ASD,86,87 it deserves furtherinvestigation as a therapeutic agent in ASD.3,86,87

Summary. This study has demonstrated that B17% ofchildren with ASD manifest biomarkers of abnormal mito-chondrial fatty-acid metabolism that parallel similar biomar-kers in the PPA rodent model of ASD. Detailed examination

Acyl-carnitine profiles in autismRE Frye et al

8

Translational Psychiatry

Page 50: Derrick MacFabe M.D.

of a subset of these patients indicates that these metabolicabnormalities are at least partly due to TCAC and ETCdysfunction. Genetic disorders do not appear to account forthe majority of these cases and the two individuals withabnormalities in mtDNA suggest dysfunction in the inter-action of complex III with complex I and/or II. For the cases inwhich genetic abnormalities have not been found it is verylikely that MD is acquired. As this subgroup of ASD patientshave several parallels with the rodent PPA model of ASD,13

this rodent model may be a useful tool to further examine thetemporal relation of behavioral bouts in relation to carnitine-acyl-carnitine fluctuations, and their possible response totherapeutic compounds thought to be useful in the treatmentof ASD and mitochondrial dysfunction such as carnitinesupplementation.

It is important to note that PPA affects multiple systems in acomplex manner and the evidence of increased PPA or othershort-chain fatty-acids being involved in the pathophysiologyof ASD, although compelling, is circumstantial at this stage.Thus, future studies should identify additional parallelsbetween the PPA rodent model of ASD and individuals withASD who manifest similar biomarkers. Further study of thismodel and this subgroup of ASD patients should improve ourunderstanding of the pathophysiology and potential riskfactors that lead to the metabolic, brain and behaviorabnormalities associated with ASD.

Conflict of interest

The authors declare no conflict of interest.

Acknowledgements. This research was supported in part from generouscontributions from GoodLife’s Children’s Foundation and Autism Research Instituteto DFM. We thank Dr Raymond Thomas (postdoctoral fellow) for published data onbrain lipids and Drs Stephan Kahler and Anirudh Saronwala for their insight intopropionic metabolism.

1. APA Diagnostic and statistical manual of mental disorders. 4th edn DC: AmericanPsychiatric Association: Washington, 1994.

2. Frye RE, Rossignol DA. Mitochondrial dysfunction can connect the diversemedical symptoms associated with autism spectrum disorders. Pediatr Res 2011; 69(5Pt 2): 41R–47RR.

3. Rossignol D, Frye RE. Mitochondrial dysfunction in autism spectrum disorders: Asystematic review and meta-analysis. Mol Psychiatry 2012; 17: 290–314.

4. Rossignol DA, Frye RE. A review of research trends in physiological abnormalities inautism spectrum disorders: immune dysregulation, inflammation, oxidative stress,mitochondrial dysfunction and environmental toxicant exposures. Mol Psychiatry 2012;17: 389–401.

5. Weissman JR, Kelley RI, Bauman ML, Cohen BH, Murray KF, Mitchell RL et al.Mitochondrial disease in autism spectrum disorder patients: a cohort analysis. PLoS ONE2008; 3: e3815.

6. Giulivi C, Zhang YF, Omanska-Klusek A, Ross-Inta C, Wong S, Hertz-Picciotto I et al.Mitochondrial dysfunction in autism. JAMA 2010; 304: 2389–2396.

7. Hallmayer J, Cleveland S, Torres A, Phillips J, Cohen B, Torigoe T et al. Genetic heritabilityand shared environmental factors among twin pairs with autism. Arch Gen Psychiatry2011; 68: 1095–1102.

8. Herbert MR, Russo JP, Yang S, Roohi J, Blaxill M, Kahler SG et al. Autism andenvironmental genomics. Neurotoxicology 2006; 27: 671–684.

9. Ashwood P, Anthony A, Torrente F, Wakefield AJ. Spontaneous mucosal lymphocytecytokine profiles in children with autism and gastrointestinal symptoms: mucosal immuneactivation and reduced counter regulatory interleukin-10. J Clin Immunol 2004; 24: 664 73.

10. MacFabe DF, Rodrı́guez-Capote K, Hoffman JE. A novel rodent model of autism:intraventricular infusions of propionic acid increase locomotor activity and induceneuroinflammation and oxidative stress in discrete regions of adult rat brain. Am J BiochemBiotechnol 2008; 4: 146–166.

11. MacFabe DF, Cain DP, Rodriguez-Capote K, Franklin AE, Hoffman JE, Boon F et al.

Neurobiological effects of intraventricular propionic acid in rats: possible role of short chain

fatty acids on the pathogenesis and characteristics of autism spectrum disorders. Behav

Brain Res 2007; 176: 149–169.12. Shultz SR, MacFabe DF, Ossenkopp KP, Scratch S, Whelan J, Taylor R et al.

Intracerebroventricular injection of propionic acid, an enteric bacterial metabolic end-

product, impairs social behavior in the rat: implications for an animal model of autism.

Neuropharmacology 2008; 54: 901–911.13. Thomas RH, Foley KA, Mepham JR, Tichenoff LJ, Possmayer F, MacFabe DF. Altered

brain phospholipid and acylcarnitine profiles in propionic acid infused rodents: further

development of a potential model of autism spectrum disorders. J Neurochem 2010; 113:

515–529.14. MacFabe DF, Cain NE, Boon F, Ossenkopp KP, Cain DP. Effects of the enteric bacterial

metabolic product propionic acid on object-directed behavior, social behavior, cognition,

and neuroinflammation in adolescent rats: relevance to autism spectrum disorder. Behav

Brain Res 2011; 217: 47–54.15. Shultz SR, Macfabe DF, Martin S, Jackson J, Taylor R, Boon F et al. Intracerebroven-

tricular injections of the enteric bacterial metabolic product propionic acid impair cognition

and sensorimotor ability in the Long-Evans rat: further development of a rodent model of

autism. Behav Brain Res 2009; 200: 33–41.16. Thomas RH, Meeking MM, Mepham JR, Tichenoff L, Possmayer F, Liu S et al. The enteric

bacterial metabolite propionic acid alters brain and plasma phospholipid molecular species:

further development of a rodent model of autism spectrum disorders. J Neuroinflammation

2012; 9: 153.17. MacFabe DF. Short-chain fatty acid fermentation products of the gut microbiome:

implications in autism spectrum disorders. Microb Ecol Health Dis 2012; 23: 19260.18. Finegold SM, Dowd SE, Gontcharova V, Liu C, Henley KE, Wolcott RD et al.

Pyrosequencing study of fecal microflora of autistic and control children. Anaerobe 2010;

16: 444–453.19. Finegold SM, Molitoris D, Song Y, Liu C, Vaisanen ML, Bolte E et al. Gastrointestinal

microflora studies in late-onset autism. Clin infect Dis 2002; 35(Suppl 1): S6–S16.20. Wajner M, Latini A, Wyse AT, Dutra-Filho CS. The role of oxidative damage in the

neuropathology of organic acidurias: insights from animal studies. J Inherit Metab Dis 2004;

27: 427–448.21. Coulter DL. Carnitine, valproate, and toxicity. J Child Neurol 1991; 6: 7–14.22. Calabrese V, Rizza V. Formation of propionate after short-term ethanol treatment and its

interaction with the carnitine pool in rat. Alcohol 1999; 19: 169–176.23. Al-Owain M, Kaya N, Al-Shamrani H, Al-Bakheet A, Qari A, Al-Muaigl S et al. Autism

spectrum disorder in a child with propionic acidemia. JIMD Reports 2012; 7: 63–66.24. Mostafa GA, El-Gamal HA, El-Wakkad ASE, El-Shorbagy OE, Hamza MM. Polyunsatu-

rated fatty acids, carnitine and lactate as biological markers of brain energy in autistic

children. Int J Child Neuropsychiatry 2005; 2: 179–188.25. Haas RH, Parikh S, Falk MJ, Saneto RP, Wolf NI, Darin N et al. The in-depth evaluation of

suspected mitochondrial disease. Mol Genet Metab 2008; 94: 16–37.26. Pastural E, Ritchie S, Lu Y, Jin W, Kavianpour A, Khine Su-Myat K et al. Novel plasma

phospholipid biomarkers of autism: mitochondrial dysfunction as a putative causative

mechanism. Prostaglandins Leukot Essent Fatty Acids 2009; 81: 253–264.27. Clark-Taylor T, Clark-Taylor BE. Is autism a disorder of fatty acid metabolism? possible

dysfunction of mitochondrial beta-oxidation by long chain acyl-CoA dehydrogenase. Med

Hypotheses 2004; 62: 970–975.28. Frye RE. Biomarkers of abnormal energy metabolism in children with autism spectrum

disorder. NAJ Med Sci 2012; 5: 141–147.29. Munnich A, Rustin P. Clinical spectrum and diagnosis of mitochondrial disorders. Am J Med

Genet 2001; 106: 4–17.30. Mitochondrial Medicine Society’s Committee on DHaas RH, Parikh S, Falk MJ, Saneto RP,

Wolf NI, Darin N et al. The in-depth evaluation of suspected mitochondrial disease.

Mol Genet Metab 2008; 94: 16–37.31. Wong LJ. Pathogenic mitochondrial DNA mutations in protein-coding genes. Muscle Nerve

2007; 36: 279–293.32. Wong LJ, Cobb BR, Chen TJ. Molecular analysis of mitochondrial DNA point mutations by

polymerase chain reaction. Methods Mol Biol 2006; 336: 135–143.33. Bai RK, Perng CL, Hsu CH, Wong LJ. Quantitative PCR analysis of mitochondrial DNA

content in patients with mitochondrial disease. Ann N Y Acad Sci 2004; 1011: 304–309.34. Bai RK, Wong LJ. Simultaneous detection and quantification of mitochondrial DNA

deletion(s), depletion, and over-replication in patients with mitochondrial disease. J Mol

Diagn 2005; 7: 613–622.35. Kirby DM, Thorburn DR, Turnbull DM, Taylor RW. Biochemical assays of respiratory chain

complex activity. Methods Cell Biol 2007; 80: 93–119.36. Roe CR, Roe DS. Recent developments in the investigation of inherited metabolic

disorders using cultured human cells. Mol Genet Metab 1999; 68: 243–257.37. James SJ, Melnyk S, Fuchs G, Reid T, Jernigan S, Pavliv O et al. Efficacy of

methylcobalamin and folinic acid treatment on glutathione redox status in children with

autism. Am J Clin Nutr 2009; 89: 425–430.38. Melnyk S, Pogribna M, Pogribny I, Hine RJ, James SJ. A new HPLC method for the

simultaneous determination of oxidized and reduced plasma aminothiols using coulometric

electrochemical detection. J Nutr Biochem 1999; 10: 490–497.

Acyl-carnitine profiles in autismRE Frye et al

9

Translational Psychiatry

Page 51: Derrick MacFabe M.D.

39. Corydon MJ, Vockley J, Rinaldo P, Rhead WJ, Kjeldsen M, Winter V et al. Role of common

gene variations in the molecular pathogenesis of short-chain acyl-CoA dehydrogenasedeficiency. Pediatr Res 2001; 49: 18–23.

40. Morava E, van den Heuvel L, Hol F, de Vries MC, Hogeveen M, Rodenburg RJ et al.Mitochondrial disease criteria: diagnostic applications in children. Neurology 2006; 67:

1823–1826.41. Graf WD, Marin-Garcia J, Gao HG, Pizzo S, Naviaux RK, Markusic D et al. Autism

associated with the mitochondrial DNA G8363A transfer RNA(Lys) mutation. J Child Neurol

2000; 15: 357–361.42. Frye RE, Naviaux RK. Autistic disorder with complex IV overactivity: a new mitochondrial

syndrome. J Ped Neurol 2011; 9: 427–434.43. Gil Borlado MC, Moreno Lastres D, Gonzalez Hoyuela M, Moran M, Blazquez A, Pello R et

al. Impact of the mitochondrial genetic background in complex III deficiency. PloS one

2010; 5: e12801.44. Schwab MA, Sauer SW, Okun JG, Nijtmans LG, Rodenburg RJ, van den Heuvel LP et al.

Secondary mitochondrial dysfunction in propionic aciduria: a pathogenic role for

endogenous mitochondrial toxins. Biochem J 2006; 398: 107–112.45. Gil Borlado MC, Moreno Lastres D, Gonzalez Hoyuela M, Moran M, Blazquez A, Pello R et

al. Impact of the mitochondrial genetic background in complex III deficiency. PLoS ONE2010; 5: e12801.

46. Frye RE. Novel cytochrome b gene mutations causing mitochondrial disease in autism.

J Ped Neurol 2012; 10: 1–6.47. Ashwood P, Van de Water J. Is autism an autoimmune disease? autoimmunity reviews

2004; 3: 557–562.48. Ashwood P, Van de Water J. A review of autism and the immune response. Clin Dev

Immunol 2004; 11: 165–174.49. James SJ, Melnyk S, Jernigan S, Cleves MA, Halsted CH, Wong DH et al. Metabolic

endophenotype and related genotypes are associated with oxidative stress in children with

autism. Am J Med Genet B, Neuropsychiatr Genet 2006; 141B: 947–956.50. Chauhan A, Chauhan V. Oxidative stress in autism. Pathophysiology 2006; 13: 171–181.51. Wegiel J, Kuchna I, Nowicki K, Imaki H, Wegiel J, Marchi E et al. The neuropathology of

autism: defects of neurogenesis and neuronal migration, and dysplastic changes. Acta

Neuropathol 2010; 119: 755–770.52. Faber S, Zinn GM, Kern JC 2nd, Kingston HM. The plasma zinc/serum copper ratio as a

biomarker in children with autism spectrum disorders. Biomarkers 2009; 14: 171–180.53. Yasuda H, Yoshida K, Yasuda Y, Tsutsui T. Infantile zinc deficiency: association with

autism spectrum disorders. Scientific reports 2011; 1: 129.54. Russo AJ, Devito R. Analysis of copper and zinc plasma concentration and the

efficacy of zinc therapy in individuals with asperger’s syndrome, pervasive developmental

disorder not otherwise specified (PDD-NOS) and autism. Biomark Insights 2011; 6:

127–133.55. Sikora SK, Spady D, Prosser C, El-Matary W. Trace elements and vitamins at diagnosis in

pediatric-onset inflammatory bowel disease. Clin Pediatr 2011; 50: 488–492.56. Iwaya H, Kashiwaya M, Shinoki A, Lee JS, Hayashi K, Hara H et al. Marginal zinc

deficiency exacerbates experimental colitis induced by dextran sulfate sodium in rats.J Nutr 2011; 141: 1077–1082.

57. Scrimgeour AG, Condlin ML. Zinc and micronutrient combinations to combat

gastrointestinal inflammation. Curr Opin Clin Nutr Metab Care 2009; 12: 653–660.58. Barollo M, Medici V, D’Inca R, Banerjee A, Ingravallo G, Scarpa M et al. Antioxidative

potential of a combined therapy of anti TNFalpha and Zn acetate in experimental colitis.World J Gastroenterol 2011; 17: 4099–4103.

59. Tran CD, Ball JM, Sundar S, Coyle P, Howarth GS. The role of zinc and metallothionein inthe dextran sulfate sodium-induced colitis mouse model. Dig Dis Sci 2007; 52: 2113–2121.

60. Zhang Y, Okamura S, Kudo T, Masuo T, Mori M. Calcineurin inhibition by polaprezinc in

rats with experimentally-induced colitis. Life Sci 2011; 88: 432–439.61. Williams BL, Hornig M, Buie T, Bauman ML, Cho Paik M, Wick I et al. Impaired

carbohydrate digestion and transport and mucosal dysbiosis in the intestines of childrenwith autism and gastrointestinal disturbances. PLoS One 2011; 6: e24585.

62. Wang L, Christophersen CT, Sorich MJ, Gerber JP, Angley MT, Conlon MA. Elevated fecalshort chain fatty acid and ammonia concentrations in children with autism spectrum

disorder. Dig Dis Sci 2012; 57: 2096–2102.63. Barton LL, Fauque GD. Biochemistry, physiology and biotechnology of sulfate-reducing

bacteria. Adv Appl Microbiol 2009; 68: 41–98.64. Lagoutte E, Mimoun S, Andriamihaja M, Chaumontet C, Blachier F, Bouillaud F. Oxidation

of hydrogen sulfide remains a priority in mammalian cells and causes reverse electron

transfer in colonocytes. Biochim Biophys Acta 2010; 1797: 1500–1511.65. Fallon J. Could one of the most widely prescribed antibiotics amoxicillin/clavulanate

"augmentin" be a risk factor for autism? Med Hypotheses 2005; 64: 312–315.

66. Sandler RH, Finegold SM, Bolte ER, Buchanan CP, Maxwell AP, Vaisanen ML et al. Short-term benefit from oral vancomycin treatment of regressive-onset autism. J Child Neurol2000; 15: 429–435.

67. Mellon AF, Deshpande SA, Mathers JC, Bartlett K. Effect of oral antibiotics on intestinalproduction of propionic acid. Arch Dis Child 2000; 82: 169–172.

68. Yorifuji T, Kawai M, Muroi J, Mamada M, Kurokawa K, Shigematsu Y et al. Unexpectedlyhigh prevalence of the mild form of propionic acidemia in Japan: presence of a commonmutation and possible clinical implications. Hum Genet 2002; 111: 161–165.

69. Perez B, Desviat LR, Rodriguez-Pombo P, Clavero S, Navarrete R, Perez-Cerda C et al.Propionic acidemia: identification of twenty-four novel mutations in Europe and NorthAmerica. Mol Genet Metab 2003; 78: 59–67.

70. DeCastro M, Nankova BB, Shah P, Patel P, Mally PV, Mishra R et al. Short chain fattyacids regulate tyrosine hydroxylase gene expression through a cAMP-dependent signalingpathway. Brain Res Mol Brain Res 2005; 142: 28–38.

71. Nakao S, Moriya Y, Furuyama S, Niederman R, Sugiya H. Propionic acid stimulatessuperoxide generation in human neutrophils. Cell Biol Int 1998; 22: 331–337.

72. Nguyen NH, Morland C, Gonzalez SV, Rise F, Storm-Mathisen J, Gundersen V et al.Propionate increases neuronal histone acetylation, but is metabolized oxidatively by glia.Relevance for propionic acidemia. J Neurochem 2007; 101: 806–814.

73. Conn AR, Fell DI, Steele RD. Characterization of alpha-keto acid transport across blood-brain barrier in rats. Am J Physiol 1983; 245: E253–E260.

74. Hara H, Haga S, Aoyama Y, Kiriyama S. Short-chain fatty acids suppress cholesterolsynthesis in rat liver and intestine. J Nutr 1999; 129: 942–948.

75. Le Poul E, Loison C, Struyf S. Functional characterization of human receptors for shortchain fatty acids and their role in polymorphonuclear cell activation. J Biol Chem 2003; 278:25481–25489.

76. Rorig B, Klausa G, Sutor B. Intracellular acidification reduced gap junctioncoupling between immature rat neocortical pyramidal neurones. J Physiol 1996; 490(Pt 1): 31–49.

77. Brass EP. Interaction of carnitine and propionate with pyruvate oxidation by hepatocytesfrom clofibrate-treated rats: importance of coenzyme A availability. J Nutr 1992; 122:234–240.

78. Chandler RJ, Zerfas PM, Shanske S, Sloan J, Hoffmann V, DiMauro S et al.Mitochondrial dysfunction in mut methylmalonic acidemia. The FASEB Journal 2009; 23:1252–1261.

79. Pettenuzzo LF, Ferreira Gda C, Schmidt AL, Dutra-Filho CS, Wyse AT, Wajner M.Differential inhibitory effects of methylmalonic acid on respiratory chain complex activities inrat tissues. International J Dev Neurosci 2006; 24: 45–52.

80. Brusque AM, Borba Rosa R, Schuck PF, Dalcin KB, Ribeiro CA, Silva CG et al. Inhibition ofthe mitochondrial respiratory chain complex activities in rat cerebral cortex bymethylmalonic acid. Neurochem Int 2002; 40: 593–601.

81. Miecz D, Januszewicz E, Czeredys M, Hinton BT, Berezowski V, Cecchelli R et al.Localization of organic cation/carnitine transporter (OCTN2) in cells forming the blood-brainbarrier. J Neurochem 2008; 104: 113–123.

82. Celestino-Soper PB, Violante S, Crawford EL, Luo R, Lionel AC, Delaby E et al. A commonX-linked inborn error of carnitine biosynthesis may be a risk factor for nondysmorphicautism. Proc Nat Acad Sci USA 2012; 109: 7974–7981.

83. Scafidi S, Racz J, Hazelton J, McKenna MC, Fiskum G. Neuroprotection by acetyl-L-carnitine after traumatic injury to the immature rat brain. Dev Neurosci 2010; 32: 480–487.

84. Patel SP, Sullivan PG, Lyttle TS, Rabchevsky AG. Acetyl-L-carnitine amelioratesmitochondrial dysfunction following contusion spinal cord injury. J Neurochem 2010;114: 291–301.

85. Fortin G, Yurchenko K, Collette C, Rubio M, Villani AC, Bitton A et al. L-carnitine, adiet component and organic cation transporter OCTN ligand, displays immuno-suppressive properties and abrogates intestinal inflammation. Clin Exp Immunol 2009;156: 161–171.

86. Geier DA, Kern JK, Davis G, King PG, Adams JB, Young JL et al. A prospective double-blind, randomized clinical trial of levocarnitine to treat autism spectrum disorders. Med SciMonit 2011; 17: PI15–PI23.

87. Rossignol DA. Novel and emerging treatments for autism spectrum disorders: a systematicreview. Ann Clin Psychiatr 2009; 21: 213–236.

Translational Psychiatry is an open-access journalpublished by Nature Publishing Group. This work is

licensed under the Creative Commons Attribution-NonCommercial-NoDerivative Works 3.0 Unported License. To view a copy of this license,visit http://creativecommons.org/licenses/by-nc-nd/3.0/

Supplementary Information accompanies the paper on the Translational Psychiatry website (http://www.nature.com/tp)

Acyl-carnitine profiles in autismRE Frye et al

10

Translational Psychiatry


Recommended