We have investigated the utility of bone marrow–derived mesenchymal stem cells (MSCs) as targets for gene therapy of the autosomal recessive disorder mucopolysaccharidosis type IH (MPS-IH, Hurler syndrome). Cultures of MSCs were initially exposed to a green fluorescent protein–expressing retrovirus. Green fluorescent protein–positive cells maintained their proliferative and differentiation capacity. Next we used a vector encoding α-l-iduronidase (IDUA), the enzyme that is defective in MPS-IH. Following transduction, MPS-IH MSCs expressed high levels of IDUA and secreted supernormal levels of this enzyme into the extracellular medium. Exogenous IDUA expression led to a normalization of glycosaminoglycan storage in MPS-IH cells, as evidenced by a dramatic decrease in the amount of 35SO4sequestered within the heparan sulfate and dermatan sulfate compartments of these cells. Finally, gene-modified MSCs were able to cross-correct the enzyme defect in untransduced MPS-IH fibroblasts via protein transfer.

Mucopolysaccharidosis type IH (MPS-IH, Hurler syndrome) is an autosomal recessive disorder resulting from defects in the gene encoding the lysosomal enzyme α-l-iduronidase (IDUA). This leads to ineffective degradation of the glycosaminoglycans (GAGs) heparan sulfate and dermatan sulfate. Individuals with very low levels of IDUA present in infancy and early childhood as a consequence of the deleterious accumulation of these GAGs in different organ systems, including the central nervous system, reticuloendothelial system, and the skeleton. Such severely affected patients usually die within the first decade.1,2 

Current therapy for MPS-IH focuses on allogeneic bone marrow transplantation from an unaffected, HLA-compatible donor. This provides normal, enzyme-competent leukocytes that secrete IDUA that can be taken up by enzyme-deficient cells via mannose-6-phosphate receptors.3 The utility of this approach is significantly limited by the availability of donors and significant toxicity of the intense immunosuppressive conditioning therapy that the recipient requires for donor hemopoiesis to become established without rejection. Even where donor hemopoiesis is fully established (ie, all hemopoietic cells have normal enzyme levels), symptoms (particularly defects in the skeleton and central nervous system) are incompletely and variably corrected.4,5 

Mesenchymal stem cells (MSCs) are multipotent progenitors that can be isolated from bone marrow and are capable of contributing to multiple mesenchymal tissues in vivo.6-10 In this paper we demonstrate, for the first time, retroviral gene transfer leading to correction of these MSCs in an inherited disorder. Furthermore, there is maintenance of the proliferative and multilineage differentiation potential of these modified cells, and they are able to cross-correct non–gene-modified cells.

Numerous studies have demonstrated the presence of donor mesenchymal cells in multiple tissues following transplantation, and MSCs injected into brain are able to differentiate into nerve cells. Taken with these, our data indicate that MSCs may prove a better target than hematopoietic stem cells in the context of gene therapy of multisystem, lysosomal storage disorders.

Isolation and culture of MSCs

Bone marrow samples were obtained from MPS-IH patients and unaffected individuals aged from 0 to 18 years, following informed parental consent and approval from the local research ethics committee. MSCs were isolated and cultured as previously described.11For differentiation assays, cells were plated at 5 × 103per well in 6-well plates in growth medium with either osteogenic12 or adipogenic13 supplements. For differentiation along the neuronal lineage, cells were preincubated for 24 hours in Dulbecco modified Eagle medium/20% fetal calf serum/1 mM β-mercapotethanol and then switched into Dulbecco modified Eagle medium/5 mM β-mercapotethanol.14 Mineralized bone was stained by the von Kossa technique,15 and adipocytes were stained using oil-red-O.16 Neurons were stained for the neuron-specific tyrosine kinase trkA using monoclonal sc-118 (Santa Cruz Biotechnology, Santa Cruz, CA).14 

Transduction of MSCs

The Lid vector expressing IDUA has been previously described.17 L– enhanced green fluorescent protein (L-EGFP) was derived by replacing the IDUA complementary DNA in Lid with a complementary DNA encoding L-EGFP. Cells (30%-40% confluent) were transduced using cell-free retroviral supernatant supplemented with 2 μg/mL polybrene. After 24 hours, medium was replaced and cells left for 5 to 7 days prior to use in assays for transgene expression, phenotypic correction, and differentiation.

Assays of IDUA activity

IDUA activity in cell homogenates and media was assayed as previously described18 using 4-methylumbelliferyl-α-l-iduronidase (Glycosynth, Cheshire, United Kingdom) as substrate. Total protein was measured according to the Lowry method.19 

Sulfate sequestration assay

Confluent MSCs were exposed to 35S-labeled Na2SO4 (NEN Life Science Products, Boston, MA) at 20 μCi/mL (0.74 MBq/mL) in Dulbecco modified Eagle medium/fetal calf serum for 1 week. Cells were then trypsinized and washed in phosphate-buffered saline to remove external GAGs. Following centrifugation at 800g for 10 minutes, cell pellets were solubilized in 2 mL of 6 M urea/0.15 M sodium phosphate, pH 7.0, containing 1% (vol/vol) Triton X-100 at 4°C for 1 hour. Extracts were filtered through a 0.2-μm syringe filter before application to a fast protein liquid chromatography Mono-Q HR 5/5 anion-exchange column (Pharmacia, St Albans, United Kingdom).

Nonincorporated 35SO4 was removed by washing through with 0.15 M NaCl/20 mM phosphate, pH 7.0, containing 1% (vol/vol) Triton X-100. Bound 35S-labeled material was eluted using a 60 mL linear gradient of 0.15 to 1.5 M NaCl in 20 mM phosphate, pH 7.0, containing 1% Triton X-100 at a flow rate of 1 mL/min and collecting 1 mL fractions. The 35S content of fractions was determined by liquid scintillation counting.

Following retroviral transduction of MSCs with the L-EGFP vector, transduced MSCs maintained the same growth rate as untransduced cells (not shown) and retained the ability to differentiate into osteoblasts (Figure 1A,B), adipocytes (Figure 1C,D), and neurons (Figure 1E,F). GFP-transduced, MSC-derived osteoblasts exhibited mineral deposits that could be visualized by von Kossa staining (Figure 1A). Transduced MSC-derived adipocytes stained with oil-red-O (Figure 1C), and neurons stained positively for trkA (Figure1E) and tau (not shown). Thus, the transduction conditions used did not compromise the proliferation and differentiation potential of the MSCs.

Fig. 1.

GFP-transduced MSCs maintain their multipotentiality.

GFP-transduced MSCs following differentiation down the osteoblast (A,B), adipocyte (C,D), or neuronal lineages (E,F). Lineage-specific staining for bone (von Kossa staining [A]), fat (oil-red-O [C]), or neuronal (trkA [E]) markers is shown along with UV visualization of GFP (B,D,F). Original magnification A-B, × 100; C-F, × 200.

Fig. 1.

GFP-transduced MSCs maintain their multipotentiality.

GFP-transduced MSCs following differentiation down the osteoblast (A,B), adipocyte (C,D), or neuronal lineages (E,F). Lineage-specific staining for bone (von Kossa staining [A]), fat (oil-red-O [C]), or neuronal (trkA [E]) markers is shown along with UV visualization of GFP (B,D,F). Original magnification A-B, × 100; C-F, × 200.

Close modal

Following transduction of MPS-IH MSCs with the IDUA retrovirus, levels of enzyme activity were measured that equaled or exceeded those detected in normal MSCs (Table 1). In contrast, no detectable IDUA was seen in untransduced MPS-IH MSCs. When cell-free medium was assayed (Table 2), no IDUA was detectable from untransduced MPS-IH MSC cultures. IDUA could be detected in medium from normal MSCs and in higher (around 7- to 200-fold) amounts from transduced MPS-IH MSCs. This higher level of secretion of recombinant IDUA is consistent with the inclusion of a rat pre-proinsulin leader at the 5′ end of the construct we have used, resulting in more efficient targeting of IDUA into the secretory pathway.20 

Table 1.

Enzyme and GAG levels in transduced and cross-corrected cells

Lysosomal enzyme levels in transduced MSCs (patients A–C, passages P1–P3)
MPS-IH AMPS-IH BMPS-IH CNormal (mean ± SEM)
Intracellular IDUA (μM/g.h)     
 Untransduced 134.8 ± 37.8 
 Transduced P1 254 568 918 (n = 5) 
 Transduced P2 280 346 409  
 Transduced P3 nd 192 605  
Extracellular IDUA (nM/mL.h)     
 Untransduced 0.90 ± 0.37 
 Transduced P1 141 148 (n = 4) 
 Transduced P2 133 167  
 Transduced P3 81 197  
Lysosomal enzyme levels in transduced MSCs (patients A–C, passages P1–P3)
MPS-IH AMPS-IH BMPS-IH CNormal (mean ± SEM)
Intracellular IDUA (μM/g.h)     
 Untransduced 134.8 ± 37.8 
 Transduced P1 254 568 918 (n = 5) 
 Transduced P2 280 346 409  
 Transduced P3 nd 192 605  
Extracellular IDUA (nM/mL.h)     
 Untransduced 0.90 ± 0.37 
 Transduced P1 141 148 (n = 4) 
 Transduced P2 133 167  
 Transduced P3 81 197  

nd indicates not done.

Table 2.

IDUA levels in MPS-1H fibroblasts exposed to different conditioned media (μM/g.h)

Conditioned medium Fresh mediumNormal MSCsMPS-IH MSCsTransduced MPS-IH MSCsTransduced MPS-IH MSCs + m-6-pTransduced MPS-IH MSCs + g-6-p
MPS-IH fibroblasts 1.5 208.9 4.9 360.7  
Normal fibroblasts 71.9 nd nd nd nd nd 
Conditioned medium Fresh mediumNormal MSCsMPS-IH MSCsTransduced MPS-IH MSCsTransduced MPS-IH MSCs + m-6-pTransduced MPS-IH MSCs + g-6-p
MPS-IH fibroblasts 1.5 208.9 4.9 360.7  
Normal fibroblasts 71.9 nd nd nd nd nd 

m-6-p indicates mannose-6-phosphate; g-6-p, glucose-6-phosphate.

Cell-free medium from MSCs was next applied to cultures of MPS-IH fibroblasts with a view to testing the cross-correction potential of the secreted enzyme (Table 2). As expected, medium from uncorrected MPS-IH MSCs did not correct the defect in MPS-IH fibroblasts. Medium from normal MSCs did correct to a small extent but, most strikingly, medium from gene-modified MPS-IH MSCs conferred high levels of IDUA levels on MPS-IH fibroblasts. This cross-correction was inhibited by mannose-6-phosphate but not by the structural analog glucose-6-phosphate, confirming that uptake was dependent on the mannose-6-phosphate receptor.3 Thus, gene-corrected MPS-IH MSCs secrete IDUA in an appropriate form that may be taken up by non–gene-corrected cells.

To test the effect of exogenous IDUA expression on the storage of GAGs in transduced MPS-IH MSCs, cultures were exposed to35SO4 to radiolabel proteoglycans (Table3). MSCs from 2 separate individuals with MPS-IH were tested. MPS-IH MSCs showed significant amounts of35SO4 sequestration due to its accumulation in the GAGs dermatan and heparan sulfate and a lack of subsequent catabolism of these. IDUA-expressing MPS-IH MSC cultures, however, showed levels of 35SO4 sequestration similar to those in MSCs from unaffected individuals, indicating a normalization of GAG levels in Lid-transduced MPS-IH MSC cultures. Thus, not only were IDUA levels corrected, but the levels of the pathological effector (namely, stored GAGs) were also corrected.

Table 3.

35SO4 incorporation (cpm) into heparan sulfate and dermatan sulfate

Normal MSCsMPS-IH AMPS-IH A (transduced)MPS-IH BMPS-IH B (transduced)
Heparan sulfate 4 973 54 671 14 957 34 515 6 120 
Dermatan sulfate 32 567 227 044 19 923 175 135 22 151 
Total GAGs 37 540 281 715 34 880 209 650 28 271 
Normal MSCsMPS-IH AMPS-IH A (transduced)MPS-IH BMPS-IH B (transduced)
Heparan sulfate 4 973 54 671 14 957 34 515 6 120 
Dermatan sulfate 32 567 227 044 19 923 175 135 22 151 
Total GAGs 37 540 281 715 34 880 209 650 28 271 

Thus, MSCs may provide a useful platform for the production of lysosomal enzymes and other bioactive molecules in patients. Clinical utility of this approach in the transplantation of gene-modified MSCs will depend upon achievement of sufficient donor chimerism in affected tissue. This has varied in studies to date, although the experience of allogeneic bone marrow transplantation in osteogenesis imperfecta has demonstrated that even very low levels of engraftment can result in clinical benefit.21 

The authors are grateful to Mr Steve Bagley for assistance with imaging.

Supported by the Royal Manchester Children's Hospital R&D fund, the Jeans for Genes appeal (Mucopolysaccharidosis Society, Amersham, United Kingdom), and the Cancer Research Campaign, London, United Kingdom.

The publication costs of this article were defrayed in part by page charge payment. Therefore, and solely to indicate this fact, this article is hereby marked “advertisement” in accordance with 18 U.S.C. section 1734.

1
Hopwood
JJ
Morris
CP
The mucopolysaccharidoses. Diagnosis, molecular genetics and treatment.
Mol Biol Med.
7
1990
381
404
2
Neufeld
EF
Meunzer
J
The mucopolysaccharidoses. In Scriver CH, Beandet AL, Sly WS, Valle D, eds.
The Metabolic Basis of Inherited Disease.
1995
565
1587
McGraw Hill
New York, NY
3
Stewart
K
Brown
OA
Morelli
AE
et al
Uptake of alpha-(L)-iduronidase produced by retrovirally transduced fibroblasts into neuronal and glial cells in vitro.
Gene Ther.
4
1997
63
75
4
Hopwood
JJ
Vellodi
A
Scott
HS
et al
Long-term clinical progress in bone marrow transplanted mucopolysaccharidosis type I patients with a defined genotype.
J Inherit Metab Dis.
16
1993
1024
1033
5
Peters
C
Balthazor
M
Shapiro
EG
et al
Outcome of unrelated donor bone marrow transplantation in 40 children with Hurler syndrome.
Blood.
87
1996
4894
4902
6
Banfi
A
Muraglia
A
Dozin
B
Mastrogiacomo
M
Cancedda
R
Quarto
R
Proliferation kinetics and differentiation potential of ex vivo expanded human bone marrow stromal cells: implications for their use in cell therapy.
Exp Hematol.
28
2000
707
715
7
Horwitz
EM
Prockop
DJ
Fitzpatrick
LA
et al
Transplantability and therapeutic effects of bone marrow-derived mesenchymal cells in children with osteogenesis imperfecta.
Nat Med.
5
1999
309
313
8
Kopen
GC
Prockop
DJ
Phinney
DG
Marrow stromal cells migrate throughout forebrain and cerebellum, and they differentiate into astrocytes after injection into neonatal mouse brains.
Proc Natl Acad Sci U S A.
96
1999
10711
10716
9
Liechty
KW
MacKenzie
TC
Shaaban
AF
et al
Human mesenchymal stem cells engraft and demonstrate site-specific differentiation after in utero transplantation in sheep.
Nat Med.
6
2000
1282
1286
10
Pereira
RF
O'Hara
MD
Laptev
AV
et al
Marrow stromal cells as a source of progenitor cells for nonhematopoietic tissues in transgenic mice with a phenotype of osteogenesis imperfecta.
Proc Natl Acad Sci U S A.
95
1998
1142
1147
11
Bruder
SP
Jaiswal
N
Haynesworth
SE
Growth kinetics, self-renewal, and the osteogenic potential of purified human mesenchymal stem cells during extensive subcultivation and following cryopreservation.
J Cell Biochem.
64
1997
278
294
12
Jaiswal
N
Haynesworth
SE
Caplan
AI
Bruder
SP
Osteogenic differentiation of purified, culture-expanded human mesenchymal stem cells in vitro.
J Cell Biochem.
64
1997
295
312
13
Nuttall
ME
Patton
AJ
Olivera
DL
Nadeau
DP
Gowen
M
Human trabecular bone cells are able to express both osteoblastic and adipocytic phenotype: implications for osteopenic disorders.
J Bone Miner Res.
13
1998
371
382
14
Woodbury
D
Schwarz
EJ
Prockop
DJ
Black
IB
Adult rat and human bone marrow stromal cells differentiate into neurons.
J Neurosci Res.
61
2000
364
370
15
Page
K
Stevens
A
Lowe
J
Bancroft
JD
Bone.
Theory and Practice of Histological Techniques.
Bancroft
JD
Stevens
A
1996
309
340
Churchill Livingston
New York, NY
16
Bayliss High
OB
Lake
B
Lipids.
Theory and Practice of Histological Techniques.
Bancroft
JD
Stevens
A
1996
213
242
Churchill Livingston
New York, NY
17
Fairbairn
LJ
Lashford
LS
Spooncer
E
et al
Long-term in vitro correction of alpha-L-iduronidase deficiency (Hurler syndrome) in human bone marrow.
Proc Natl Acad Sci U S A.
93
1996
2025
2030
18
Stirling
JL
Robinson
D
Fensom
AH
Benson
PF
Baker
JE
Fluorimetric assay for prenatal detection of Hurler and Scheie homozygotes or heterozygotes.
Lancet.
1
1978
147
19
Lowry
OH
Rosebrough
NJ
Farrand
AL
Randall
RJ
Protein measurement with the Folin-Phenol reagents.
J Biol Chem.
193
1951
265
275
20
Anson
DS
Bielicki
J
Hopwood
JJ
Correction of mucopolysaccharidosis type I fibroblasts by retroviral-mediated transfer of the human alpha-L-iduronidase gene.
Hum Gene Ther.
3
1992
371
379
21
Horwitz
EM
Prockop
DJ
Gordon
PL
et al
Clinical responses to bone marrow transplantation in children with severe osteogenesis imperfecta.
Blood.
97
2001
1227
1231

Author notes

Leslie J. Fairbairn, Paterson Institute for Cancer Research, Christie Hospital NHS Trust, Manchester, M20 4BX, United Kingdom; e-mail: lfairbairn@picr.man.ac.uk.

Sign in via your Institution