• No results found

Translational studies in Zellweger spectrum disorders - Chapter 1: General introduction

N/A
N/A
Protected

Academic year: 2021

Share "Translational studies in Zellweger spectrum disorders - Chapter 1: General introduction"

Copied!
13
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

UvA-DARE is a service provided by the library of the University of Amsterdam (https://dare.uva.nl)

Translational studies in Zellweger spectrum disorders

Berendse, K.

Publication date

2016

Document Version

Final published version

Link to publication

Citation for published version (APA):

Berendse, K. (2016). Translational studies in Zellweger spectrum disorders.

General rights

It is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s)

and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an open

content license (like Creative Commons).

Disclaimer/Complaints regulations

If you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please

let the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the material

inaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letter

to: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. You

will be contacted as soon as possible.

(2)

Chapter 1

(3)

1

Peroxisomes

Peroxisomes are multipurpose organelles surrounded by a membrane containing

different transport systems that control in- and efflux of metabolites, are present

in almost all eukaryotic cells

1

and are involved in multiple different metabolic

pathways, both catabolic and anabolic

2

. The name of the organelle originates

from the description by De Duve and Baudhuin

3

, after Rhodin first described the

discovery of a new organelle in mouse kidney cells in 1954

4

. In 1966, De Duve

and Baudhuin were the first to describe the presence of catalase and hydrogen

peroxide generating enzymes in specific microbodies and therefore named these

organelles “peroxisomes”. In humans, peroxisomes are found in every cell, except

red blood cells. Peroxisomes are approximately 1µm in width and have a half-life of

approximately 2 days in cultured mammalian cells

5

. The number of peroxisomes

varies considerably among species and in different cell types, and also depends

on the developmental stage and environmental conditions. Peroxisomes adapt to

metabolic needs as they can proliferate and multiply or be degraded in response to

various stimuli

67

. Because this thesis only comprises research in humans and mice,

yeast peroxisomes will not be discussed in detail, although yeast is an excellent

model to study more fundamental aspects of peroxisomal function. For a detailed

review see

89

.

Peroxisome biogenesis

Peroxins are essential for the biogenesis of normal peroxisomes. There are 14

human peroxins, encoded by 14 different

PEX

genes

10

. These peroxins are involved

in either import of peroxisomal membrane proteins (PMPs, i.e.

PEX3, PEX11, PEX16

and PEX19) or import of peroxisomal matrix proteins (i.e.

PEX1,

PEX2, PEX5, PEX6,

PEX7, PEX10, PEX12, PEX13, PEX14, PEX26) (figure 1). Within the PMP import

process, PEX3 is located in the peroxisomal membrane and acts as a docking site

for PEX19, which is a protein located in the cytosol and responsible for binding

PMPs

11

. The exact function of PEX16 is unknown, but it has been hypothesized

that it functions as a receptor for PEX19

10

. When defects involve these proteins,

peroxisome biogenesis is disturbed and often no peroxisomal remnants can be

found in the cell

12

.

Transportation of peroxisomal matrix proteins into peroxisomes is mediated by two

peroxisomal targeting sequences (i.e. PTS-1 and PTS-2)

1314

, which are recognized

by specific cytosolic receptor proteins, being the PEX5 small isoform (PEX5S) for

C-terminal tripeptide PTS1 motifs and the PEX5 long form (PEX5L)- PEX7 complex

for N-terminal octapeptide PTS2 proteins, respectively

15 16 17

(figure 1). PEX5L

contains 37 additional amino acids that allows it to bind to PEX7, transporting PEX7

to the peroxisomal membrane. Only three proteins are currently known to have a

PTS2 signal in humans, i.e. thiolase (for

β

-oxidation), phytanoyl-CoA 2-hydroxylase

(for

α

-oxidation) and alkyl-dihydroxyacetone-phosphate synthase (involved in the

ether lipid synthesis pathway)

2

. The import machinery of peroxisomal matrix proteins

consists of the PEX14- PEX13 complex, which is the docking complex for PEX5 and

PEX7

1819

. Thereafter the targeted proteins are transported across the peroxisomal

membrane via PEX2, PEX10 and PEX12

20

. Notably, it was recently reported that

the import of PTS-1 and PTS-2 proteins is performed by distinct pores in yeast

21

.

Subsequently, PEX5 and PEX7 are released from the membrane for another import

cycle (via mono-ubiquitination) or targeted to the proteasome for degradation (via

poly-ubiquitination)

22

, so called pexophagy

23

. Pexophagy is a process that has

extensively been studied in yeast

24

. The PEX1, PEX6 and PEX26 complex releases

PEX5 from the peroxisomal membrane

25

. When a defect is located in peroxins

involved in the matrix protein import, remnants of peroxisomes, either missing a

group of matrix proteins or without any content (“ peroxisomal ghosts”) can be

observed

12

. PEX11, which consists of three isoforms, including PEX11

α

, PEX11

β

and PEX11

γ

, is involved in the elongation and constriction of peroxisomes. It is

hypothesized that PEX11

α

is involved in the proliferation and PEX11

β

in growth

and division of peroxisomes

26

. The exact function of these isoforms remains to be

elucidated and will not be discussed further.

In recent years, several theories were proposed regarding the formation/division

of peroxisomes, mainly based on research in yeast models

2728

. The current view

Peroxisomal lumen Cytosol 7 7 5S 5S 5L 5L pts-1 pts-1 pts-2 pts-2 7 14 13 2 12 10 26 6 1 16 3 PMP 19 PMP 19 PMP Matrix protein import Membrane protein import

(4)

1

is that peroxisomes are formed

de novo

from the endoplasmic reticulum, but they

can also be formed from pre-existing peroxisomes via division

6 29

. Peroxisomal

division occurs via three different processes, including elongation, constriction

and fission. Several proteins, i.e. PEX11, dynamin-like protein 1/dynamin-related

protein 1 (DLP1/DRP1)

30

, fission 1 (FIS1)

31

, mitochondrial fission factor (Mff)

32

and ganglioside-induced differentiation associated protein 1 (GDAP1)

33

have been

discovered to play a role in this process in humans. Notably, all these proteins,

except for PEX11, are also involved in the fission of mitochondria.

Several co-factors, substrates and products, such as NADH and ATP are necessary

for proper peroxisomal function. Transportation of small solute molecules occurs

via the channel forming protein PXMP2

34

, whereas transport of large molecules

(i.e. so called “bulky” solutes), but also ATP and cofactors, requires active transport

35

. Among others, this active transport is mediated by three ATP-binding cassettes

(ABC) transporters, classified as “subfamily D”; ABCD-1, -2 and -3

36 37

. These

transporters have a hydrophobic transmembrane domain and hydrophilic region,

and are also known as adrenoleukodystrophy protein

38

, adrenoleukodystrophy

related protein

39

and peroxisomal membrane protein 70

40

, respectively. Research

in yeast, mice and humans revealed that very-long chain acyl-CoAs are transported

via ABCD1

41 42

, whereas ABCD2 is believed to play a role in the

β

-oxidation of

polyunsaturated fatty acids (e.g. DHA), as illustrated by the

ABCD2

knockout

mice

43

. ABCD2 also accepts certain substrates for ABCD1

44

. ABCD3 is the most

abundant and mediates the transport of branched-chain acyl-CoAs

45

.

Peroxisomal functions

As mentioned before, peroxisomes play an important role in several metabolic

processes. Knowledge of many of these processes was gained from the discovery

of patients with peroxisomal defects and associated biochemical alterations in

plasma, urine and skin fibroblasts. Below, the main peroxisomal metabolic pathways

in humans will be described briefly.

Fatty acid

α

-oxidation

Fatty acids with a methyl group at the third position have to be shortened by one

carbon atom via the peroxisomal

α

-oxidation system to produce 2-methyl fatty

acids. Phytanic acid (3,7,11,14-tetramethylhexadecanoic acid) is a typical example

of a 3-methyl branched chain fatty acid that is converted into the corresponding

2-methyl fatty acid pristanic acid (2,4,6,10-tetramethylpentadecanoic acid)

46

, which

subsequently can be degraded via the peroxisomal

β

-oxidation system

4748

. Before

α

-oxidation can occur, the substrate needs to be activated to an acyl-CoA ester.

Fatty acid

β

-oxidation

Fatty acid

β

-oxidation (i.e. the removal of the carboxyl end of the fatty acid) is one of

the main functions of peroxisomes. Similar to the mitochondrial fatty acid

β

-oxidation,

the oxidation consists of four different steps, including dehydrogenation, subsequent

hydration, dehydrogenation and thiolytic cleavage

49

. In each cycle, 2 carbon atoms

of different substrates are removed. Because the peroxisomal

β

-oxidation system

cannot degrade fatty acids fully into acetyl-CoA units, the shortened fatty acids

are shuttled to the mitochondria for further degradation, which may occur via two

different routes including a carnitine dependent route and a free fatty acid route

(see

2

for detailed discussion). One of the most abundant omega-3 fatty acids is

docosahexaenoic acid (DHA). DHA is considered to be essential for normal brain

and retinal development and function

50 5152

. Its precursor tetracosahexanaenoic

acid undergoes one cycle of peroxisomal

β

-oxidation to form DHA

53

. In addition,

the peroxisomal

β

-oxidation system can handle other substrates such as

branched chain fatty acids like pristanic acid and the bile acid intermediates di-

and trihydroxycholestanoic acid, long chain dicarboxylic acids, the side chains of

eicosanoids and other substrates described below.

Bile acid synthesis

Liver peroxisomes are involved in the production of bile acids. The bile acid

intermediates dihydroxycholestanoic acid (DHCA) and trihydroxycholestanoic acid

(THCA) undergo one cycle of peroxisomal

β

-oxidation to form the CoA esters of the

primarily bile acids chenodeoxycholic acid (CDCA) and cholic acid (CA), respectively

54

. These bile acids are then conjugated to their tauro- or glyco-conjugated forms

by the peroxisomal/cytosolic enzyme bile acid-CoA:amino acid N-acyltransferase

(BAAT)

55

. Subsequently these bile acids are exported from the peroxisomes into bile

canaliculi to be further excreted into the bile.

Glyoxylate detoxification

Glycolate is consumed via vegetables or formed in the mitochondria after

catabolism of hydroxyl proline, derived from meat or endogenous collagen turnover.

Glycolate is metabolized to glyoxylate in peroxisomes. The end point of glyoxylate

detoxification in liver and kidney, as catalysed by the peroxisomal enzyme

alanine-glyoxylate aminotransferase (AGT), is glycine

56

. When peroxisomes are absent and

AGT is localized in the cytosol, glyoxylate is oxidized to oxalate by cytosolic lactate

dehydrogenase. Oxalate cannot be degraded further in vertebrate species and can

readily precipitate as calcium oxalate thus forming insoluble crystals. Because of its

insolubility, oxalate will deposit in tissue and body fluids, with detrimental effects on

many organs, including the kidneys.

56

.

(5)

1

Ether phospholipids synthesis

Phospholipids are major components of the cell membrane and play an important

role in a variety of processes

57 58

. Phospholipids have a glycerol or sphingosyl

backbone with a polar or hydrophilic head group and nonpolar or hydrophobic

fatty acid side chains

59

. The majority of phospholipids have an ester bond at the

sn-1 position of the glycerol backbone. Ether phospholipids, however, have an

ether bond at this latter position. In mammals, ether phospholipids usually occur

in their plasmalogen form, as characterized by an unsaturated 1-0-alkenyl (vinyl

ether) group at the sn-1 position. Plasmalogens are found in myelin, heart muscle,

erythrocytes and retina

60

. Formation of this characteristic ether bond, requires the

obligatory participation of peroxisomes by virtue of the fact that the two enzymes

dihydroxyacetonephosphate transferase (DHAPAT) and alkylglycerone phosphate

synthase (AGPS) are strictly peroxisomal enzymes. The additional steps for full

synthesis of plasmalogens are performed in the endoplasmic reticulum

60

.

Other peroxisomal functions

Other metabolic processes in which peroxisomes play a role include: cellular

redox metabolism, with pro-oxidant enzymes (FAD-linked oxidases) and the

anti-oxidant enzymes catalase and Cu/Zn-superoxide dismutase

61

. Other functions of

peroxisomes include: the oxidation of L-pipecolic acid and D-amino acids, fatty acid

chain elongations and the oxidation of polyamines, although these are all (minor)

peroxisomal processes. For a detailed review see

2

.

Interplay with other organelles

Peroxisomes are known to interact with different organelles

62

. For example,

because peroxisomes cannot degrade fatty acids to completion, due to the lack

of a respiratory chain, the products of peroxisomal β-oxidation have to be shuttled

to mitochondria for full oxidation to CO2 and H2O

2

. Moreover, peroxisomes and

mitochondria interplay in the lipid homeostasis, the production and scavenging of

reactive oxygen species, maintenance of redox states, thermogenesis and they share

similar proteins for division and fission, which will not be discussed in detail

63 64 65

. An

example is the fission protein DLP1. When defective it results in both a peroxisomal

and mitochondrial fission defect with devastating results on clinical phenotype

66

.

There are also enzymes that are located in both mitochondria and peroxisomes such

as 3-hydroxy-3-methylglutaryl-CoA lyase

67

and alpha- methylacyl-CoA racemase

(AMACR)

68

. It is important to note that since the discovery of peroxisomal diseases,

concomitant mitochondrial alterations and/or dysfunction have also been reported

69 70 71

.

Peroxisomal disorders

After the initial discovery of peroxisomes in 1954, Goldfischer was the first to

associate disease with peroxisomal dysfunction/absence in 1973

69

. Since 1973,

multiple inherited metabolic diseases of varying extent, related to peroxisomal

dysfunction, were discovered. The peroxisomal disorders can be grouped into two

different classes based on the underlying defect in (1) the biogenesis of peroxisomes

or (2) defects in single peroxisomal enzymes (table 1).

Peroxisome biogenesis disorders

The group of peroxisome biogenesis disorders (PBDs) is comprised of two

main entities, including the Zellweger spectrum disorders (ZSDs) and rhizomelic

chondrodysplasia punctata (RCDP, type 1 and 5). Other PBDs include defects

in the genes for peroxisomal fission, i.e. mutations in

DLP1

66

,

MFF

72

,

GDAP1

33

and

PEX11

β 73

. Details about ZSDs will be described in Chapter 2. Briefly, the

ZSD are characterized by a spectrum of disease severity ranging from the very

mild Heimler syndrome

74

and relatively mild infantile Refsum disease (IRD)

75

to

neonatal adrenoleukodystrophy (NALD)

76

and the most severe phenotype, classical

Zellweger syndrome (ZS)

77

. As a result of peroxisomal dysfunction of varying extent,

several biochemical markers can be measured in plasma, urine and erythrocytes,

including among others elevated levels of very-long chain fatty acids (VLCFAs) and

bile acid intermediates DHCA and THCA.

The pathology within the RCDP spectrum is due to defects in the plasmalogen

synthesis, leading to reduced levels of plasmalogens. Clinically these patients have

cataract, skeletal dysplasia (i.e. shortening of proximal long bones, rhizomelia) and

calcification of cartilage (chondrodysplasia punctata) on X-ray

78

. The life expectancy

is usually short

79

. However, milder phenotypes are also described

80

. RCDP 1 is

caused by mutations in

PEX7 and is the most common form of RCDP

80

. RCPD

type 5 was recently added to the RCDP spectrum and results from a mutation in

PEX5L

81

.

Single peroxisomal enzyme deficiencies

Because this thesis is primarily focused on the peroxisome biogenesis disorders,

the single peroxisomal enzyme deficiencies will only be described briefly. For a

detailed review see

26

.

X-linked adrenoleukodystrophy (X-ALD): X-ALD is the peroxisomal disease with the

highest prevalence and affects both men and women

82

. In males, the first disease

(6)

1

Table 1 Classification of peroxisomal disorders

Disorder Gene

Peroxisome biogenesis disorders

Dynamin-like protein 1 (DLP1) deficiency DLP1 Ganglioside-induced differentiation associated protein 1

(GDAP1) deficiency GDAP1

Mitochondrial fission factor (Mff) deficiency MFF Rhizomelic chondrodysplasia punctata type 1 (RCDP1) PEX7

Rhizomelic chondrodysplasia punctata type 5 (RCDP5) PEX5 (long isoform) Zellweger spectrum disorders (ZSDs) PEX1, 2, 3, 5, 6, 10, 11β, 12, 13, 14, 16,

19, 26

Single peroxisomal enzyme deficiencies

70-kDa peroxisomal membrane protein (PMP70)

deficiency ABCD3

α-methylacyl-CoA racemase (AMACR) deficiency AMACR

Acatalasemia CAT

Acyl-CoA oxidase 1 (ACOX1) deficiency ACOX1 Bile acid-CoA:amino acid N-acyltransferase (BAAT)

deficiency BAAT

Classic Refsum disease PHYH/PEX7

D-bifunctional protein (DBP) deficiency HSD17B4

Isolated glycolic aciduria HAO1

Primary hyperoxaluria type I AGXT

Rhizomelic chondrodysplasia punctata type 2 (RCDP2) GNPAT Rhizomelic chondrodysplasia punctata type 3 (RCDP3) AGPS Rhizomelic chondrodysplasia punctata type 4 (RCDP4) FAR1 Sterol carrier protein X (SCPx) deficiency SCP2

X-adrenoleukodystrophy (X-ALD) ABCD1

progressive cerebral demyelination and/or myelopathy. Female X-ALD carriers

develop signs of myelopathy and/or peripheral neuropathy

83

. Overall, the symptoms

are progressive with age. The disease is caused by a mutation in the

ABCD1 gene

84

and characterised by accumulation of VLCFAs in tissues and plasma. Recently,

X-ALD was implemented in the newborn screening in New York, USA

85

and it

will soon be implemented in The Netherlands for male neonates. Because levels

of C26:0-lysophosphatidylcholine (C26:0-lysoPC) in bloodspots are determined for

this newborn screening, ZSD patients will also be diagnosed, as they also present

with elevated levels of C26:0-lysoPC.

Peroxisomal acyl-CoA oxidase 1 (ACOX1) deficiency is caused by mutations in

the

ACOX1 gene

86

. Because

ACOX1 is involved in the

β

-oxidation of VLCFAs,

patients present with elevated levels of VLCFAs in plasma and fibroblasts. Clinical

presentations range from a fatal, early onset form with hypotonia, frequent seizures,

impaired vision and hearing to a late-onset form with cerebellar abnormalities.

To date multiple patients with ACOX deficiency have been reported

8788

. Proper

diagnosis requires enzyme testing followed by mutation analysis.

2-Methyl-CoA racemase (AMACR) processes fatty acids with a methyl group in

the (R)-configuration to the (S)-configuration, making them ready for peroxisomal

β

-oxidation. The AMACR enzyme is located in both peroxisomes and mitochondria,

and the two forms are produced from the same gene

8968

. Patients with AMACR

deficiency accumulate the (R)-configuration of pristanic acid and bile acid

intermediates

90

. Clinical symptoms include, among others, liver disease, retinopathy,

peripheral neuropathy and epilepsy in old and/or young

90919293

.

D-bifunctional protein (DBP) deficiency is caused by a defect in the enzyme

catalysing the second and third step of the peroxisomal

β

-oxidation. Mutations in the

HSD17B4 gene result in DBP deficiency. The clinical phenotype is broad

94

ranging

from patients dying in the first year of life to patients surviving into adolescence,

albeit with severe handicaps (Prof. Poll-The, unpublished data). Recently, several

adult patients, with only minor clinical symptoms were diagnosed using whole

exome sequencing

9596

. Because DBP is involved in the

β

-oxidation of straight chain

VLCFAs as well as branched-chain FAs, patients accumulate VLCFAs, pristanic and

bile acid intermediates in plasma. The enzyme activity can be measured in skin

fibroblasts, differentiating DBP from ZSD.

Different mutations in the

PHYH gene result in Refsum disease

97

, leading to a

dysfunctional phytanoyl-CoA 2-hydroxylase, an enzyme involved in the first step

of the peroxisomal

α

-oxidation

98

. Furthermore, specific mutations in

PEX7 are also

reported to cause Refsum disease

99

. As phytanic acid is present in daily food,

(7)

1

levels in plasma

100

. Therefore the therapy for this disease is relatively simple: a

phytanic acid restricted diet. The clinical phenotype is broad ranging from retinitis

pigmentosa, deafness, polyneuropathy and anosmia to liver disease

101

.

Currently three different types of RCDP, caused by defects in single peroxisomal

enzymes, exist. A mutation in the DHAPAT gene results in RCDP type 2

102

. RCDP

type 3 is caused by mutations in the alkyl-dihydroxyacetone phosphate synthase

gene

103

. Specific mutations in fatty acyl-CoA reductase 1 are described to cause a

deficiency in the production of fatty acid alcohols for the plasmalogens biosynthesis

104

, which is referred to as RCDP type 4

105

. In general, these diseases cannot

be distinguished based on clinical phenotype and patients present with similar

symptoms as the RCDP type 1 phenotype

57

, except that no rhizomelia appears to

be present in RCDP type 4 despite the very low plasmalogen levels.

Other enzyme deficiencies with divergent clinical phenotypes include; primary

hyperoxaluria type I

106

, isolated glycolic aciduria

107

, sterol carrier protein X (SCPx)

deficiency

108

, 70-kDa peroxisomal membrane protein (PMP70) deficiency

45

, bile

acid-CoA:amino acid N-acyltransferase (BAAT) deficiency

109

and acatalasemia

110 111

, which will not be discussed in detail.

Mouse models

In order to study the pathogenic mechanism of peroxisomal dysfunction and the

consequences of different

PEX mutations and associated biochemical alterations,

several mouse models have been generated. Because this thesis is aimed at PBDs,

mice with defects in single peroxisomal enzymes will not be discussed here. For a

detailed review see

112113

. The main symptoms of mice with selective inactivation

of peroxisomes in several tissues (e.g. brain and liver);

Pex5-alfp

114

,

Pex5-albumin

115116117

,

Pex5-Nestin

114118

,

Pex5-NEX

119120

,

Pex5-gfap

119

,

Pex5-Cnp

121122123

,

Pex13-Nestin

124

,

Pex5-AMH

125

,

Pex5-aP2

126

, are summarized in table 2. Mouse

models for PBD, except the RCDP type 1

127

, will be introduced briefly. It should be

noted that the majority of mice had a severe phenotype with mice dying soon after

birth.

Mice with a defect in

PEX2 (global knockout) were created in different backgrounds

(i.e. C57BL/6 and Swiss Webster)

128129

. Marked embryonic lethality was reported

in the C57BL/6 background. When pups were born alive, they suffered from

growth retardation, hypotonia, impaired feeding and eventually death within 24

hours. Swiss mice, however, sporadically survived for approximately 1-2 weeks.

Biochemically, the most pronounced abnormalities were increased VLCFAs, altered

bile acid metabolism and decreased levels of DHA and plasmalogens. Both models

resemble the human ZS phenotype.

The

PEX5 mouse model (global knockout) was the first mouse model for ZSD and

also mimics the severe ZS phenotype

130

. Mice were born with intra-uterine growth

retardation, were severely hypotonic, and died shortly after birth. Biochemically

these mice had increased levels of VLCFAs and depletion of plasmalogens.

A mouse model with a defect in

PEX10 was created using N-ethyl-N-nitrosourea

(ENU) mutagenesis

131

. The majority of mice died within a few hours after birth,

but some mice did survive. These mice were severely growth retarded and had an

ataxic gait. Biochemical results, obtained from E18.5 and 2 adult mice, showed

significantly increased levels of the bile acid intermediates and low plasmalogens

at E18.5.

Two different models were made with a global knockout of

PEX11

α

. In the model

generated by Li et al

132

, all peroxisomal biomarkers in plasma were normal and

the same is true for the mouse model developed by Weng et al

133

. Only a reduced

number of peroxisomes in livers of

PEX11

α

mice were found. When stressed with

fibrates or a high fat diet, the phenotype, such as fatty liver, was more pronounced.

Mice with a defect in

PEX11

β

(global knockout) were born in a normal Mendelian

ratio but displayed neonatal lethality, hypotonia and growth retardation. Only a minor

increase in the hepatic VLCFAs levels and decreased brain plasmalogens were

reported

134

.

Mice with defective

PEX13 (global knockout) presented with severe hypotonia and

growth retardation and died soon after birth. No intra-uterine death was documented.

VLCFAs were elevated and plasmalogens were decreased

135

.

Similar to our model (see Chapter 4), another group created a mouse model with a

homozygous missense mutation in Pex1

136

, which is the most common mutation in

ZSD patients

137

and is associated with a relatively mild phenotype

138

with survival

into adulthood (see Chapter 7). Mice with this defect were born in normal Mendelian

ratios. Some survived into adulthood, but the majority died of a natural cause before

reaching adulthood. All mice were growth retarded. Biochemical abnormalities

comprised increased plasma levels of VLCFAs, DHCA and THCA, C26:0-lysoPC

in bloodspots and low levels of plasmalogens in bloodspots. Overall, this model

represents the milder end of the ZSD spectrum.

(8)

1

Mouse gene/model (loxP)

Cell type with per

oxisomal deficiency Main symptoms/pathology Biochemical abnormalities Remarks Liver Pex5 -alfp-cr e

Inactivation of hepatocytes during fetal development

Gr

owth r

etar

dation, cholestasis,

steatosis, steatorrhea and mild tremor

Not r

eported

Mitochondrial abnormalities, 80% died within 8 days

Pex5

-albumin-cr

e

Hepatocytes >1week postnatal

Gr

owth r

etar

dation untill 3

months of age. Hepatomegaly

, steatosis, fibr osis and car cinogenesis C24 BA, glycogen ↓, C 24

BA, phytanic-, pristanic-acid, lactate

Mitochondrial abnormalities, impair

ed de novo lipogenesis, PP AR α activation, perturbed carbohydrate metabolism Brain Pex5 -Nestin-cr e Neur ons, astr ocytes, oligodendr ocytes

Survival into adulthood. Motor and coor

dination deficits and

cognitive impairment, evolving in lethargy and death befor

e

the age of 6 months. Delays in neocortex and cer

ebellum

formation, impairment in the formation and maintenance of myelin, axonal degeneration

VLCF A ↑, plasmalogens ↓ Astr o- and micr ogliosis, accumulation of lipid dr oplets Pex5 -NEX-cr e For ebrain pr ojection neur ons No obvious symptoms None Pex5 -gfap-cr e Astr ocytes, pr ecursor cells of neur

ons and oligodendr

ocytes No obvious symptoms VLCF A ↑, plasmalogens ↓ Accumulation of lipid dr oplets Table 2

Overview of the main symptoms/pathology of mice with selective inactivation of per

oxisomes in several tissues

BA bile acids, VLCF

A very long-chain fatty acids

Mouse gene/model (loxP)

Cell type with per

oxisomal deficiency Main symptoms/pathology Biochemical abnormalities Remarks Pex5 -Cnp Oligodendr ocytes

Myelin and axonal abnormalities, late-onset progr

essive neur

odegenerative

disease, peripheral neur

opathy with muscle

weakness VLCF A ↑, plasmalogens ↓

91% died within 12 months, r

eactive gliosis and

neur oinflammation, paranodal swellings Pex13 -nestin-cr e Neur ons, astr ocytes, oligodendr ocytes Gr owth r etar dation, impair ed

reflex and motor behaviour development, abnormal cer

ebellar formation

Plasmalogens

70% died within 5 weeks, mitchondrial dysfunction, oxidative str

ess, glyosis

Testis Pex5

-AMH-cr

e

Sertoli

Testicular degeneration and infertility

Lipid dr

oplets

White adipose tissue Pex5-aP2 White adipose tissue and non-adipogenic tissues, such as chondr

ocytes, myocytes,

neur

ons and osteocytes.

Lower body weight, incr

eased

fat mass, r

educed lipolysis,

impair

ed adr

energic tone which

impacts on muscle function, impair

ed thermogenesis of

skeletal muscle when fasted, degr

ee of insulin r esistance Plasmalogens ↓ Table 2 Part 2 BA bile acids, VLCF

(9)

1

Outline of this thesis

This thesis focusses on translational studies in patients with a ZSD, with the aim to

discover and test new treatments and to create a new mouse model to study the

pathogenesis of the milder ZSD phenotype in more detail. The first part of the thesis,

Chapter 1, briefly introduces peroxisomes, peroxisomal biogenesis and function,

diseases associated with peroxisomal dysfunction and lastly existing PBD mouse

models are described. The second part of the introduction (Chapter 2) provides insight

in the history of ZSDs, different clinical phenotypes, diagnosis, clinical management

and therapeutic options. In general, no treatment for ZSDs exists and management

is mainly based on supportive therapy.

To accomplish the original goal of the thesis we tested several promising compounds

in skin fibroblasts of patients with a mild ZSD. When immunofluorescent staining

for the PTS-1 protein catalase is performed, these skin fibroblasts of mild patients

with a specific mutation in

PEX1, display a mixed population of cells either with

or without import competent peroxisomes, the so called “peroxisomal mosaicism”

139

. These skin fibroblasts regain catalase import competent peroxisomes when

cultured at 30°C, resembling healthy control cells lines. In contrast, when cells are

cultured at 40°C, all cells lose import competent peroxisomes. One compound (i.e.

arginine) was able to (partially) restore peroxisomal function in skin fibroblasts of

mild ZSD patients and was tested in more detail. Hence, detailed results of these

experiments are described in Chapter 3, of part II laboratory studies. As described in

Chapter 1, several mouse models for PBDs have been created. The majority of these

models, however, represent the severe end of the ZSD spectrum, whereas only

limited data is available on the first mouse model for mild ZSD patients. In Chapter 4,

the results of a comprehensive characterisation of our Pex1 mouse model, a model

that represents the milder end of the ZSD spectrum, is presented.

Cholic acid, marketed as Cholbam® in the United States, was recently approved by

the FDA for treatment in patients with a ZSD under a ‘rare paediatric disease priority

review voucher’. Until now, cholic acid therapy has not been studied in a large

cohort of ZSD patients and approval was based on case reports only. In Chapter

5, we report the results of the first large clinical trial of nine months cholic acid

supplementation in 19 patients with a ZSD. Briefly, our results indicate that cholic

acid does suppress the formation of bile acids in the majority of patients but it can

be harmful for patients suffering from severe liver disease.

ZSDs are generally associated with the severe classical ZS phenotype with patients

presenting symptoms such as hypotonia, prolonged jaundice, severe cognitive

impairment and death within the first year of life. In Chapter 6, part IV, we focus on

extension of the clinical phenotype with some important implications for diagnostics

and general knowledge for all those involved in the care of ZSD patients. We describe

the clinical phenotype of patients with a ZSD surviving into adulthood, and present

age specific symptoms such as peripheral neuropathy with or without pyramidal

signs. Moreover, we show that some patients with prolonged survival present an

insidiously progressive disease course, despite normalization of biomarkers for

peroxisomal disease measured in plasma and erythrocytes. The disease should be

considered as a slowly progressive disease with a subgroup of patients surviving

into adulthood and not solely as a devastating paediatric disease with death within

the first years of life. In addition, recent years many patients were extensively

examined during several visits at our outpatient clinic. During these visits we noted

that a subgroup of patients developed adrenal insufficiency. Therefore we started a

cohort study to determine the prevalence of adrenal insufficiency in (mild) ZSD. The

results of this study are described in Chapter 7. Finally, all findings are discussed and

future prospectives are provided in Chapter 8.

(10)

1

References

1. Gabaldón, T. Peroxisome diversity and evolution. Philos. Trans. R. Soc. Lond. B. Biol. Sci.

365, 765–73 (2010).

2. Wanders, R. J. A. & Waterham, H. R. Biochemistry of mammalian peroxisomes revisited.

Annu. Rev. Biochem. 75, 295–332 (2006).

3. De Duve, C. & Baudhuin, P. Peroxisomes (microbodies and related particles). Physiol. Rev. 46,

323–57 (1966).

4. Rhodin, J. Correlation of ultrastructural organization and function in normal and experimentally

changed proximal convoluted tubule cells of the mouse kidney [Ph.D. thesis]. (Aktiebolaget, Godvil, Stockholm, Sweden, 1954).

5. Huybrechts, S. J.,Van Veldhoven, P. P.,Brees, C.,Mannaerts, G. P.,Los, G. V & Fransen, M.

Peroxisome dynamics in cultured mammalian cells. Traffic 10, 1722–33 (2009).

6. Schrader, M.,Bonekamp, N. A. & Islinger, M. Fission and proliferation of peroxisomes.

Biochim. Biophys. Acta 1822, 1343–57 (2012).

7. Ezaki, J.,Kominami, E. & Ueno, T. Peroxisome degradation in mammals. IUBMB Life 63,

1001–8 (2011).

8. Smith, J. J. & Aitchison, J. D. Peroxisomes take shape. Nat. Rev. Mol. Cell Biol. 14, 803–17

(2013).

9. Fagarasanu, A.,Mast, F. D.,Knoblach, B. & Rachubinski, R. A. Molecular mechanisms of

organelle inheritance: lessons from peroxisomes in yeast. Nat. Rev. Mol. Cell Biol. 11, 644–54

(2010).

10. Fujiki, Y.,Okumoto, K.,Mukai, S.,Honsho, M. & Tamura, S. Peroxisome biogenesis in

mammalian cells. Front. Physiol. 5, 307 (2014).

11. Pinto, M. P.,Grou, C. P.,Fransen, M.,Sá-Miranda, C. & Azevedo, J. E. The cytosolic domain of

PEX3, a protein involved in the biogenesis of peroxisomes, binds membrane lipids. Biochim.

Biophys. Acta 1793, 1669–75 (2009).

12. Waterham, H. R. & Ebberink, M. S. Genetics and molecular basis of human peroxisome

biogenesis disorders. Biochim. Biophys. Acta - Mol. Basis Dis. (2012).

13. Subramani, S. Targeting of proteins into the peroxisomal matrix. J. Membr. Biol. 125, 99–106

(1992).

14. Lazarow, P. B. The import receptor Pex7p and the PTS2 targeting sequence. Biochim.

Biophys. Acta 1763, 1599–604 (2006).

15. Braverman, N.,Dodt, G.,Gould, S. J. & Valle, D. An isoform of pex5p, the human PTS1

receptor, is required for the import of PTS2 proteins into peroxisomes. Hum. Mol. Genet. 7,

1195–205 (1998).

16. Dodt, G.,Warren, D.,Becker, E.,Rehling, P. & Gould, S. J. Domain mapping of human PEX5

reveals functional and structural similarities to Saccharomyces cerevisiae Pex18p and

Pex21p. J. Biol. Chem. 276, 41769–81 (2001).

17. Matsumura, T.,Otera, H. & Fujiki, Y. Disruption of the interaction of the longer isoform of

Pex5p, Pex5pL, with Pex7p abolishes peroxisome targeting signal type 2 protein import in

mammals. Study with a novel Pex5-impaired Chinese hamster ovary cell mutant. J. Biol.

Chem. 275, 21715–21 (2000).

18. Dammai, V. & Subramani, S. The human peroxisomal targeting signal receptor, Pex5p, is

translocated into the peroxisomal matrix and recycled to the cytosol. Cell 105, 187–96 (2001).

19. Dodt, G. Multiple PEX genes are required for proper subcellular distribution and stability

of Pex5p, the PTS1 receptor: evidence that PTS1 protein import is mediated by a cycling

receptor. J. Cell Biol. 135, 1763–1774 (1996).

20. Azevedo, J. E.,Costa-Rodrigues, J.,Guimarães, C. P.,Oliveira, M. E. & Sã-Miranda, C. Protein

translocation across the peroxisomal membrane. Cell Biochem. Biophys. 41, 451–68 (2004).

21. Montilla-Martinez, M.,Beck, S.,Klümper, J.,Meinecke, M.,Schliebs, W.,Wagner, R. & Erdmann,

R. Distinct Pores for Peroxisomal Import of PTS1 and PTS2 Proteins. Cell Rep. 13, 2126–34

(2015).

22. Grou, C. P.,Carvalho, A. F.,Pinto, M. P.,Alencastre, I. S.,Rodrigues, T. A.,Freitas, M.

O.,Francisco, T.,Sá-Miranda, C. & Azevedo, J. E. The peroxisomal protein import

machinery--a case report of transient ubiquitination with a new flavor. Cell. Mol. Life Sci. 66, 254–62

(2009).

23. Kim, P. K.,Hailey, D. W.,Mullen, R. T. & Lippincott-Schwartz, J. Ubiquitin signals autophagic

degradation of cytosolic proteins and peroxisomes. Proc. Natl. Acad. Sci. U. S. A. 105,

20567–74 (2008).

24. Manjithaya, R.,Nazarko, T. Y.,Farré, J.-C. & Subramani, S. Molecular mechanism and

physiological role of pexophagy. FEBS Lett. 584, 1367–73 (2010).

25. Gouveia, A. M.,Guimarães, C. P.,Oliveira, M. E.,Reguenga, C.,Sá-Miranda, C. & Azevedo,

J. E. Characterization of the peroxisomal cycling receptor Pex5p import pathway. Adv. Exp.

Med. Biol. 544, 219–20 (2003).

26. Waterham, H. R.,Ferdinandusse, S. & Wanders, R. J. A. Human disorders of peroxisome

metabolism and biogenesis. Biochim. Biophys. Acta 1863, 922–33 (2016).

27. Hettema, E. H.,Erdmann, R.,van der Klei, I. & Veenhuis, M. Evolving models for peroxisome

biogenesis. Curr. Opin. Cell Biol. 29, 25–30 (2014).

28. Tabak, H. F.,Braakman, I. & van der Zand, A. Peroxisome formation and maintenance are

dependent on the endoplasmic reticulum. Annu. Rev. Biochem. 82, 723–44 (2013).

29. Fujiki, Y.,Okumoto, K.,Mukai, S.,Honsho, M. & Tamura, S. Peroxisome biogenesis in

mammalian cells. Front. Physiol. 5, 307 (2014).

30. Koch, A.,Thiemann, M.,Grabenbauer, M.,Yoon, Y.,McNiven, M. A. & Schrader, M.

Dynamin-like protein 1 is involved in peroxisomal fission. J. Biol. Chem. 278, 8597–605 (2003).

31. Koch, A.,Yoon, Y.,Bonekamp, N. A.,McNiven, M. A. & Schrader, M. A role for Fis1 in both

mitochondrial and peroxisomal fission in mammalian cells. Mol. Biol. Cell 16, 5077–86 (2005).

32. Gandre-Babbe, S. & van der Bliek, A. M. The novel tail-anchored membrane protein Mff

controls mitochondrial and peroxisomal fission in mammalian cells. Mol. Biol. Cell 19, 2402–

12 (2008).

33. Huber, N.,Guimaraes, S.,Schrader, M.,Suter, U. & Niemann, A. Charcot-Marie-Tooth

disease-associated mutants of GDAP1 dissociate its roles in peroxisomal and mitochondrial fission.

EMBO Rep. 14, 545–52 (2013).

34. Rokka, A.,Antonenkov, V. D.,Soininen, R.,Immonen, H. L.,Pirilä, P. L.,Bergmann,

U.,Sormunen, R. T.,Weckström, M.,Benz, R. & Hiltunen, J. K. Pxmp2 is a channel-forming

protein in Mammalian peroxisomal membrane. PLoS One 4, e5090 (2009).

35. Visser, W. F.,van Roermund, C. W. T.,Ijlst, L.,Waterham, H. R. & Wanders, R. J. A. Metabolite

transport across the peroxisomal membrane. Biochem. J. 401, 365–75 (2007).

36. Kemp, S.,Theodoulou, F. L. & Wanders, R. J. A. Mammalian peroxisomal ABC transporters:

from endogenous substrates to pathology and clinical significance. Br. J. Pharmacol. 164,

1753–66 (2011).

37. Morita, M. & Imanaka, T. Peroxisomal ABC transporters: Structure, function and role in

disease. Biochim. Biophys. Acta 1822, 1387–96 (2012).

38. Mosser, J.,Douar, A. M.,Sarde, C. O.,Kioschis, P.,Feil, R.,Moser, H.,Poustka, A. M.,Mandel, J.

L. & Aubourg, P. Putative X-linked adrenoleukodystrophy gene shares unexpected homology

with ABC transporters. Nature 361, 726–30 (1993).

39. Holzinger, A.,Kammerer, S.,Berger, J. & Roscher, A. A. cDNA cloning and mRNA expression

of the human adrenoleukodystrophy related protein (ALDRP), a peroxisomal ABC transporter.

Biochem. Biophys. Res. Commun. 239, 261–4 (1997).

40. Kamijo, K.,Kamijo, T.,Ueno, I.,Osumi, T. & Hashimoto, T. Nucleotide sequence of the human

70 kDa peroxisomal membrane protein: a member of ATP-binding cassette transporters.

Biochim. Biophys. Acta 1129, 323–7 (1992).

41. Wiesinger, C.,Kunze, M.,Regelsberger, G.,Forss-Petter, S. & Berger, J. Impaired very

long-chain acyl-CoA β-oxidation in human X-linked adrenoleukodystrophy fibroblasts is a direct

consequence of ABCD1 transporter dysfunction. J. Biol. Chem. 288, 19269–79 (2013).

42. van Roermund, C. W. T.,Visser, W. F.,Ijlst, L.,van Cruchten, A.,Boek, M.,Kulik, W.,Waterham,

H. R. & Wanders, R. J. A. The human peroxisomal ABC half transporter ALDP functions as a

(11)

1

43. Fourcade, S.,Ruiz, M.,Camps, C.,Schlüter, A.,Houten, S. M.,Mooyer, P. A. W.,Pàmpols,

T.,Dacremont, G.,Wanders, R. J. A.,Giròs, M. & Pujol, A. A key role for the peroxisomal

ABCD2 transporter in fatty acid homeostasis. Am. J. Physiol. Endocrinol. Metab. 296, E211–

21 (2009).

44. Pujol, A.,Ferrer, I.,Camps, C.,Metzger, E.,Hindelang, C.,Callizot, N.,Ruiz, M.,Pàmpols,

T.,Giròs, M. & Mandel, J. L. Functional overlap between ABCD1 (ALD) and ABCD2 (ALDR)

transporters: a therapeutic target for X-adrenoleukodystrophy. Hum. Mol. Genet. 13, 2997–

3006 (2004).

45. Ferdinandusse, S.,Jimenez-Sanchez, G.,Koster, J.,Denis, S.,Van Roermund, C.

W.,Silva-Zolezzi, I.,Moser, A. B.,Visser, W. F.,Gulluoglu, M.,Durmaz, O.,Demirkol, M.,Waterham, H. R.,Gökcay, G.,Wanders, R. J. A. & Valle, D. A novel bile acid biosynthesis defect due to a

deficiency of peroxisomal ABCD3. Hum. Mol. Genet. 24, 361–70 (2015).

46. Verhoeven, N. M. & Jakobs, C. Human metabolism of phytanic acid and pristanic acid. Prog.

Lipid Res. 40, 453–66 (2001).

47. Wanders, R. J. A.,van Roermund, C. W. T.,Visser, W. F.,Ferdinandusse, S.,Jansen, G. A.,van

den Brink, D. M.,Gloerich, J. & Waterham, H. R. Peroxisomal fatty acid alpha- and

beta-oxidation in health and disease: new insights. Adv. Exp. Med. Biol. 544, 293–302 (2003).

48. Casteels, M.,Foulon, V.,Mannaerts, G. P. & Van Veldhoven, P. P. Alpha-oxidation of

3-methyl-substituted fatty acids and its thiamine dependence. Eur. J. Biochem. 270, 1619–27 (2003).

49. Wanders, R. J. A.,Vreken, P.,Ferdinandusse, S.,Jansen, G. A.,Waterham, H. R.,van

Roermund, C. W. & Van Grunsven, E. G. Peroxisomal fatty acid alpha- and beta-oxidation in humans: enzymology, peroxisomal metabolite transporters and peroxisomal diseases.

Biochem. Soc. Trans. 29, 250–67 (2001).

50. Horrocks, L. A. & Yeo, Y. K. Health benefits of docosahexaenoic acid (DHA). Pharmacol. Res.

40, 211–25 (1999).

51. Birch, E. E.,Hoffman, D. R.,Uauy, R.,Birch, D. G. & Prestidge, C. Visual acuity and the

essentiality of docosahexaenoic acid and arachidonic acid in the diet of term infants. Pediatr.

Res. 44, 201–9 (1998).

52. Bazinet, R. P. & Layé, S. Polyunsaturated fatty acids and their metabolites in brain function

and disease. Nat. Rev. Neurosci. 15, 771–785 (2014).

53. Ferdinandusse, S.,Denis, S.,Mooijer, P. a,Zhang, Z.,Reddy, J. K.,Spector, a a & Wanders,

R. J. Identification of the peroxisomal beta-oxidation enzymes involved in the biosynthesis of

docosahexaenoic acid. J. Lipid Res. 42, 1987–95 (2001).

54. Ferdinandusse, S. & Houten, S. M. Peroxisomes and bile acid biosynthesis. Biochim.

Biophys. Acta 1763, 1427–40 (2006).

55. Pircher, P. C.,Kitto, J. L.,Petrowski, M. L.,Tangirala, R. K.,Bischoff, E. D.,Schulman, I. G. &

Westin, S. K. Farnesoid X receptor regulates bile acid-amino acid conjugation. J. Biol. Chem.

278, 27703–11 (2003).

56. Salido, E.,Pey, A. L.,Rodriguez, R. & Lorenzo, V. Primary hyperoxalurias: disorders of

glyoxylate detoxification. Biochim. Biophys. Acta 1822, 1453–64 (2012).

57. Braverman, N. E. & Moser, A. B. Functions of plasmalogen lipids in health and disease.

Biochim. Biophys. Acta 1822, 1442–52 (2012).

58. Zoeller, R. A.,Lake, A. C.,Nagan, N.,Gaposchkin, D. P.,Legner, M. A. & Lieberthal, W.

Plasmalogens as endogenous antioxidants: somatic cell mutants reveal the importance of the

vinyl ether. Biochem. J. 338 ( Pt 3, 769–76 (1999).

59. Akoh, C. C. & Min, D. B. in Food Lipids: Chemistry, Nutrition, and Biotechnology, Third

Edition, 39–55 (2008).

60. Brites, P.,Waterham, H. R. & Wanders, R. J. A. Functions and biosynthesis of plasmalogens in

health and disease. Biochim. Biophys. Acta 1636, 219–31 (2004).

61. Fransen, M.,Nordgren, M.,Wang, B. & Apanasets, O. Role of peroxisomes in

ROS/RNS-metabolism: implications for human disease. Biochim. Biophys. Acta 1822, 1363–73 (2012).

62. Wanders, R. J. A. Peroxisomes in human health and disease: metabolic pathways, metabolite

transport, interplay with other organelles and signal transduction. Subcell. Biochem. 69,

23–44 (2013).

63. Schrader, M. & Yoon, Y. Mitochondria and peroxisomes: are the ‘big brother’ and the ‘little

sister’ closer than assumed? Bioessays 29, 1105–14 (2007).

64. Lismont, C.,Nordgren, M.,Van Veldhoven, P. P. & Fransen, M. Redox interplay between

mitochondria and peroxisomes. Front. cell Dev. Biol. 3, 35 (2015).

65. Demarquoy, J. & Le Borgne, F. Crosstalk between mitochondria and peroxisomes. World J.

Biol. Chem. 6, 301–9 (2015).

66. Waterham, H. R.,Koster, J.,van Roermund, C. W. T.,Mooyer, P. a W.,Wanders, R. J. A. &

Leonard, J. V. A lethal defect of mitochondrial and peroxisomal fission. N. Engl. J. Med. 356,

1736–41 (2007).

67. Ashmarina, L. I.,Robert, M. F.,Elsliger, M. A. & Mitchell, G. A. Characterization of the

hydroxymethylglutaryl-CoA lyase precursor, a protein targeted to peroxisomes and

mitochondria. Biochem. J. 315, 71–75 (1996).

68. Schmitz, W.,Albers, C.,Fingerhut, R. & Conzelmann, E. Purification and Characterization of an

alpha-Methylacyl-CoA Racemase from Human Liver. Eur. J. Biochem. 231, 815–822 (1995).

69. Goldfischer, S.,Moore, C. L.,Johnson, A. B.,Spiro, A. J.,Valsamis, M. P.,Wisniewski, H.

K.,Ritch, R. H.,Norton, W. T.,Rapin, I. & Gartner, L. M. Peroxisomal and Mitochondrial Defects

in the Cerebro-Hepato-Renal Syndrome. Science. 182, 62–64 (1973).

70. Baumgart, E.,Vanhorebeek, I.,Grabenbauer, M.,Borgers, M.,Declercq, P. E.,Fahimi, H. D. &

Baes, M. Mitochondrial alterations caused by defective peroxisomal biogenesis in a mouse

model for Zellweger syndrome (PEX5 knockout mouse). Am. J. Pathol. 159, 1477–94 (2001).

71. Salpietro, V.,Phadke, R.,Saggar, A.,Hargreaves, I. P.,Yates, R.,Fokoloros, C.,Mankad,

K.,Hertecant, J.,Ruggieri, M.,McCormick, D. & Kinali, M. Zellweger syndrome and secondary

mitochondrial myopathy. Eur. J. Pediatr. 174, 557–63 (2015).

72. Shamseldin, H. E.,Alshammari, M.,Al-Sheddi, T.,Salih, M. A.,Alkhalidi, H.,Kentab, A.,Repetto,

G. M.,Hashem, M. & Alkuraya, F. S. Genomic analysis of mitochondrial diseases in a

consanguineous population reveals novel candidate disease genes. J. Med. Genet. 49, 234–

41 (2012).

73. Ebberink, M. S.,Koster, J.,Visser, G.,Spronsen, F. Van,Stolte-Dijkstra, I.,Smit, G. P. a,Fock,

J. M.,Kemp, S.,Wanders, R. J. a & Waterham, H. R. A novel defect of peroxisome division

due to a homozygous non-sense mutation in the PEX11β gene. J. Med. Genet. 49, 307–13

(2012).

74. Ratbi, I.,Falkenberg, K. D.,Sommen, M.,Al-Sheqaih, N.,Guaoua, S.,Vandeweyer, G.,Urquhart,

J. E.,Chandler, K. E.,Williams, S. G.,Roberts, N. A.,El Alloussi, M.,Black, G. C.,Ferdinandusse, S.,Ramdi, H.,Heimler, A.,Fryer, A.,Lynch, S.-A.,Cooper, N.,Ong, K. R.,Smith, C. E.

L.,Inglehearn, C. F.,Mighell, A. J.,Elcock, C.,Poulter, J. A.,Tischkowitz, M.,Davies, S. J.,Sefiani, A.,Mironov, A. A.,Newman, W. G.,Waterham, H. R. & Van Camp, G. Heimler Syndrome Is Caused by Hypomorphic Mutations in the Peroxisome-Biogenesis Genes PEX1 and PEX6.

Am. J. Hum. Genet. 97, 535–545 (2015).

75. Poll-The, B. T.,Saudubray, J. M.,Ogier, H. a,Odièvre, M.,Scotto, J. M.,Monnens,

L.,Govaerts, L. C.,Roels, F.,Cornelis, a & Schutgens, R. B. Infantile Refsum disease: an inherited peroxisomal disorder. Comparison with Zellweger syndrome and neonatal

adrenoleukodystrophy. Eur. J. Pediatr. 146, 477–83 (1987).

76. Kelley, R. I.,Datta, N. S.,Dobyns, W. B.,Hajra, A. K.,Moser, A. B.,Noetzel, M. J.,Zackai, E.

H. & Moser, H. W. Neonatal adrenoleukodystrophy: new cases, biochemical studies, and

differentiation from Zellweger and related peroxisomal polydystrophy syndromes. Am. J. Med.

Genet. 23, 869–901 (1986).

77. Bowen, P.,Lee, C. S.,Zellweger, H. & Lindenberg, R. A familial syndrome of multiple congenital

defects. Bull. Johns Hopkins Hosp. 114, 402–14 (1964).

78. Braverman, N. E.,D’Agostino, M. D. & Maclean, G. E. Peroxisome biogenesis disorders:

Biological, clinical and pathophysiological perspectives. Dev. Disabil. Res. Rev. 17, 187–96

(2013).

79. White, A. L.,Modaff, P.,Holland-Morris, F. & Pauli, R. M. Natural history of rhizomelic

(12)

1

80. Braverman, N.,Chen, L.,Lin, P.,Obie, C.,Steel, G.,Douglas, P.,Chakraborty, P. K.,Clarke, J. T.

R.,Boneh, A.,Moser, A.,Moser, H. & Valle, D. Mutation analysis of PEX7 in 60 probands with rhizomelic chondrodysplasia punctata and functional correlations of genotype with phenotype.

Hum. Mutat. 20, 284–97 (2002).

81. Braverman, N. E.,Moser, A. B. & Steinberg, S. J. Rhizomelic Chondrodysplasia Punctata

Type 1. 2001 Nov 16 [Updated 2012 Sep 12]. In Pagon RA, Adam MP, Ardinger HH, et al.,

editors. GeneReviews® [Internet]. Seattle (WA): University of Washington, Seattle; 1993-2016 Available at: http://www.ncbi.nlm.nih.gov/books/NBK1270/.

82. Engelen, M.,Kemp, S.,de Visser, M.,van Geel, B. M.,Wanders, R. J. a,Aubourg, P. &

Poll-The, B. T. X-linked adrenoleukodystrophy (X-ALD): clinical presentation and guidelines for

diagnosis, follow-up and management. Orphanet J. Rare Dis. 7, 51 (2012).

83. Engelen, M.,Barbier, M.,Dijkstra, I. M. E.,Schür, R.,de Bie, R. M. A.,Verhamme, C.,Dijkgraaf,

M. G. W.,Aubourg, P. A.,Wanders, R. J. A.,van Geel, B. M.,de Visser, M.,Poll-The, B. T. &

Kemp, S. X-linked adrenoleukodystrophy in women: a cross-sectional cohort study. Brain

137, 693–706 (2014).

84. Mosser, J.,Douar, A. M.,Sarde, C. O.,Kioschis, P.,Feil, R.,Moser, H.,Poustka, A. M.,Mandel, J.

L. & Aubourg, P. Putative X-linked adrenoleukodystrophy gene shares unexpected homology

with ABC transporters. Nature 361, 726–30 (1993).

85. Vogel, B. H.,Bradley, S. E.,Adams, D. J.,D’Aco, K.,Erbe, R. W.,Fong, C.,Iglesias, A.,Kronn,

D.,Levy, P.,Morrissey, M.,Orsini, J.,Parton, P.,Pellegrino, J.,Saavedra-Matiz, C. A.,Shur, N.,Wasserstein, M.,Raymond, G. V & Caggana, M. Newborn screening for X-linked adrenoleukodystrophy in New York State: Diagnostic protocol, surveillance protocol and

treatment guidelines. Mol. Genet. Metab. 114, 599–603 (2015).

86. Poll-The, B. T.,Roels, F.,Ogier, H.,Scotto, J.,Vamecq, J.,Schutgens, R. B.,Wanders, R.

J.,van Roermund, C. W.,van Wijland, M. J. & Schram, A. W. A new peroxisomal disorder with enlarged peroxisomes and a specific deficiency of acyl-CoA oxidase (pseudo-neonatal

adrenoleukodystrophy). Am. J. Hum. Genet. 42, 422–34 (1988).

87. Ferdinandusse, S.,Denis, S.,Hogenhout, E. M.,Koster, J.,van Roermund, C. W. T.,IJlst,

L.,Moser, A. B.,Wanders, R. J. A. & Waterham, H. R. Clinical, biochemical, and mutational

spectrum of peroxisomal acyl-coenzyme A oxidase deficiency. Hum. Mutat. 28, 904–12

(2007).

88. Ferdinandusse, S.,Barker, S.,Lachlan, K.,Duran, M.,Waterham, H. R.,Wanders, R. J. A. &

Hammans, S. Adult peroxisomal acyl-coenzyme A oxidase deficiency with cerebellar and

brainstem atrophy. J. Neurol. Neurosurg. Psychiatry 81, 310–2 (2010).

89. Ferdinandusse, S.,Denis, S.,IJlst, L.,Dacremont, G.,Waterham, H. R. & Wanders, R. J.

Subcellular localization and physiological role of alpha-methylacyl-CoA racemase. J. Lipid

Res. 41, 1890–6 (2000).

90. Ferdinandusse, S.,Denis, S.,Clayton, P. T.,Graham, A.,Rees, J. E.,Allen, J. T.,McLean, B.

N.,Brown, A. Y.,Vreken, P.,Waterham, H. R. & Wanders, R. J. Mutations in the gene encoding peroxisomal alpha-methylacyl-CoA racemase cause adult-onset sensory motor neuropathy.

Nat. Genet. 24, 188–91 (2000).

91. Clarke, C. E.,Alger, S.,Preece, M. A.,Burdon, M. A.,Chavda, S.,Denis, S.,Ferdinandusse, S. &

Wanders, R. J. A. Tremor and deep white matter changes in alpha-methylacyl-CoA racemase

deficiency. Neurology 63, 188–9 (2004).

92. Thompson, S. A.,Calvin, J.,Hogg, S.,Ferdinandusse, S.,Wanders, R. J. A. & Barker, R. A.

Relapsing encephalopathy in a patient with α-methylacyl-CoA racemase deficiency. BMJ

Case Rep. 2009, (2009).

93. Smith, E. H.,Gavrilov, D. K.,Oglesbee, D.,Freeman, W. D.,Vavra, M. W.,Matern, D. & Tortorelli,

S. An adult onset case of alpha-methyl-acyl-CoA racemase deficiency. J. Inherit. Metab. Dis.

33 Suppl 3, S349–53 (2010).

94. Ferdinandusse, S.,Denis, S.,Mooyer, P. a W.,Dekker, C.,Duran, M.,Soorani-Lunsing, R.

J.,Boltshauser, E.,Macaya, A.,Gärtner, J.,Majoie, C. B. L. M.,Barth, P. G.,Wanders, R. J. A.

& Poll-The, B. T. Clinical and biochemical spectrum of D-bifunctional protein deficiency. Ann.

Neurol. 59, 92–104 (2006).

95. Lines, M. A.,Jobling, R.,Brady, L.,Marshall, C. R.,Scherer, S. W.,Rodriguez, A. R.,Lee, L.,Lang,

A. E.,Mestre, T. A.,Wanders, R. J. A.,Ferdinandusse, S. & Tarnopolsky, M. A. Peroxisomal D-bifunctional protein deficiency: Three adults diagnosed by whole-exome sequencing.

Neurology (2014).

96. Lieber, D. S.,Hershman, S. G.,Slate, N. G.,Calvo, S. E.,Sims, K. B.,Schmahmann, J. D. &

Mootha, V. K. Next generation sequencing with copy number variant detection expands the

phenotypic spectrum of HSD17B4-deficiency. BMC Med. Genet. 15, 30 (2014).

97. Jansen, G. A.,Ofman, R.,Ferdinandusse, S.,Ijlst, L.,Muijsers, A. O.,Skjeldal, O. H.,Stokke,

O.,Jakobs, C.,Besley, G. T.,Wraith, J. E. & Wanders, R. J. A. Refsum disease is caused by

mutations in the phytanoyl-CoA hydroxylase gene. Nat. Genet. 17, 190–3 (1997).

98. Jansen, G. A.,Waterham, H. R. & Wanders, R. J. A. Molecular basis of Refsum disease:

sequence variations in phytanoyl-CoA hydroxylase (PHYH) and the PTS2 receptor (PEX7).

Hum. Mutat. 23, 209–18 (2004).

99. Van den Brink, D. M.,Brites, P.,Haasjes, J.,Wierzbicki, A. S.,Mitchell, J.,Lambert-Hamill, M.,de

Belleroche, J.,Jansen, G. A.,Waterham, H. R. & Wanders, R. J. A. Identification of PEX7 as

the second gene involved in Refsum disease. Adv. Exp. Med. Biol. 544, 69–70 (2003).

100. Wanders, R. J. A.,Komen, J. & Ferdinandusse, S. Phytanic acid metabolism in health and

disease. Biochim. Biophys. Acta 1811, 498–507 (2011).

101. Wanders, R. J. A.,Waterham, H. R. & Leroy, B. P. Refsum disease. 2006 Mar 20 [Updated

2015 Jun 11]. In Pagon RA, Adam MP, Ardinger HH, et al., editors. GeneReviews® [Internet]. Seattle (WA): University of Washington, Seattle; 1993-2016 Available at: http://

www.ncbi.nlm.nih.gov/books/NBK1353/.

102. Ofman, R.,Hettema, E. H.,Hogenhout, E. M.,Caruso, U.,Muijsers, A. O. & Wanders, R. J.

A. Acyl-CoA:dihydroxyacetonephosphate acyltransferase: cloning of the human cDNA and

resolution of the molecular basis in rhizomelic chondrodysplasia punctata type 2. Hum. Mol.

Genet. 7, 847–53 (1998).

103. de Vet, E. C.,Ijlst, L.,Oostheim, W.,Wanders, R. J. A. & van den Bosch, H.

Alkyl-dihydroxyacetonephosphate synthase. Fate in peroxisome biogenesis disorders and

identification of the point mutation underlying a single enzyme deficiency. J. Biol. Chem. 273,

10296–301 (1998).

104. Buchert, R.,Tawamie, H.,Smith, C.,Uebe, S.,Innes, A. M.,Al Hallak, B.,Ekici, A. B.,Sticht,

H.,Schwarze, B.,Lamont, R. E.,Parboosingh, J. S.,Bernier, F. P. & Abou Jamra, R. A peroxisomal disorder of severe intellectual disability, epilepsy, and cataracts due to fatty

acyl-CoA reductase 1 deficiency. Am. J. Hum. Genet. 95, 602–10 (2014).

105. Wanders, R. J. A. & Poll-The, B. T. ‘Role of peroxisomes in human lipid metabolism and its

importance for neurological development’. Neurosci. Lett. S0304-3940, 00461–9 [Epub ahead

of print] (2015).

106. Danpure, C. J. & Jennings, P. R. Peroxisomal alanine:glyoxylate aminotransferase deficiency

in primary hyperoxaluria type I. FEBS Lett. 201, 20–4 (1986).

107. Frishberg, Y.,Zeharia, A.,Lyakhovetsky, R.,Bargal, R. & Belostotsky, R. Mutations in HAO1

encoding glycolate oxidase cause isolated glycolic aciduria. J. Med. Genet. 51, 526–9 (2014).

108. Ferdinandusse, S.,Kostopoulos, P.,Denis, S.,Rusch, H.,Overmars, H.,Dillmann, U.,Reith,

W.,Haas, D.,Wanders, R. J. A.,Duran, M. & Marziniak, M. Mutations in the gene encoding peroxisomal sterol carrier protein X (SCPx) cause leukencephalopathy with dystonia and

motor neuropathy. Am. J. Hum. Genet. 78, 1046–52 (2006).

109. Hadžić, N.,Bull, L. N.,Clayton, P. T. & Knisely, A. S. Diagnosis in bile acid-CoA: amino acid

N-acyltransferase deficiency. World J. Gastroenterol. 18, 3322–6 (2012).

110. Takahara, S. Progressive oral gangrene probably due to lack of catalase in the blood

(acatalasaemia); report of nine cases. Lancet 2, 1101–4 (1952).

111. Góth, L. & Nagy, T. Inherited catalase deficiency: is it benign or a factor in various age related

disorders? Mutat. Res. 753, 147–54 (2013).

112. Baes, M. & Van Veldhoven, P. P. Mouse models for peroxisome biogenesis defects and

β-oxidation enzyme deficiencies. Biochim. Biophys. Acta 1822, 1489–500 (2012).

113. Baes, M. & Van Veldhoven, P. P. Generalised and conditional inactivation of Pex genes in

(13)

1

114. Krysko, O.,Hulshagen, L.,Janssen, A.,Schütz, G.,Klein, R.,De Bruycker, M.,Espeel,

M.,Gressens, P. & Baes, M. Neocortical and cerebellar developmental abnormalities in

conditions of selective elimination of peroxisomes from brain or from liver. J. Neurosci. Res.

85, 58–72 (2007).

115. Dirkx, R.,Vanhorebeek, I.,Martens, K.,Schad, A.,Grabenbauer, M.,Fahimi, D.,Declercq,

P.,Van Veldhoven, P. P. & Baes, M. Absence of peroxisomes in mouse hepatocytes causes

mitochondrial and ER abnormalities. Hepatology 41, 868–78 (2005).

116. Peeters, A.,Swinnen, J. V,Van Veldhoven, P. P. & Baes, M. Hepatosteatosis in peroxisome

deficient liver despite increased β-oxidation capacity and impaired lipogenesis. Biochimie 93,

1828–38 (2011).

117. Peeters, A.,Fraisl, P.,van den Berg, S.,Ver Loren van Themaat, E.,Van Kampen, A.,Rider, M.

H.,Takemori, H.,van Dijk, K. W.,Van Veldhoven, P. P.,Carmeliet, P. & Baes, M. Carbohydrate metabolism is perturbed in peroxisome-deficient hepatocytes due to mitochondrial dysfunction, AMP-activated protein kinase (AMPK) activation, and peroxisome

proliferator-activated receptor γ coactivator 1α (PGC-1α) suppression. J. Biol. Chem. 286, 42162–79

(2011).

118. Hulshagen, L.,Krysko, O.,Bottelbergs, A.,Huyghe, S.,Klein, R.,Van Veldhoven, P. P.,De Deyn,

P. P.,D’Hooge, R.,Hartmann, D. & Baes, M. Absence of functional peroxisomes from mouse

CNS causes dysmyelination and axon degeneration. J. Neurosci. 28, 4015–27 (2008).

119. Bottelbergs, A.,Verheijden, S.,Hulshagen, L.,Gutmann, D. H.,Goebbels, S.,Nave,

K.-A.,Kassmann, C. & Baes, M. Axonal integrity in the absence of functional peroxisomes from

projection neurons and astrocytes. Glia 58, 1532–43 (2010).

120. Goebbels, S.,Bormuth, I.,Bode, U.,Hermanson, O.,Schwab, M. H. & Nave, K.-A. Genetic

targeting of principal neurons in neocortex and hippocampus of NEX-Cre mice. Genesis 44,

611–21 (2006).

121. Kassmann, C. M.,Lappe-Siefke, C.,Baes, M.,Brügger, B.,Mildner, A.,Werner, H. B.,Natt,

O.,Michaelis, T.,Prinz, M.,Frahm, J. & Nave, K.-A. Axonal loss and neuroinflammation caused

by peroxisome-deficient oligodendrocytes. Nat. Genet. 39, 969–76 (2007).

122. Kassmann, C. M.,Quintes, S.,Rietdorf, J.,Möbius, W.,Sereda, M. W.,Nientiedt, T.,Saher,

G.,Baes, M. & Nave, K.-A. A role for myelin-associated peroxisomes in maintaining paranodal

loops and axonal integrity. FEBS Lett. 585, 2205–11 (2011).

123. Ahlemeyer, B.,Neubert, I.,Kovacs, W. J. & Baumgart-Vogt, E. Differential expression of

peroxisomal matrix and membrane proteins during postnatal development of mouse brain. J.

Comp. Neurol. 505, 1–17 (2007).

124. Müller, C. C.,Nguyen, T. H.,Ahlemeyer, B.,Meshram, M.,Santrampurwala, N.,Cao, S.,Sharp,

P.,Fietz, P. B.,Baumgart-Vogt, E. & Crane, D. I. PEX13 deficiency in mouse brain as a model of Zellweger syndrome: abnormal cerebellum formation, reactive gliosis and oxidative stress.

Dis. Model. Mech. 4, 104–19 (2011).

125. Huyghe, S.,Schmalbruch, H.,De Gendt, K.,Verhoeven, G.,Guillou, F.,Van Veldhoven, P. P. &

Baes, M. Peroxisomal multifunctional protein 2 is essential for lipid homeostasis in Sertoli cells

and male fertility in mice. Endocrinology 147, 2228–36 (2006).

126. Martens, K.,Bottelbergs, A.,Peeters, A.,Jacobs, F.,Espeel, M.,Carmeliet, P.,Van Veldhoven,

P. P. & Baes, M. Peroxisome deficient aP2-Pex5 knockout mice display impaired white

adipocyte and muscle function concomitant with reduced adrenergic tone. Mol. Genet.

Metab. 107, 735–47 (2012).

127. Brites, P.,Motley, A. M.,Gressens, P.,Mooyer, P. a W.,Ploegaert, I.,Everts, V.,Evrard,

P.,Carmeliet, P.,Dewerchin, M.,Schoonjans, L.,Duran, M.,Waterham, H. R.,Wanders, R. J. A. & Baes, M. Impaired neuronal migration and endochondral ossification in Pex7 knockout mice:

a model for rhizomelic chondrodysplasia punctata. Hum. Mol. Genet. 12, 2255–67 (2003).

128. Faust, P. L. Abnormal cerebellar histogenesis in PEX2 Zellweger mice reflects multiple

neuronal defects induced by peroxisome deficiency. J. Comp. Neurol. 461, 394–413 (2003).

129. Faust, P. L. & Hatten, M. E. Targeted deletion of the PEX2 peroxisome assembly gene in mice

provides a model for Zellweger syndrome, a human neuronal migration disorder. J. Cell Biol.

139, 1293–305 (1997).

130. Baes, M.,Gressens, P.,Baumgart, E.,Carmeliet, P.,Casteels, M.,Fransen, M.,Evrard, P.,Fahimi,

D.,Declercq, P. E.,Collen, D.,van Veldhoven, P. P. & Mannaerts, G. P. A mouse model for

Zellweger syndrome. Nat. Genet. 17, 49–57 (1997).

131. Hanson, M. G.,Fregoso, V. L.,Vrana, J. D.,Tucker, C. L. & Niswander, L. A. Peripheral nervous

system defects in a mouse model for peroxisomal biogenesis disorders. Dev. Biol. 395, 84–95

(2014).

132. Li, X.,Baumgart, E.,Dong, G.-X.,Morrell, J. C.,Jimenez-Sanchez, G.,Valle, D.,Smith, K.

D. & Gould, S. J. PEX11alpha is required for peroxisome proliferation in response to 4-phenylbutyrate but is dispensable for peroxisome proliferator-activated receptor

alpha-mediated peroxisome proliferation. Mol. Cell. Biol. 22, 8226–40 (2002).

133. Weng, H.,Ji, X.,Naito, Y.,Endo, K.,Ma, X.,Takahashi, R.,Shen, C.,Hirokawa, G.,Fukushima, Y.

& Iwai, N. Pex11α deficiency impairs peroxisome elongation and division and contributes to

nonalcoholic fatty liver in mice. Am. J. Physiol. Endocrinol. Metab. 304, E187–96 (2013).

134. Li, X.,Baumgart, E.,Morrell, J. C.,Valle, D.,Gould, S. J. & Jimenez-sanchez, G. PEX11

beta deficiency is lethal and impairs neuronal migration but does not abrogate peroxisome

function. Mol Cell Biol 22, 4358–4365 (2002).

135. Maxwell, M.,Bjorkman, J.,Nguyen, T.,Finnie, J.,Paterson, C.,Tonks, I.,Barbara, C.,Kay, G.

F.,Crane, D. I.,Sharp, P. & Paton, B. C. Pex13 inactivation in the mouse disrupts peroxisome

biogenesis and leads to a Zellweger syndrome phenotype. Mol Cell Biol 23, 5947–5957

(2003).

136. Hiebler, S.,Masuda, T.,Hacia, J. G.,Moser, A. B.,Faust, P. L.,Liu, A.,Chowdhury, N.,Huang,

N.,Lauer, A.,Bennett, J.,Watkins, P. a,Zack, D. J.,Braverman, N. E.,Raymond, G. V & Steinberg, S. J. The Pex1-G844D mouse: a model for mild human Zellweger spectrum

disorder. Mol. Genet. Metab. 111, 522–32 (2014).

137. Ebberink, M. S.,Mooijer, P. a W.,Gootjes, J.,Koster, J.,Wanders, R. J. A. & Waterham, H. R.

Genetic classification and mutational spectrum of more than 600 patients with a Zellweger

syndrome spectrum disorder. Hum. Mutat. 32, 59–69 (2011).

138. Rosewich, H.,Ohlenbusch, a & Gärtner, J. Genetic and clinical aspects of Zellweger spectrum

patients with PEX1 mutations. J. Med. Genet. 42, e58 (2005).

139. Imamura, A.,Tamura, S.,Shimozawa, N.,Suzuki, Y.,Zhang, Z.,Tsukamoto, T.,Orii, T.,Kondo,

N.,Osumi, T. & Fujiki, Y. Temperature-sensitive mutation in PEX1 moderates the phenotypes of

Referenties

GERELATEERDE DOCUMENTEN

Voorr de patrociniumwisseling van de Oudmunsterkerk zijn verschillende verklarin- genn te geven. Stellig heeft men door St-Bonifatius zo nadrukkelijk op de voorgrond te plaatsen,,

Twaalfde-eeuw- sese bronnen maken ook melding van kapittelgoed (wijngaarden) van St-Marie in het nabijgelegen Bendorf: MGHH DFI nr.. Net als bij de koningsprebende in het

Er moeten in Utrecht dus nog andere factoren in het spel zijnn geweest In dit verband dient nogmaals de aandacht te worden gevestigd op de gevolgenvann de diepgaande en

Ter vergelijking zijn ook de vermeldingen opgenomenn van Utrechtse heiligen in de litanie uit het psalterium van Wolbodo, vervaardigd omstreeks 1000, inn de tijd dat hij scholaster

Ordinarius sancri Martini (ca. Vellekoopp noemt St.-Bonifatius als wijdingsdag maar dit is onjuist:: Vellekoop 1999,7. 1300), ad 10 mei (vriendelijk mededeling Bram van den Hovenn

44 Ervan uitgaande dat de proost Liudolf, die na Ansfried wordt genoemd proost was vann een buitenkapittel, vermoedelijkk Oldenzaal (zie hierna n. 5), en dat Ansfried geen proost

6. Willibalds Leben des Bonifatius, nebst einige zeitgenössigen Dolcumen- ten.. Darmstadt Ausgewahlte Quellen zur deutschen Geschichte des Mittelalters, Freiherr vom

1100 233 Utrecht: de nederzetting met de kerken eind elfde eeuw 244 Bisdom Utrecht: de indeling in aartsdiakonaten 255 Middelburg: de nederzetting en de kerken