Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Jan 1.
Published in final edited form as: BioDrugs. 2010;24(1):9–21. doi: 10.2165/11530550-000000000-00000

Glycosylation of Therapeutic Proteins: An Effective Strategy to Optimize Efficacy

Ricardo J Solá 1, Kai Griebenow 1
PMCID: PMC2805475  NIHMSID: NIHMS150932  PMID: 20055529

Abstract

During their development and administration, protein-based drugs routinely display suboptimum therapeutic efficacies due to their poor physicochemical and pharmacological properties. These innate liabilities have driven the development of molecular level strategies to improve the therapeutic behavior of protein drugs. Among, the currently developed approaches, glycoengineering is one of the most promising due fact that it has been shown to simultaneously afford improvements over most of the parameters necessary for optimization of protein drug in vivo efficacy (e.g., in vitro and in vivo molecular stability, pharmacodynamic responses, and pharmacokinetic profiles) while allowing for targeting to the desired site of action. The intent of this article is to provide an account of the effects that glycosylation has on the therapeutic efficacy of protein drugs and to describe the current understanding of the mechanisms by which glycosylation leads to such effects.


In recent decades there has been an accelerated drive towards the increased development of protein based drugs due to their great economic and clinical importance. Although proteins display multiple therapeutically favorable properties (e.g., higher target specificities, pharmacological potencies, and frequently lower side effects) their development and employment is often hindered as these routinely display suboptimum therapeutic efficacies due to intrinsic limitations in their physicochemical and pharmacological properties. (117) As a result, there is great interest in the development and employment of molecular level approaches to improve the therapeutic efficacy of protein drugs by engineering their physicochemical and pharmacological properties.(1723) A promising approach being currently employed involves the strategic manipulation of the protein’s surface glycosylation patterns through glycoengineering.(13, 2429) Even though a vast amount of studies have demonstrated that glycosylation can lead to enhanced therapeutic efficacies for protein drugs, many aspects regarding the mechanisms by which glycosylation induces such effects remain unclear. The intent of this article is therefore to provide an account of the current understanding of the mechanisms by which glycosylation improves the therapeutic efficacy of protein drugs. This is achieved by presenting a survey of the principal physicochemical and pharmacological aspects limiting the therapeutic effectiveness of protein drugs, by addressing which of these can be improved by glycosylation, and by discussing the currently proposed mechanisms for such effects.

1. Intrinsic Limitations to Protein Therapeutic Efficacy

Protein drugs routinely display suboptimum therapeutic efficacies due to their inherently poor physicochemical and pharmacological properties. While poor physicochemical properties for protein drugs mainly arise from low in vitro and in vivo molecular stabilities their poor pharmacological properties are due to adverse pharmacodynamic (PD) responses and limited pharmacokinetic (PK) profiles.(9, 1416) All of these liabilities can diminish the clinical effectiveness of protein drugs by affecting their systemic bioavailability.

1.1 Molecular Instability

Proteins drugs generally display low in vitro molecular stabilities during their pharmaceutical development lifecycle due to their inherently liable structural elements coupled with several innate physical and chemical instabilities.(13) This problem is further compounded in a pharmaceutical setting as protein drugs are routinely exposed to several destabilizing environments during their production, purification, storage, and delivery. (4, 11, 30, 31) For example, the backbone and amino acid side chains of protein drugs can be subject to several chemical instability issues (e.g., chemical hydrolysis, fragmentation, crosslinking, oxidation, deamidation, β-elimination, and racemization) due to their potential to undergo acid-base and redox chemistries. (9, 3234) Additionally, the secondary and tertiary structural elements of proteins which are requisite for function can also be affected by physical instability issues; such as, irreversible conformational changes, local and global unfolding, due to their non-covalent nature.(4, 6, 30, 35) Protein drugs are also prone to pH, temperature, and concentration dependant precipitation, surface adsorption, and non-native supramolecular aggregation as a result of their colloidal properties.(3, 11, 3640) Furthermore, once administered their in vivo molecular stability becomes a limiting issue as their structure is susceptible to extra- and intra-cellular enzymatic degradation.(16) If left unaddressed these in vitro and in vivo molecular instability issues can adversely impact the therapeutic efficacy of protein drugs due to the direct dependency of pharmacological properties (PK/PD) on the amount of functionally active protein that is administered and persistent in circulation.(4, 13, 16)

1.2 Adverse Pharmacodynamic and Pharmacokinetic Profiles

Achieving optimum therapeutic efficacy is dependant on maintaining a proper balance between drug exposure and effect. Therefore, the PK and PD parameters for therapeutics are often tuned through the drug design lifecycle in a manner assuring that desired in vivo responses are achieved. PK refers to the time dependency of drug action (dose/metabolic profiles) and is influenced by drug absorption, distribution, and excretion as well as initial response times and duration of effects. PK parameters usually determined for protein drugs include circulatory half-lifes, volumes of distribution, clearance rates, and total bioavailability. PD examines the potency of drugs (dose/response profiles) through the study of in vitro activities. For protein based drugs PD parameters usually determined are enzymatic rates and receptor binding affinities. Protein based drugs usually display limited PK profiles and sharp PD responses as a result of their poor physicochemical and pharmacological properties.(15) Limited PK profiles are mainly evidenced by adverse local adsorption and systemic distribution patterns for subcutaneously (SC) administered protein drugs as result of variable protein hydropathy (hydrophilic/hydrophobic surface balance) and by fast excretion rates (e.g., short circulation half-lifes) for intravenously (IV) administered ones due to rapid elimination from the body through proteolytic, renal, hepatic, and receptor mediated clearance mechanisms.(15, 16)

All protein drugs are susceptible to some level of clearance through non-specific proteolytic degradation due to the ubiquitous nature and systemic distribution of proteases. Therefore, protein catabolism is not limited to the gastrointestinal, renal, and hepatic tissues but can also occur in the blood and at other tissues. (16) Such non-specific cleavage events negatively affect efficacy through inactivation of the protein drug.(13) For smaller sized protein drugs clearance occurs mainly at the kidneys with renal clearance occurring through three main routes.(16, 41) The first route, glomerular filtration, is controlled primarily by the kidney’s size permeaselectivity with the urine/blood filterability being greatly reduced for proteins with molecular weights and hydrodynamic radius greater that 50 kDa and 60 A.(4244) Proteins exceeding this molecular weight limit are eliminated by other pathways; mainly proteolytic degradation, hepatic uptake, and immune clearance. The second factor affecting kidney permeaselectivity is the protein surface charge. Filtration of highly charged proteins (both anionic and cationic) is retarded by the presence of negatively and positively charged proximal and distal elements within the renal glomerular basement membrane and epithelium.(16, 43) The second route of renal elimination; which applies mainly to small linear peptides, occurs through hydrolysis by brush border enzymes located on the luminal membrane followed by reabsorption. A third and less frequent route involves peritubular extraction from postglomerular capillaries followed by intracellular degradation. For larger sized proteins clearance occurs mainly at the liver through both specific and non-specific hepatic uptake mechanisms.(16) Specific clearance occurs through receptor-mediated (e.g. asialoglycoprotein and low density lipoprotein receptors) endocytosis at the hepatocytes. This process is interestingly regulated by the glycosylation state of proteins (discussed further on). Alternatively, non-specific hepatic clearance of proteins can also occur through phagocytosis in the reticuloendothelial system. Additional non-hepatic receptor mediated specific elimination mechanisms can occur whereas the protein drug is removed by endocytosis after binding to its therapeutic target receptor. This process is influenced by the protein drug potency with stronger receptor binders being removed from the systemic circulation faster.

As a result of the innate susceptibility of proteins to all of these clearance mechanisms, protein drugs generally display limited plasma persistence lifetimes.(15) Higher protein concentrations and increased dosing frequencies are therefore often employed to achieve favorable therapeutic responses. Ironically, such frequent treatment regimes coupled with the high target specificities and potencies of protein drugs can lead to inappropriately sharp dose/response profiles. This can lead to overstimulation of the targeted pathway and in many instances trigger autoregulatory feedback inhibition mechanisms that can be therapeutically counterproductive in the long run by leading to loss of in vivo efficacy.(45, 46) To counteract these innate limitations it has become routine practice to integrate molecular level technologies to engineer the physicochemical and pharmacological properties of protein drugs (second-generation biopharmaceuticals) in the early stages of their development lifecycle.(47) Established technologies that have been shown to significantly improve the efficacy of protein drugs by increasing their molecular stabilities and plasma persistence times and by decreasing their PD responses through various mechanisms include: targeted mutations, generation of fusion proteins and conjugates, glycosylation engineering, and pegylation.(1719, 2124, 48) Of these, engineered glycosylation is one of the most promising due fact that it has been shown to simultaneously afford improvements over most of the molecular parameters necessary for optimization of therapeutic efficacy while allowing for targeting to the desired site of action.(13, 2429)

2. Protein Glycosylation

A substantial fraction of the currently approved protein pharmaceuticals need to be properly glycosylated to exhibit optimum therapeutic efficacy (Table I). This is due to the fact that glycosylation can influence a variety of physiological processes at both the cellular (e.g. intracellular targeting) and protein levels (e.g. protein-protein binding, protein molecular stability).(24, 49, 50) Glycosylation refers to the covalent attachment of carbohydrate based molecules (glycans) to the protein surface. Glycosylation is the most prevalent and structurally complex of the chemical modifications that occur naturally in proteins.(5156) In this context glycosylation can display structural heterogeneity with respect to both the site of glycan attachment (macro-heterogeneity) and with respect to the glycan’s structure (micro-heterogeneity). Additionally, since all of the potential glycosylation sites are not occupied simultaneously this can lead to the formation of glycoforms with differences in the number of attached glycans.

Table I.

Partial List of Approved Glycoprotein-based BioDrugs

INN Brand Name (Company) Indication Number of Glycans Production System
Agalsidase alfa (galactosidase) Replagal® (Shire) Treatment of Fabry disease 6 N-Linked HF cells
Agalsidase beta (galactosidase) Fabrazyme® (Genzyme) Treatment of Fabry disease 6 N-Linked CHO cells
Alglucosidase alfa (glucosidase) Myozyme® (Shire) Treatment of Pompe disease 7 N-Linked CHO cells
Alpha 1-antitrypsin (α1AT) Prolastin® (Talecris Biotherapeutics) Treatment of congenital α1AT deficiency with emphysema 3 N-Linked Tissue fractionation (human placenta)
Antithrombin III Atryn® (Ovation Pharmaceutics)
Berinert® (CSL)
Prevention of peri-operative and peri-partum thromboelitic events 3–4 N-Linked Milk fractionation (transgenic goats)
C1-esterase-inhibitor Cinryze® (CSL) Treatment of hereditary angioedema 6 N-Linked
7 O-Linked
Plasma fractionation (human)
Choriogonadotropin alfa Ovidrel® (EMD Serono) Treatment of female infertility 4 N-Linked
4 O-Linked
CHO cells
Darbopoetin alfa ARANESP® (Amgen) Treatment of anemia associated with chronic renal failure 5 N-Linked
1 O-Linked
CHO cells
Dornase alfa (Dnase) Pulmozyme® (Genzyme) Treatment of cystic fibrosis 2 N-Linked CHO cells
Drotrecogin alfa (CF-XIV, Protein C) Xigris® (Eli Lilly) Treatment of severe sepsis 4 N-Linked HEK cells
Epoetin alfa, beta, delta, omega, zeta * Treatment of anemia associated with chronic renal failure (CRF) 3 N-Linked
1 O-Linked
CHO cells
Eptacog alfa (CF VIIa) NovoSeven® (Novo Nordisk) Treatment of spontaneous and surgical bleedings in haemophilia A and B 2 N-Linked
2 O-Linked
BHK cells
Fibrinogen Haemocomplettan® (CSL) Haemorrhagic diatheses in hypo-, dys-, or afibrinogenaemia 5 N-Linked Plasma fractionation (human)
Follitropin alfa Gonal-F® (EMD Serono) Treatment of female infertility 4 N-Linked CHO cells
Follitropin beta Follistim AQ® (Schering Plough) Treatment of female infertility 4 N-Linked CHO cells
Galsulfase Naglazyme® (Genzyme) Treatment of Maroteaux-Lamy syndrome 6 N-Linked CHO cells
Glucocerebrosidase Cerezyme® (Genzyme) Treatment of Type I Gaucher disease 4 N-Linked CHO cells
Hyaluronidase Hylenex® (Baxter/Halozyme Therapeutics) Ophthalmic surgery 3 N-Linked CHO cells
Iduronidase alfa (laronidase) Aldurazyme® (Genzyme) Treatment of Mucopolysaccharidosis I 6 N-Linked CHO cells
Idursulfase Elaprase® (Shire) Treatment of Mucopolysaccharidosis I 8 N-Linked Human cells
Imiglucerase Ceredase® (Genzyme) Treatment of Type I Gaucher disease 4 N-Linked Tissue fractionation (human placenta)
Insulin * Treatment of diabetes * *
Interferon-alfa-n3 Alferon N® (Hemisphere Rx)
Avonex® (Biogen)
Treatment of external condylomata acuminata 1 O-Linked HL cells
Interferon beta-1a Rebif® (Pfizer/EMD Serono) Treatment of multiple sclerosis 1 N-Linked CHO cells
Lenograstim (G-CSF) Granocyte® (Chugai Pharma) Treatment of chemotherapy induced neutropenia 1 O-Linked CHO cells
Lutropin alfa Luveris® (EMD Serono) Treatment of female infertility 3 N-Linked CHO cells
Monoclonal antibodies * Multiple indications * *
Moroctocog alfa (CF VIII) * Prevention and control of hemorrhagic episodes in patients with haemophilia B 7 N-Linked CHO cells
BHK cells
Nonocog alfa (CF IX) Benefix® (Wyeth) Treatment of hemorrhagic episodes in patients with haemophilia B 2 N-Linked
4 O-Linked
CHO cells
Sargramostim (G-CSF) Leukine® (Genzyme) Treatment after induction chemotherapy with acute myelogenus leukemia 2 N-Linked
4 O-Linked
Yeast cells
Tenecteplase TNKase® (Genentech) Treatment of acute myocardial infarction 3 N-Linked
1 O-Linked
CHO cells
Thyrotropin alfa Thyrogen® (Genzyme) Detection of thyroid cancer and hypothyroidism 3 N-Linked CHO cells
Tissue plasminogen activator Alteplase® (Genentech) Treatment of acute myocardial infarction 3 N-Linked CHO cells
Urokinase alfa Abbokinase® (ImaRx Therapeutics) Treatment of acute massive pulmonary emboli 1 N-Linked
1 O-Linked
HK cells

Information was obtained from the Prescribing Information (PI) for each product. Further information available at www.fda.gov and www.emea.europa.eu.

*

Multiple approved products. Further information available at www.biopharma.com.

INN = International nonproprietary name.

CHO = Chinese hamster ovary

BHK = Baby hamster kidney

HEK = Human embryonic kidney

HF = Human fibroblast

HL = Human leukocytes

In humans the most prevalent glycosylation sites occur at asparagine residues (N-linked glycosylation through the Asn-X-Thr/Ser consensus sequence) and at serine or threonine residues (O-linked glycosylation).(57, 58) Further structural complexity occurs due to variability in the glycan’s monosaccharide sequence order, branching pattern, and length. Nonetheless, certain glycan core structures have been identified with these being formed by the enzymatic bridging and remodeling of the following monosaccharides: fucose (Fuc), galactose (Gal), mannose (Man), N-acetylglucosamine (GlcNAc), N-acetylgalactosamine (GalNAc), and N-acetylneuraminic acid (sialic acid).(55, 59) In humans three principal N-linked core glycan structures are formed with these being classified according to their monosaccharide content and structure: high mannose type (Man2-6Man3GlcNAc2), mixed type (GlcNAc2Man3GlcNAc2), and hybrid type (Man3GlcNAcMan3GlcNAc2).(59) For O-linked glycans four principal core structures have been identified: core 1 (GalGalNAc), core 2 (GalGlcNAcGalNAc), core 3 (GlcNAcGalNAc), and core 4 (GlcNAc2GalNAc). The terminal ends of these glycan core structures are often further functionalized (e.g., phosphates, sulfates, carboxylic acids) with chemically charged glycans (e.g., sialic acid) in human glycoproteins, leading to even greater structural diversity. These functionalized terminal glycans can alter the protein’s surface charge and isoelectric point (pI) which have been related to the increased circulatory lifetimes for glycoproteins.(23) Such structural diversity poses certain problems for glycoprotein based drugs as it has been found that variations in the expression system (Table I) or changes to the manufacturing process can lead to changes in glycosylation patterns.(60) Accordingly, alternative methods of glycoprotein production are being explored to design protein drugs with homogeneous structurally defined glycosylation patterns through genetic, enzymatic, and chemo enzymatic methods.(48, 6175) The reader is referred to the following recent reviews for extensive discussions on the details of different glycoprotein production systems.(62, 7688)

3. Optimization of Protein Therapeutic Efficacy by Glycosylation

3.1 Effects of Glycosylation on Protein Molecular Instability

A vast amount of studies have demonstrated that natural glycosylation increases the molecular stability of proteins (for a detailed mechanistic account refer to the recent review on glycoprotein biophysics by Solá et al.).(49) Furthermore, engineered glycosylation has been shown to stabilize a variety of protein drugs against almost all of the major physicochemical instabilities encountered during their pharmaceutical employment thus leading to enhanced in vitro molecular stabilities.(13) Pharmaceutically relevant protein instabilities which are improved by glycosylation include: oxidation; cross-linking; pH, chemical, heating, and freezing induced unfolding/denaturation; precipitation; kinetic inactivation; and aggregation. (13) Furthermore, these stabilization effects appear to be of a generalized nature since they have been shown to occur in a variety of structurally unrelated proteins.(13) Protein drugs whose stability has been reported to be increased by glycosylation include agalsidase alfa (REPLAGAL®, Shire) (aggregation, precipitation),(89) alglucosidase (MYOZYME®; Shire) (thermal denaturation),(90) alpha 1-antitrypsin (PROLASTIN®; Talecris Biotherapeutics) (chemical/thermal denaturation),(91) chymotrypsin (WOBE MUGOS®; Marlyn Nutraceuticals) (chemical/thermal/kinetic denaturation, aggregation),(92, 93) choriogonadotropin alfa (OVIDREL®; EMD Serono) (thermal denaturation),(94) epoetin alfa (EPOGEN®/PROCRIT®; Amgen/Ortho Biotech) (thermal/pH/chemical/kinetic denaturation, oxidation, aggregation),(9597) interferon beta (AVONEX®/REBIF®; Biogen/Pfizer EMD Serono) (disulfide crosslinking, precipitation, thermal denaturation, aggregation),(98100) ranpirnase (ONCONASE®; Alfacell) (thermal denaturation),(101, 102) lenograstim (GRANOCYTE®/NEUTROGIN®; Sanofi Aventis/Chugai Pharma) (thermal/pH/kinetic denaturation, disulfide crosslinking),(103105) thyrotropin alfa (THYROGEN®; Genzyme) (aggregation),(106) urokinase alfa (ABBOKINASE®; ImaRx Therapeutics) (thermal denaturation),(107) insulin (nondisulfide crosslinking, aggregation),(108) and various IgG-like antibodies (thermal denaturation).(109, 110) These in vitro stabilization effects as a result of protein glycosylation have been directly related to the amount of glycans present in the protein drug (for a detailed account of each of the different stabilization mechanisms see the recent review by Solá and Griebenow).(13) Many of these in vitro stabilization effects have been proposed to lead to increased in vivo efficacies by possibly allowing for a greater amount of functional protein to be administered to the patient and by minimizing the formation of neutralizing antibodies due to the lack of conformationally altered and aggregated species in the end formulation.(13)

In addition to these in vitro stabilizing effects, glycosylation can also result in increased in vivo molecular stability for protein drugs once administered by leading to increased lifetimes for the functional forms of these proteins due to its prevention of proteolytic degradation.(13, 111114) Some examples of therapeutically relevant proteins whose proteolytic susceptibility has been reported to be decreased by glycosylation include glucagons-like peptide 1,(115) lenograstim,(103, 116) bucelipase alfa,(117) drotrecogin alfa (XIGRIS®; Eli Lilly),(118) ranpirnase,(101, 102) thyrotropin alfa,(106) urokinase alfa,(119) interferon-γ (ACTIMMUNE®; Intermune),(120) and various IgG-like antibodies.(121) Proteolytic stabilization of proteins has also been related to the number of glycans bound to the protein surface; their length and branching; and the charges of their terminal end glycans therefore this effect can be influenced by both from steric and electrostatic repulsions induced by the surface bound glycans.(13, 122) In this context, it was recently shown by Raju et al. that negatively charged glycans (e.g., those ending with sialic acids) are more efficient in preventing antibody proteolysis.(121)

3.2 Effects of Glycosylation on Protein Pharmacodynamics and Pharmacokinetics

The initial understanding of the role of glycans on protein in vivo circulatory behavior can be attributed in large part to the discovery of the hepatic asialoglycoprotein receptor by Aswell and Morell in the 1960’s.(123126) While studying the mechanisms controlling the circulatory turnover of ceruloplasmin, a protein involved in hepatolenticular degeneration (Wilson disease), Aswell et al. noticed dramatic differences in the circulatory lifetimes between the natively glycosylated protein (sialic acid terminated glycans) (t1/2: ~ 56 hrs) and a partially deglycosylated variant of the protein (galactose terminated glycans) (t1/2: < 30 min).(127) Subsequent studies by them and others extended these findings to other proteins (α1-acid glycoprotein, α2-macroglobulin, thyroglobulin, haptoglobin, fetuin, orosomucoid, ribonuclease) validating the generality of this specific hepatic clearance mechanism.(128130) It was found that exposure of galactose terminating glycans through desialylation led to fast removal of the partially deglycosylated proteins from the circulation due to specific endocytosis mediated by asialoglycoprotein receptors expressed in the hepatocytes.(123, 124, 129, 131) Subsequently it was found that glycoproteins exposing glycans terminating in mannose, N-acetyl-glucosamine, or fucose could be also removed from the circulation due to specific interactions with other mammalian lectin-like receptors expressed at different cell types.(132136) These pioneering studies highlighted several important facts about the effects of surface glycans on the circulatory behavior of glycoproteins: (i) improperly glycosylated proteins are rapidly removed from the circulation by specific receptor-based mechanisms, (ii) natively glycosylated proteins (sialic acid terminated) have longer circulating lifetimes than non-glycosylated proteins and partially glycosylated proteins, (iii) increased sialic acid and glycan content correlates with increased circulating lifetimes, (iv) depending on their glycosylation patterns proteins can be targeted to certain tissue types and organs. Accordingly most studies on the effects of glycosylation on the in vivo efficacy of proteins have emphasized on the role of increased glycan content and glycan structure.(23) It is important to note that these lectin-like receptor based clearance mechanisms do not apply to all glycosylated biopharmaceuticals. An exception being IgG-like antibodies whose clearance is mediated via the neonatal Fc receptor (FcRn) and is not influenced by antibody glycosylation or glycoforms.(137) Glycosylation of protein drugs has been found to lead to improved therapeutic efficacy by increasing in vivo bioavailability, ambient circulating levels, and duration of effects through the modulation of their PK/PD properties. Changes to protein PK parameters induced by glycosylation include: improved absorption for small peptides,(138142) modulated absorption for larger proteins,(143) improved distribution,(26, 144) longer circulation lifetimes,(26, 27, 130, 145153) and decreased clearance rates.(2427, 29, 42, 82, 130, 140, 145171) Alternatively glycosylation modulates protein PK parameters by leading to altered potencies as a result of diminished in vitro enzymatic activities and altered receptor binding affinities.(49, 137, 140, 172, 173) Therefore similarly as to what has been described to occur for other protein engineering methodologies (e.g. pegylation), glycosylation appears to modulate the in vivo efficacy of protein drugs by altering the balance between their potencies (PD) and exposure times (PK).(174) In the specific case of IgG-like antibody based therapeutics, glycosylation has been shown to improve their in-vivo therapeutic efficacy by altering their effector functions through modulated binding affinities for the FcγR receptor (for a mechanistic discussion see the following reviews).(21, 175, 176)

Examples of therapeutically relevant proteins whose in vivo efficacies have been reported to be increased by their natural glycosylation include: agalsidase alfa,(177179) agalsidase beta (FABRAZYME®; Genzyme),(177, 179181) epoetin alfa and epoetin beta,(143, 168, 172, 182, 183) follitropin alfa (GONAL-F®; Merck/Serono) and follitropin beta (FOLLISTIM®; Schering-Plough),(159, 160, 169, 184, 185) insulin growth factor binding protein 6 (IGFBP-6),(163) lutropin alfa (LUVERIS®; Merck/Serono),(186191) transforming growth factor β1,(192) antithrombin (ATryn®/TROMBATE-III®; Genzyme/Talecris Biotherapeutics),(162) thyrotropin alfa (THYROGEN®; Genzyme),(166) lenograstim,(103, 193) sargramostim (LEUKINE®; Genzyme),(154, 194, 195) interleukin-3,(196) prourokinase,(151) lymphotoxin,(152, 197) C1-esterase inhibitor (Berinert®; CSL),(198200) IgG-like antibodies,(72, 121, 201) interferon beta,(98, 202, 203) coagulation factor VIIa (NOVOSEVEN®; Novo Nordisk),(204) coagulation factor VIII (moroctocog alfa),(155, 156, 205) coagulation factor IX (nonacog alfa) (BENEFIX®; Wyeth), and the p55 tumor necrosis receptor fusion protein.(158) In most of these studies increased circulatory lifetimes and improved in vivo activities have been attributed to reduced hepatic and renal clearance as well as diminished proteolytic degradation as a result of the presence of the charged glycans (terminal sialic acid). Expanding on this concept several studies have shown that increasing the number of sialic acid containing glycans beyond those of the native protein through engineered hyperglycosylation can effectively be employed as a technology to further optimize the circulatory half-life and in vivo activity of proteins.(24, 27, 206208) Examples of pharmaceutically relevant proteins whose circulatory half-lifes were shown to be increased by hyperglycosylation include: interferon alfa and gamma,(26, 120) luteinizing hormone,(149) Fv antibody fragments,(209) asparaginase,(210, 211) cholinesterase,(164, 165) darbepoetin alfa (AraNESP®; Amgen),(25, 27, 161, 212214) trombopoietin,(25, 27, 215) leptin,(25, 27) FSH,(159, 184, 216, 217) IFN-α2,(26) serum albumin,(145) and corifollitropin alfa.(218222) Engineered glycosylation has been also employed to further optimize the in vivo pharmacological behavior of protein drugs by allowing for targeted delivery to disease-affected tissues.(48) This methodology has been mainly applied to treat the lysosomal storage diseases (e.g. Gaucher, Pompe, and Fabry disease; Hurler and Maroteaux-Lamy syndrome).(223) The employed strategy involves enzyme transport to the lysosomes by receptor-mediated endocytosis after targeting the mannose and IGF-II/cation-independent mannose 6-phosphate receptors.(224228) This is achieved through glycoengineering of the protein glycans to expose at their terminal either mannose or mannose 6-phosphate. This strategy has been employed successfully with the following enzymes: β-glucocerebrosidase (CEREZYME®; Genzyme),(229231) α-glucosidase (MYOZYME®; Genzyme),(232) α-galactosidase (FABRAZYME®/REPLAGAL® Genzyme/Shire),(177, 178) galsulfase (NAGLAZYME®; Biomarin Pharmaceuticals),(233) and α-L-iduronidase (ALDURAZYME®; Genzyme/Biomarin Pharmaceuticals).(234) For all of these enzymes increase therapeutic efficacy has been achieved by targeting the protein drug to the desired site of action.

4. Conclusions and Future Prospects

Design of protein therapeutics with optimized in vivo efficacy can be achieved through the simultaneous optimization of drug molecular stability, pharmacokinetics, pharmacodynamics, and targeting by engineered glycosylation. This technology can be employed to ameliorate a multitude of pharmaceutically relevant physicochemical and pharmacological problems. Mechanistically, it appears that the different glycosylation parameters (e.g., number of glycans attached, glycan’s molecular size, sequence, and charge) can modulate the pharmacological properties of protein drugs to different extents. Engineered glycosylation could therefore provide ample future opportunities towards the improvement of protein drugs since in principle all of these glycosylation parameters can be simultaneously optimized.

While the pharmaceutical application of glycosylation still suffers from some technical challenges due to the intrinsically complex nature of glycan structures and the difficulties related to glycoprotein production in host-expression systems (e.g., low glycoprotein expression yields, glycan structural macro- and micro-heterogeneity), further advancements in the understanding of chemical- and enzyme- based glycan remodeling strategies being currently pursued by glycoengineering companies, will allow for the rational design of targeted glycoprotein structures. The significant potential that glycosylation engineering holds towards improving the therapeutic efficacy of protein drugs should lead to further research towards the understanding of the fundamental effects that glycans have on protein physicochemical and pharmacological properties.

Acknowledgments

This publication was made possible by a grant to K.G. (SC1 GM086240) from the National Institute of General Medical Sciences (NIGMS) at the National Institutes of Health (NIH) through the Support of Competitive Research (SCORE) program. Its contents are solely the responsibility of the authors and do not necessarily represent the official views of NIGMS. The authors have no conflicts of interest that are directly relevant to the contents of this article. The article will be available as the final edited version at: http://adisonline.com/biodrugs/Pages/default.aspx.

References

  • 1.Andersen DC, Krummen L. Recombinant protein expression for therapeutic applications. Curr Opin Biotechnol. 2002;13(2):117–23. doi: 10.1016/s0958-1669(02)00300-2. [DOI] [PubMed] [Google Scholar]
  • 2.Carpenter JF, Manning MC, Randolph TW. Long-term storage of proteins. Curr Protoc Protein Sci. 2002;Chapter 4(Unit 46) doi: 10.1002/0471140864.ps0406s27. [DOI] [PubMed] [Google Scholar]
  • 3.Hawe A, Friess W. Formulation development for hydrophobic therapeutic proteins. Pharm Dev Technol. 2007;12(3):223–37. doi: 10.1080/10837450701247350. [DOI] [PubMed] [Google Scholar]
  • 4.Wang W. Instability, stabilization, and formulation of liquid protein pharmaceuticals. International Journal Of Pharmaceutics. 1999;185(2):129–188. doi: 10.1016/s0378-5173(99)00152-0. [DOI] [PubMed] [Google Scholar]
  • 5.Frokjaer S, Otzen DE. Protein drug stability: A formulation challenge. Nature Reviews Drug Discovery. 2005;4(4):298–306. doi: 10.1038/nrd1695. [DOI] [PubMed] [Google Scholar]
  • 6.Manning MC, Patel K, Borchardt RT. Stability of Protein Pharmaceuticals. Pharmaceutical Research. 1989;6(11):903. doi: 10.1023/a:1015929109894. [DOI] [PubMed] [Google Scholar]
  • 7.Davis GC. Protein stability: impact upon protein pharmaceuticals. Biologicals. 1993;21(2):105. doi: 10.1006/biol.1993.1057. [DOI] [PubMed] [Google Scholar]
  • 8.Krishnamurthy R, Manning MC. The stability factor: importance in formulation development. Curr Pharm Biotechnol. 2002;3(4):361–71. doi: 10.2174/1389201023378229. [DOI] [PubMed] [Google Scholar]
  • 9.Arakawa T, Prestrelski SJ, Kenney WC, Carpenter JF. Factors affecting short-term and long-term stabilities of proteins. Advanced Drug Delivery Reviews. 2001;46(1–3):307–326. doi: 10.1016/s0169-409x(00)00144-7. [DOI] [PubMed] [Google Scholar]
  • 10.Lee JC. Biopharmaceutical formulation. Current Opinion In Biotechnology. 2000;11(1):81–84. doi: 10.1016/s0958-1669(99)00058-0. [DOI] [PubMed] [Google Scholar]
  • 11.Wang W. Protein aggregation and its inhibition in biopharmaceutics. International Journal Of Pharmaceutics. 2005;289(1–2):1–30. doi: 10.1016/j.ijpharm.2004.11.014. [DOI] [PubMed] [Google Scholar]
  • 12.Patro SY, Freund E, Chang BS. Protein formulation and fill-finish operations. Biotechnol Annu Rev. 2002;8:55–84. doi: 10.1016/s1387-2656(02)08004-3. [DOI] [PubMed] [Google Scholar]
  • 13.Sola RJ, Griebenow K. Effects of glycosylation on the stability of protein pharmaceuticals. J Pharm Sci. 2009;98(4):1223–45. doi: 10.1002/jps.21504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Brown LR. Commercial challenges of protein drug delivery. Expert Opin Drug Deliv. 2005;2(1):29–42. doi: 10.1517/17425247.2.1.29. [DOI] [PubMed] [Google Scholar]
  • 15.Mahmood I, Green MD. Pharmacokinetic and pharmacodynamic considerations in the development of therapeutic proteins. Clin Pharmacokinet. 2005;44(4):331–47. doi: 10.2165/00003088-200544040-00001. [DOI] [PubMed] [Google Scholar]
  • 16.Tang L, Persky AM, Hochhaus G, Meibohm B. Pharmacokinetic aspects of biotechnology products. J Pharm Sci. 2004;93(9):2184–204. doi: 10.1002/jps.20125. [DOI] [PubMed] [Google Scholar]
  • 17.Beals JM, Shanafelt AB. Enhancing exposure of protein therapeutics. Drug Discovery Today: Technologies. 2006;3(1):87. doi: 10.1016/j.ddtec.2006.03.001. [DOI] [PubMed] [Google Scholar]
  • 18.Marshall SA, Lazar GA, Chirino AJ, Desjarlais JR. Rational design and engineering of therapeutic proteins. Drug Discovery Today. 2003;8(5):212–221. doi: 10.1016/s1359-6446(03)02610-2. [DOI] [PubMed] [Google Scholar]
  • 19.Lazar GA, Marshall SA, Plecs JJ, Mayo SL, Desjarlais JR. Designing proteins for therapeutic applications. Curr Opin Struct Biol. 2003;13(4):513–8. doi: 10.1016/s0959-440x(03)00104-0. [DOI] [PubMed] [Google Scholar]
  • 20.Jefferis R. Glycosylation of recombinant antibody therapeutics. Biotechnology Progress. 2005;21(1):11–16. doi: 10.1021/bp040016j. [DOI] [PubMed] [Google Scholar]
  • 21.Jefferis R. Glycosylation as a strategy to improve antibody-based therapeutics. Nat Rev Drug Discov. 2009;8(3):226–34. doi: 10.1038/nrd2804. [DOI] [PubMed] [Google Scholar]
  • 22.Jefferis R. Glycosylation of antibody therapeutics: optimization for purpose. Methods Mol Biol. 2009;483:223–38. doi: 10.1007/978-1-59745-407-0_13. [DOI] [PubMed] [Google Scholar]
  • 23.Byrne B, Donohoe GG, O’Kennedy R. Sialic acids: carbohydrate moieties that influence the biological and physical properties of biopharmaceutical proteins and living cells. Drug Discovery Today. 2007;12(7–8):319. doi: 10.1016/j.drudis.2007.02.010. [DOI] [PubMed] [Google Scholar]
  • 24.Sinclair AM, Elliott S. Glycoengineering: The effect of glycosylation on the properties of therapeutic proteins. Journal Of Pharmaceutical Sciences. 2005;94(8):1626–1635. doi: 10.1002/jps.20319. [DOI] [PubMed] [Google Scholar]
  • 25.Elliott S, Lorenzini T, Asher S, Aoki K, Brankow D, Buck L, et al. Enhancement of therapeutic protein in vivo activities through glycoengineering. Nat Biotechnol. 2003;21(4):414–21. doi: 10.1038/nbt799. [DOI] [PubMed] [Google Scholar]
  • 26.Ceaglio N, Etcheverrigaray M, Kratje R, Oggero M. Novel long-lasting interferon alpha derivatives designed by glycoengineering. Biochimie. 2008;90(3):437–49. doi: 10.1016/j.biochi.2007.10.013. [DOI] [PubMed] [Google Scholar]
  • 27.Koury MJ. Sugar coating extends half-lives and improves effectiveness of cytokine hormones. Trends Biotechnol. 2003;21(11):462–4. doi: 10.1016/j.tibtech.2003.09.002. [DOI] [PubMed] [Google Scholar]
  • 28.Raju TS, Briggs JB, Chamow SM, Winkler ME, Jones AJ. Glycoengineering of therapeutic glycoproteins: in vitro galactosylation and sialylation of glycoproteins with terminal N-acetylglucosamine and galactose residues. Biochemistry. 2001;40(30):8868–76. doi: 10.1021/bi010475i. [DOI] [PubMed] [Google Scholar]
  • 29.Beck A, Wagner-Rousset E, Bussat MC, Lokteff M, Klinguer-Hamour C, Haeuw JF, et al. Trends in glycosylation, glycoanalysis and glycoengineering of therapeutic antibodies and Fc-fusion proteins. Curr Pharm Biotechnol. 2008;9(6):482–501. doi: 10.2174/138920108786786411. [DOI] [PubMed] [Google Scholar]
  • 30.Wang W. Lyophilization and development of solid protein pharmaceuticals. International Journal Of Pharmaceutics. 2000;203(1–2):1–60. doi: 10.1016/s0378-5173(00)00423-3. [DOI] [PubMed] [Google Scholar]
  • 31.Wang W, Singh S, Zeng DL, King K, Nema S. Antibody structure, instability, and formulation. J Pharm Sci. 2007;96(1):1–26. doi: 10.1002/jps.20727. [DOI] [PubMed] [Google Scholar]
  • 32.Volkin DB, Mach H, Middaugh CR. Degradative covalent reactions important to protein stability. Molecular Biotechnology. 1997;8(2):105–122. doi: 10.1007/BF02752255. [DOI] [PubMed] [Google Scholar]
  • 33.Pace CN. Conformational stability of globular proteins. Trends Biochem Sci. 1990;15(1):14–7. doi: 10.1016/0968-0004(90)90124-t. [DOI] [PubMed] [Google Scholar]
  • 34.Xie M, Schowen RL. Secondary structure and protein deamidation. J Pharm Sci. 1999;88(1):8–13. doi: 10.1021/js9802493. [DOI] [PubMed] [Google Scholar]
  • 35.Pace CN. Measuring and increasing protein stability. Trends Biotechnol. 1990;8(4):93–8. doi: 10.1016/0167-7799(90)90146-o. [DOI] [PubMed] [Google Scholar]
  • 36.Valente JJ, Payne RW, Manning MC, Wilson WW, Henry CS. Colloidal behavior of proteins: effects of the second virial coefficient on solubility, crystallization and aggregation of proteins in aqueous solution. Curr Pharm Biotechnol. 2005;6(6):427–36. doi: 10.2174/138920105775159313. [DOI] [PubMed] [Google Scholar]
  • 37.Chi EY, Krishnan S, Kendrick BS, Chang BS, Carpenter JF, Randolph TW. Roles of conformational stability and colloidal stability in the aggregation of recombinant human granulocyte colony-stimulating factor. Protein Sci. 2003;12(5):903–13. doi: 10.1110/ps.0235703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Kueltzo LA, Wang W, Randolph TW, Carpenter JF. Effects of solution conditions, processing parameters, and container materials on aggregation of a monoclonal antibody during freeze-thawing. J Pharm Sci. 2008;97(5):1801–12. doi: 10.1002/jps.21110. [DOI] [PubMed] [Google Scholar]
  • 39.Cromwell ME, Hilario E, Jacobson F. Protein aggregation and bioprocessing. Aaps J. 2006;8(3):E572–9. doi: 10.1208/aapsj080366. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Roberts CJ. Non-native protein aggregation kinetics. Biotechnol Bioeng. 2007;98(5):927–38. doi: 10.1002/bit.21627. [DOI] [PubMed] [Google Scholar]
  • 41.Strober W, Waldmann TA. The role of the kidney in the metabolism of plasma proteins. Nephron. 1974;13(1):35–66. doi: 10.1159/000180368. [DOI] [PubMed] [Google Scholar]
  • 42.Caliceti P, Veronese FM. Pharmacokinetic and biodistribution properties of poly(ethylene glycol)-protein conjugates. Adv Drug Deliv Rev. 2003;55(10):1261–77. doi: 10.1016/s0169-409x(03)00108-x. [DOI] [PubMed] [Google Scholar]
  • 43.Weinstein T, Gafter U, Chagnac A, Skutelsky E. Distribution of glycosaminoglycans in rat renal tubular epithelium. J Am Soc Nephrol. 1997;8(4):586–95. doi: 10.1681/ASN.V84586. [DOI] [PubMed] [Google Scholar]
  • 44.Choi HS, Liu W, Misra P, Tanaka E, Zimmer JP, Itty Ipe B, et al. Renal clearance of quantum dots. Nat Biotechnol. 2007;25(10):1165–70. doi: 10.1038/nbt1340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Montastruc JL, Galitzky J, Berlan M, Montastruc P. [Mechanism of receptor regulation during repeated administration of drugs] Therapie. 1993;48(5):421–6. [PubMed] [Google Scholar]
  • 46.Laduron PM. From receptor internalization to nuclear translocation. New targets for long-term pharmacology. Biochem Pharmacol. 1994;47(1):3–13. doi: 10.1016/0006-2952(94)90431-6. [DOI] [PubMed] [Google Scholar]
  • 47.Walsh G. Second-generation biopharmaceuticals. Eur J Pharm Biopharm. 2004;58(2):185–96. doi: 10.1016/j.ejpb.2004.03.012. [DOI] [PubMed] [Google Scholar]
  • 48.Grabenhorst E, Schlenke P, Pohl S, Nimtz M, Conradt HS. Genetic engineering of recombinant glycoproteins and the glycosylation pathway in mammalian host cells. Glycoconj J. 1999;16(2):81–97. doi: 10.1023/a:1026466408042. [DOI] [PubMed] [Google Scholar]
  • 49.Sola RJ, Rodriguez-Martinez JA, Griebenow K. Modulation of protein biophysical properties by chemical glycosylation: biochemical insights and biomedical implications. Cell Mol Life Sci. 2007;64(16):2133–52. doi: 10.1007/s00018-007-6551-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Liu DT. Glycoprotein pharmaceuticals: scientific and regulatory considerations, and the US Orphan Drug Act. Trends Biotechnol. 1992;10(4):114–20. doi: 10.1016/0167-7799(92)90192-x. [DOI] [PubMed] [Google Scholar]
  • 51.Weerapana E, Imperiali B. Asparagine-linked protein glycosylation: from eukaryotic to prokaryotic systems. Glycobiology. 2006;16(6):91R–101R. doi: 10.1093/glycob/cwj099. [DOI] [PubMed] [Google Scholar]
  • 52.Mann M, Jensen ON. Proteomic analysis of post-translational modifications. Nature Biotechnology. 2003;21(3):255–261. doi: 10.1038/nbt0303-255. [DOI] [PubMed] [Google Scholar]
  • 53.Walsh CT, Garneau-Tsodikova S, Gatto GJ. Protein posttranslational modifications: The chemistry of proteome diversifications. Angewandte Chemie-International Edition. 2005;44(45):7342–7372. doi: 10.1002/anie.200501023. [DOI] [PubMed] [Google Scholar]
  • 54.Apweiler R, Hermjakob H, Sharon N. On the frequency of protein glycosylation, as deduced from analysis of the SWISS-PROT database. Biochim Biophys Acta. 1999;1473(1):4–8. doi: 10.1016/s0304-4165(99)00165-8. [DOI] [PubMed] [Google Scholar]
  • 55.Sears P, Wong CH. Enzyme action in glycoprotein synthesis. Cell Mol Life Sci. 1998;54(3):223–52. doi: 10.1007/s000180050146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Lehle L, Strahl S, Tanner W. Protein glycosylation, conserved from yeast to man: a model organism helps elucidate congenital human diseases. Angew Chem Int Ed Engl. 2006;45(41):6802–18. doi: 10.1002/anie.200601645. [DOI] [PubMed] [Google Scholar]
  • 57.Medzihradszky KF. Characterization of protein N-glycosylation. Methods Enzymol. 2005;405:116–38. doi: 10.1016/S0076-6879(05)05006-8. [DOI] [PubMed] [Google Scholar]
  • 58.Peter-Katalinic J. Methods in enzymology: O-glycosylation of proteins. Methods Enzymol. 2005;405:139–71. doi: 10.1016/S0076-6879(05)05007-X. [DOI] [PubMed] [Google Scholar]
  • 59.Hossler P, Mulukutla BC, Hu WS. Systems analysis of N-glycan processing in mammalian cells. PLoS One. 2007;2(1):e713. doi: 10.1371/journal.pone.0000713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Sethuraman N, Stadheim TA. Challenges in therapeutic glycoprotein production. Curr Opin Biotechnol. 2006;17(4):341–6. doi: 10.1016/j.copbio.2006.06.010. [DOI] [PubMed] [Google Scholar]
  • 61.Hsieh-Wilson LC. Tailor-made glycoproteins. Trends Biotechnol. 2004;22(10):489–91. doi: 10.1016/j.tibtech.2004.08.009. [DOI] [PubMed] [Google Scholar]
  • 62.Rich JR, Withers SG. Emerging methods for the production of homogeneous human glycoproteins. Nat Chem Biol. 2009;5(4):206–15. doi: 10.1038/nchembio.148. [DOI] [PubMed] [Google Scholar]
  • 63.Tarp MA, Clausen H. Mucin-type O-glycosylation and its potential use in drug and vaccine development. Biochim Biophys Acta. 2008;1780(3):546–63. doi: 10.1016/j.bbagen.2007.09.010. [DOI] [PubMed] [Google Scholar]
  • 64.Brik A, Ficht S, Wong CH. Strategies for the preparation of homogenous glycoproteins. Curr Opin Chem Biol. 2006;10(6):638–44. doi: 10.1016/j.cbpa.2006.10.003. [DOI] [PubMed] [Google Scholar]
  • 65.Grogan MJ, Pratt MR, Marcaurelle LA, Bertozzi CR. Homogeneous glycopeptides and glycoproteins for biological investigation. Annu Rev Biochem. 2002;71:593–634. doi: 10.1146/annurev.biochem.71.110601.135334. [DOI] [PubMed] [Google Scholar]
  • 66.Gamblin DP, Scanlan EM, Davis BG. Glycoprotein synthesis: an update. Chem Rev. 2009;109(1):131–63. doi: 10.1021/cr078291i. [DOI] [PubMed] [Google Scholar]
  • 67.van Kasteren SI, Kramer HB, Gamblin DP, Davis BG. Site-selective glycosylation of proteins: creating synthetic glycoproteins. Nat Protoc. 2007;2(12):3185–94. doi: 10.1038/nprot.2007.430. [DOI] [PubMed] [Google Scholar]
  • 68.Geng J, Mantovani G, Tao L, Nicolas J, Chen G, Wallis R, et al. Site-directed conjugation of “clicked” glycopolymers to form glycoprotein mimics: binding to mammalian lectin and induction of immunological function. J Am Chem Soc. 2007;129(49):15156–63. doi: 10.1021/ja072999x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Vazquez-Dorbatt V, Maynard HD. Biotinylated glycopolymers synthesized by atom transfer radical polymerization. Biomacromolecules. 2006;7(8):2297–302. doi: 10.1021/bm060105f. [DOI] [PubMed] [Google Scholar]
  • 70.Vazquez-Dorbatt V, Tolstyka ZP, Chang CW, Maynard HD. Synthesis of a Pyridyl Disulfide End-Functionalized Glycopolymer for Conjugation to Biomolecules and Patterning on Gold Surfaces. Biomacromolecules. 2009 doi: 10.1021/bm900395h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Choi BK, Bobrowicz P, Davidson RC, Hamilton SR, Kung DH, Li H, et al. Use of combinatorial genetic libraries to humanize N-linked glycosylation in the yeast Pichia pastoris. Proc Natl Acad Sci U S A. 2003;100(9):5022–7. doi: 10.1073/pnas.0931263100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Li H, Sethuraman N, Stadheim TA, Zha D, Prinz B, Ballew N, et al. Optimization of humanized IgGs in glycoengineered Pichia pastoris. Nat Biotechnol. 2006;24(2):210–5. doi: 10.1038/nbt1178. [DOI] [PubMed] [Google Scholar]
  • 73.Hamilton SR, Bobrowicz P, Bobrowicz B, Davidson RC, Li H, Mitchell T, et al. Production of complex human glycoproteins in yeast. Science. 2003;301(5637):1244–6. doi: 10.1126/science.1088166. [DOI] [PubMed] [Google Scholar]
  • 74.Hamilton SR, Davidson RC, Sethuraman N, Nett JH, Jiang Y, Rios S, et al. Humanization of yeast to produce complex terminally sialylated glycoproteins. Science. 2006;313(5792):1441–3. doi: 10.1126/science.1130256. [DOI] [PubMed] [Google Scholar]
  • 75.Potgieter TI, Cukan M, Drummond JE, Houston-Cummings NR, Jiang Y, Li F, et al. Production of monoclonal antibodies by glycoengineered Pichia pastoris. J Biotechnol. 2009;139(4):318–25. doi: 10.1016/j.jbiotec.2008.12.015. [DOI] [PubMed] [Google Scholar]
  • 76.Chiba Y, Jigami Y. Production of humanized glycoproteins in bacteria and yeasts. Curr Opin Chem Biol. 2007;11(6):670–6. doi: 10.1016/j.cbpa.2007.08.037. [DOI] [PubMed] [Google Scholar]
  • 77.Karg SR, Kallio PT. The production of biopharmaceuticals in plant systems. Biotechnol Adv. 2009 doi: 10.1016/j.biotechadv.2009.07.002. [DOI] [PubMed] [Google Scholar]
  • 78.Hossler P, Khattak SF, Li ZJ. Optimal and consistent protein glycosylation in mammalian cell culture. Glycobiology. 2009;19(9):936–49. doi: 10.1093/glycob/cwp079. [DOI] [PubMed] [Google Scholar]
  • 79.Chiba Y, Akeboshi H. Glycan engineering and production of ‘humanized’ glycoprotein in yeast cells. Biol Pharm Bull. 2009;32(5):786–95. doi: 10.1248/bpb.32.786. [DOI] [PubMed] [Google Scholar]
  • 80.Ballew N, Gerngross T. Production of therapeutic proteins in fungal hosts. Expert Opin Biol Ther. 2004;4(5):623–6. doi: 10.1517/14712598.4.5.623. [DOI] [PubMed] [Google Scholar]
  • 81.Gerngross TU. Advances in the production of human therapeutic proteins in yeasts and filamentous fungi. Nat Biotechnol. 2004;22(11):1409–14. doi: 10.1038/nbt1028. [DOI] [PubMed] [Google Scholar]
  • 82.Bork K, Horstkorte R, Weidemann W. Increasing the sialylation of therapeutic glycoproteins: The potential of the sialic acid biosynthetic pathway. J Pharm Sci. 2009 doi: 10.1002/jps.21684. [DOI] [PubMed] [Google Scholar]
  • 83.Werner RG, Kopp K, Schlueter M. Glycosylation of therapeutic proteins in different production systems. Acta Paediatr Suppl. 2007;96(455):17–22. doi: 10.1111/j.1651-2227.2007.00199.x. [DOI] [PubMed] [Google Scholar]
  • 84.Langdon RH, Cuccui J, Wren BW. N-linked glycosylation in bacteria: an unexpected application. Future Microbiol. 2009;4(4):401–12. doi: 10.2217/fmb.09.10. [DOI] [PubMed] [Google Scholar]
  • 85.Ko K, Ahn MH, Song M, Choo YK, Kim HS, Ko K, et al. Glyco-engineering of biotherapeutic proteins in plants. Mol Cells. 2008;25(4):494–503. [PubMed] [Google Scholar]
  • 86.Mohan C, Kim YG, Koo J, Lee GM. Assessment of cell engineering strategies for improved therapeutic protein production in CHO cells. Biotechnol J. 2008;3(5):624–30. doi: 10.1002/biot.200700249. [DOI] [PubMed] [Google Scholar]
  • 87.Shi X, Jarvis DL. Protein N-glycosylation in the baculovirus-insect cell system. Curr Drug Targets. 2007;8(10):1116–25. doi: 10.2174/138945007782151360. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Hamilton SR, Gerngross TU. Glycosylation engineering in yeast: the advent of fully humanized yeast. Curr Opin Biotechnol. 2007;18(5):387–92. doi: 10.1016/j.copbio.2007.09.001. [DOI] [PubMed] [Google Scholar]
  • 89.Ioannou YA, Zeidner KM, Grace ME, Desnick RJ. Human alpha-galactosidase A: glycosylation site 3 is essential for enzyme solubility. Biochem J. 1998;332 (Pt 3):789–97. doi: 10.1042/bj3320789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Clark SE, Muslin EH, Henson CA. Effect of adding and removing N-glycosylation recognition sites on the thermostability of barley alpha-glucosidase. Protein Eng Des Sel. 2004;17(3):245–9. doi: 10.1093/protein/gzh028. [DOI] [PubMed] [Google Scholar]
  • 91.Kwon KS, Yu MH. Effect of glycosylation on the stability of alpha1-antitrypsin toward urea denaturation and thermal deactivation. Biochim Biophys Acta. 1997;1335(3):265–72. doi: 10.1016/s0304-4165(96)00143-2. [DOI] [PubMed] [Google Scholar]
  • 92.Sundaram PV, Venkatesh R. Retardation of thermal and urea induced inactivation of alpha-chymotrypsin by modification with carbohydrate polymers. Protein Eng. 1998;11(8):699–705. doi: 10.1093/protein/11.8.699. [DOI] [PubMed] [Google Scholar]
  • 93.Sola RJ, Al-Azzam W, Griebenow K. Engineering of protein thermodynamic, kinetic, and colloidal stability: Chemical Glycosylation with monofunctionally activated glycans. Biotechnol Bioeng. 2006;94(6):1072–9. doi: 10.1002/bit.20933. [DOI] [PubMed] [Google Scholar]
  • 94.van Zuylen CW, de Beer T, Leeflang BR, Boelens R, Kaptein R, Kamerling JP, et al. Mobilities of the inner three core residues and the Man(alpha 1--6) branch of the glycan at Asn78 of the alpha-subunit of human chorionic gonadotropin are restricted by the protein. Biochemistry. 1998;37(7):1933–40. doi: 10.1021/bi9718548. [DOI] [PubMed] [Google Scholar]
  • 95.Uchida E, Morimoto K, Kawasaki N, Izaki Y, Abdu Said A, Hayakawa T. Effect of active oxygen radicals on protein and carbohydrate moieties of recombinant human erythropoietin. Free Radic Res. 1997;27(3):311–23. doi: 10.3109/10715769709065769. [DOI] [PubMed] [Google Scholar]
  • 96.Narhi LO, Arakawa T, Aoki KH, Elmore R, Rohde MF, Boone T, et al. The effect of carbohydrate on the structure and stability of erythropoietin. J Biol Chem. 1991;266(34):23022–6. [PubMed] [Google Scholar]
  • 97.Tsuda E, Kawanishi G, Ueda M, Masuda S, Sasaki R. The role of carbohydrate in recombinant human erythropoietin. Eur J Biochem. 1990;188(2):405–11. doi: 10.1111/j.1432-1033.1990.tb15417.x. [DOI] [PubMed] [Google Scholar]
  • 98.Runkel L, Meier W, Pepinsky RB, Karpusas M, Whitty A, Kimball K, et al. Structural and functional differences between glycosylated and non-glycosylated forms of human interferon-beta (IFN-beta) Pharm Res. 1998;15(4):641–9. doi: 10.1023/a:1011974512425. [DOI] [PubMed] [Google Scholar]
  • 99.Karpusas M, Whitty A, Runkel L, Hochman P. The structure of human interferon-beta: implications for activity. Cell Mol Life Sci. 1998;54(11):1203–16. doi: 10.1007/s000180050248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Conradt HS, Egge H, Peter-Katalinic J, Reiser W, Siklosi T, Schaper K. Structure of the carbohydrate moiety of human interferon-beta secreted by a recombinant Chinese hamster ovary cell line. J Biol Chem. 1987;262(30):14600–5. [PubMed] [Google Scholar]
  • 101.Rudd PM, Joao HC, Coghill E, Fiten P, Saunders MR, Opdenakker G, et al. Glycoforms modify the dynamic stability and functional activity of an enzyme. Biochemistry. 1994;33(1):17–22. doi: 10.1021/bi00167a003. [DOI] [PubMed] [Google Scholar]
  • 102.Kim BM, Kim H, Raines RT, Lee Y. Glycosylation of onconase increases its conformational stability and toxicity for cancer cells. Biochem Biophys Res Commun. 2004;315(4):976–83. doi: 10.1016/j.bbrc.2004.01.153. [DOI] [PubMed] [Google Scholar]
  • 103.Nissen C. Glycosylation of recombinant human granulocyte colony stimulating factor: implications for stability and potency. Eur J Cancer. 1994;30A (Suppl 3):S12–4. [PubMed] [Google Scholar]
  • 104.Oh-eda M, Hasegawa M, Hattori K, Kuboniwa H, Kojima T, Orita T, et al. O-linked sugar chain of human granulocyte colony-stimulating factor protects it against polymerization and denaturation allowing it to retain its biological activity. J Biol Chem. 1990;265(20):11432–5. [PubMed] [Google Scholar]
  • 105.Ono M. Physicochemical and biochemical characteristics of glycosylated recombinant human granulocyte colony stimulating factor (lenograstim) Eur J Cancer. 1994;30A (Suppl 3):S7–11. [PubMed] [Google Scholar]
  • 106.Weintraub BD, Stannard BS, Meyers L. Glycosylation of thyroid-stimulating hormone in pituitary tumor cells: influence of high mannose oligosaccharide units on subunit aggregation, combination, and intracellular degradation. Endocrinology. 1983;112(4):1331–45. doi: 10.1210/endo-112-4-1331. [DOI] [PubMed] [Google Scholar]
  • 107.Yang B, Li TD. Effect of glycosylation at Asn302 of pro-urokinase on its stability in culture supernatant. Chin Med Sci J. 2006;21(2):128–30. [PubMed] [Google Scholar]
  • 108.Baudys M, Uchio T, Mix D, Wilson D, Kim SW. Physical stabilization of insulin by glycosylation. J Pharm Sci. 1995;84(1):28–33. doi: 10.1002/jps.2600840108. [DOI] [PubMed] [Google Scholar]
  • 109.Liu H, Bulseco GG, Sun J. Effect of posttranslational modifications on the thermal stability of a recombinant monoclonal antibody. Immunol Lett. 2006;106(2):144–53. doi: 10.1016/j.imlet.2006.05.011. [DOI] [PubMed] [Google Scholar]
  • 110.Ghirlando R, Lund J, Goodall M, Jefferis R. Glycosylation of human IgG-Fc: influences on structure revealed by differential scanning micro-calorimetry. Immunol Lett. 1999;68(1):47–52. doi: 10.1016/s0165-2478(99)00029-2. [DOI] [PubMed] [Google Scholar]
  • 111.Lis H, Sharon N. Protein glycosylation. Structural and functional aspects. Eur J Biochem. 1993;218(1):1–27. doi: 10.1111/j.1432-1033.1993.tb18347.x. [DOI] [PubMed] [Google Scholar]
  • 112.Varki A. Biological roles of oligosaccharides: all of the theories are correct. Glycobiology. 1993;3(2):97–130. doi: 10.1093/glycob/3.2.97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Vegarud G, Christensen TB. The resistance of glycoproteins to proteolytic inactivation. Acta Chem Scand B. 1975;29(8):887–8. doi: 10.3891/acta.chem.scand.29b-0887. [DOI] [PubMed] [Google Scholar]
  • 114.Vegarud G, Christnsen TB. Glycosylation of Proteins: a new method of enzyme stabilization. Biotechnol Bioeng. 1975;17(9):1391–7. doi: 10.1002/bit.260170918. [DOI] [PubMed] [Google Scholar]
  • 115.Ueda T, Tomita K, Notsu Y, Ito T, Fumoto M, Takakura T, et al. Chemoenzymatic synthesis of glycosylated glucagon-like peptide 1: effect of glycosylation on proteolytic resistance and in vivo blood glucose-lowering activity. J Am Chem Soc. 2009;131(17):6237–45. doi: 10.1021/ja900261g. [DOI] [PubMed] [Google Scholar]
  • 116.Carter CR, Keeble JR, Thorpe R. Human serum inactivates non-glycosylated but not glycosylated granulocyte colony stimulating factor by a protease dependent mechanism: significance of carbohydrates on the glycosylated molecule. Biologicals. 2004;32(1):37–47. doi: 10.1016/j.biologicals.2003.12.002. [DOI] [PubMed] [Google Scholar]
  • 117.Wicker-Planquart C, Canaan S, Riviere M, Dupuis L. Site-directed removal of N-glycosylation sites in human gastric lipase. Eur J Biochem. 1999;262(3):644–51. doi: 10.1046/j.1432-1327.1999.00427.x. [DOI] [PubMed] [Google Scholar]
  • 118.Grinnell BW, Walls JD, Gerlitz B. Glycosylation of human protein C affects its secretion, processing, functional activities, and activation by thrombin. J Biol Chem. 1991;266(15):9778–85. [PubMed] [Google Scholar]
  • 119.Wang P, Zhang J, Sun Z, Chen Y, Liu JN. Glycosylation of prourokinase produced by Pichia pastoris impairs enzymatic activity but not secretion. Protein Expr Purif. 2000;20(2):179–85. doi: 10.1006/prep.2000.1310. [DOI] [PubMed] [Google Scholar]
  • 120.Sareneva T, Cantell K, Pyhala L, Pirhonen J, Julkunen I. Effect of carbohydrates on the pharmacokinetics of human interferon-gamma. J Interferon Res. 1993;13(4):267–9. doi: 10.1089/jir.1993.13.267. [DOI] [PubMed] [Google Scholar]
  • 121.Raju TS, Scallon B. Fc glycans terminated with N-acetylglucosamine residues increase antibody resistance to papain. Biotechnol Prog. 2007;23(4):964–71. doi: 10.1021/bp070118k. [DOI] [PubMed] [Google Scholar]
  • 122.Clowers BH, Dodds ED, Seipert RR, Lebrilla CB. Site determination of protein glycosylation based on digestion with immobilized nonspecific proteases and Fourier transform ion cyclotron resonance mass spectrometry. J Proteome Res. 2007;6(10):4032–40. doi: 10.1021/pr070317z. [DOI] [PubMed] [Google Scholar]
  • 123.Ashwell G, Morell A. The dual role of sialic acid in the hepatic recognition and catabolism of serum glycoproteins. Biochem Soc Symp. 1974;(40):117–24. [PubMed] [Google Scholar]
  • 124.Ashwell G, Morell AG. The role of surface carbohydrates in the hepatic recognition and transport of circulating glycoproteins. Adv Enzymol Relat Areas Mol Biol. 1974;41(0):99–128. doi: 10.1002/9780470122860.ch3. [DOI] [PubMed] [Google Scholar]
  • 125.Stockert RJ, Morell AG, Ashwell G. Structural characteristics and regulation of the asialoglycoprotein receptor. Targeted Diagn Ther. 1991;4:41–64. [PubMed] [Google Scholar]
  • 126.Pricer WE, Jr, Hudgin RL, Ashwell G, Stockert RJ, Morell AG. A membrane receptor protein for asialoglycoproteins. Methods Enzymol. 1974;34:688–91. doi: 10.1016/s0076-6879(74)34090-6. [DOI] [PubMed] [Google Scholar]
  • 127.Morell AG, Irvine RA, Sternlieb I, Scheinberg IH, Ashwell G. Physical and chemical studies on ceruloplasmin. V. Metabolic studies on sialic acid-free ceruloplasmin in vivo. J Biol Chem. 1968;243(1):155–9. [PubMed] [Google Scholar]
  • 128.Gross V, Heinrich PC, vom Berg D, Steube K, Andus T, Tran-Thi TA, et al. Involvement of various organs in the initial plasma clearance of differently glycosylated rat liver secretory proteins. Eur J Biochem. 1988;173(3):653–9. doi: 10.1111/j.1432-1033.1988.tb14048.x. [DOI] [PubMed] [Google Scholar]
  • 129.Morell AG, Gregoriadis G, Scheinberg IH, Hickman J, Ashwell G. The role of sialic acid in determining the survival of glycoproteins in the circulation. J Biol Chem. 1971;246(5):1461–7. [PubMed] [Google Scholar]
  • 130.Baynes JW, Wold F. Effect of glycosylation on the in vivo circulating half-life of ribonuclease. J Biol Chem. 1976;251(19):6016–24. [PubMed] [Google Scholar]
  • 131.Wileman T, Harding C, Stahl P. Receptor-mediated endocytosis. Biochem J. 1985;232(1):1–14. doi: 10.1042/bj2320001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Schlesinger PH, Doebber TW, Mandell BF, White R, DeSchryver C, Rodman JS, et al. Plasma clearance of glycoproteins with terminal mannose and N-acetylglucosamine by liver non-parenchymal cells. Studies with beta-glucuronidase, N-acetyl-beta-D-glucosaminidase, ribonuclease B and agalacto-orosomucoid. Biochem J. 1978;176(1):103–9. doi: 10.1042/bj1760103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Schlesinger PH, Rodman JS, Doebber TW, Stahl PD, Lee YC, Stowell CP, et al. The role of extra-hepatic tissues in the receptor-mediated plasma clearance of glycoproteins terminated by mannose or N-acetylglucosamine. Biochem J. 1980;192(2):597–606. doi: 10.1042/bj1920597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Townsend R, Stahl P. Isolation and characterization of a mannose/N-acetylglucosamine/fucose-binding protein from rat liver. Biochem J. 1981;194(1):209–14. doi: 10.1042/bj1940209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Achord DT, Brot FE, Bell CE, Sly WS. Human beta-glucuronidase: in vivo clearance and in vitro uptake by a glycoprotein recognition system on reticuloendothelial cells. Cell. 1978;15(1):269–78. doi: 10.1016/0092-8674(78)90102-2. [DOI] [PubMed] [Google Scholar]
  • 136.Weigel PH, Yik JH. Glycans as endocytosis signals: the cases of the asialoglycoprotein and hyaluronan/chondroitin sulfate receptors. Biochim Biophys Acta. 2002;1572(2–3):341–63. doi: 10.1016/s0304-4165(02)00318-5. [DOI] [PubMed] [Google Scholar]
  • 137.Jefferis R. Recombinant antibody therapeutics: the impact of glycosylation on mechanisms of action. Trends Pharmacol Sci. 2009;30(7):356–62. doi: 10.1016/j.tips.2009.04.007. [DOI] [PubMed] [Google Scholar]
  • 138.Egleton RD, Mitchell SA, Huber JD, Janders J, Stropova D, Polt R, et al. Improved bioavailability to the brain of glycosylated Met-enkephalin analogs. Brain Res. 2000;881(1):37–46. doi: 10.1016/s0006-8993(00)02794-3. [DOI] [PubMed] [Google Scholar]
  • 139.Egleton RD, Mitchell SA, Huber JD, Palian MM, Polt R, Davis TP. Improved blood-brain barrier penetration and enhanced analgesia of an opioid peptide by glycosylation. J Pharmacol Exp Ther. 2001;299(3):967–72. [PubMed] [Google Scholar]
  • 140.Kihlberg J, Ahman J, Walse B, Drakenberg T, Nilsson A, Soderberg-ahlm C, et al. Glycosylated peptide hormones: pharmacological properties and conformational studies of analogues of [1-desamino,8-D-arginine]vasopressin. J Med Chem. 1995;38(1):161–9. doi: 10.1021/jm00001a021. [DOI] [PubMed] [Google Scholar]
  • 141.Haubner R, Wester HJ, Burkhart F, Senekowitsch-Schmidtke R, Weber W, Goodman SL, et al. Glycosylated RGD-containing peptides: tracer for tumor targeting and angiogenesis imaging with improved biokinetics. J Nucl Med. 2001;42(2):326–36. [PubMed] [Google Scholar]
  • 142.Albert R, Marbach P, Bauer W, Briner U, Fricker G, Bruns C, et al. SDZ CO 611: a highly potent glycated analog of somatostatin with improved oral activity. Life Sci. 1993;53(6):517–25. doi: 10.1016/0024-3205(93)90703-6. [DOI] [PubMed] [Google Scholar]
  • 143.Halstenson CE, Macres M, Katz SA, Schnieders JR, Watanabe M, Sobota JT, et al. Comparative pharmacokinetics and pharmacodynamics of epoetin alfa and epoetin beta. Clin Pharmacol Ther. 1991;50(6):702–12. doi: 10.1038/clpt.1991.210. [DOI] [PubMed] [Google Scholar]
  • 144.Sasayama S, Moriya K, Chiba T, Matsumura T, Hayashi H, Hayashi A, et al. Glycosylated human interleukin-1alpha, neoglyco IL-1alpha, coupled with N-acetylneuraminic acid exhibits selective activities in vivo and altered tissue distribution. Glycoconj J. 2000;17(6):353–9. doi: 10.1023/A:1007181929405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Iwao Y, Hiraike M, Kragh-Hansen U, Kawai K, Suenaga A, Maruyama T, et al. Altered chain-length and glycosylation modify the pharmacokinetics of human serum albumin. Biochim Biophys Acta. 2009;1794(4):634–41. doi: 10.1016/j.bbapap.2008.11.022. [DOI] [PubMed] [Google Scholar]
  • 146.Millward TA, Heitzmann M, Bill K, Langle U, Schumacher P, Forrer K. Effect of constant and variable domain glycosylation on pharmacokinetics of therapeutic antibodies in mice. Biologicals. 2008;36(1):41–7. doi: 10.1016/j.biologicals.2007.05.003. [DOI] [PubMed] [Google Scholar]
  • 147.Sheffield WP, Marques JA, Bhakta V, Smith IJ. Modulation of clearance of recombinant serum albumin by either glycosylation or truncation. Thromb Res. 2000;99(6):613–21. doi: 10.1016/s0049-3848(00)00286-3. [DOI] [PubMed] [Google Scholar]
  • 148.Cousin P, Dechaud H, Grenot C, Lejeune H, Hammond GL, Pugeat M. Influence of glycosylation on the clearance of recombinant human sex hormone-binding globulin from rabbit blood. J Steroid Biochem Mol Biol. 1999;70(4–6):115–21. doi: 10.1016/s0960-0760(99)00101-6. [DOI] [PubMed] [Google Scholar]
  • 149.Burgon PG, Stanton PG, Robertson DM. In vivo bioactivities and clearance patterns of highly purified human luteinizing hormone isoforms. Endocrinology. 1996;137(11):4827–36. doi: 10.1210/endo.137.11.8895353. [DOI] [PubMed] [Google Scholar]
  • 150.Wawrzynczak EJ, Cumber AJ, Parnell GD, Jones PT, Winter G. Blood clearance in the rat of a recombinant mouse monoclonal antibody lacking the N-linked oligosaccharide side chains of the CH2 domains. Mol Immunol. 1992;29(2):213–20. doi: 10.1016/0161-5890(92)90102-4. [DOI] [PubMed] [Google Scholar]
  • 151.Henkin J, Dudlak D, Beebe DP, Sennello L. igh sialic acid content slows prourokinase turnover in rabbits. Thromb Res. 1991;63(2):215–25. doi: 10.1016/0049-3848(91)90285-5. [DOI] [PubMed] [Google Scholar]
  • 152.Kawatsu M, Takeo K, Kajikawa T, Funahashi I, Asahi T, Kakutani T, et al. The pharmacokinetic pattern of glycosylated human recombinant lymphotoxin (LT) in rats after intravenous administration. J Pharmacobiodyn. 1990;13(9):549–57. doi: 10.1248/bpb1978.13.549. [DOI] [PubMed] [Google Scholar]
  • 153.Gross V, Steube K, Tran-Thi T, Haussinger D, Legler G, Decker K, et al. The role of N-glycosylation for the plasma clearance of rat liver secretory glycoproteins. Eur J Biochem. 1987;162(1):83–8. doi: 10.1111/j.1432-1033.1987.tb10545.x. [DOI] [PubMed] [Google Scholar]
  • 154.Kim HJ, Lee DH, Kim DK, Han GB, Kim HJ. The glycosylation and in vivo stability of human granulocyte-macrophage colony-stimulating factor produced in rice cells. Biol Pharm Bull. 2008;31(2):290–4. doi: 10.1248/bpb.31.290. [DOI] [PubMed] [Google Scholar]
  • 155.Denis CV, Christophe OD, Oortwijn BD, Lenting PJ. Clearance of von Willebrand factor. Thromb Haemost. 2008;99(2):271–8. doi: 10.1160/TH07-10-0629. [DOI] [PubMed] [Google Scholar]
  • 156.Millar CM, Riddell AF, Brown SA, Starke R, Mackie I, Bowen DJ, et al. Survival of von Willebrand factor released following DDAVP in a type 1 von Willebrand disease cohort: influence of glycosylation, proteolysis and gene mutations. Thromb Haemost. 2008;99(5):916–24. doi: 10.1160/TH07-09-0565. [DOI] [PubMed] [Google Scholar]
  • 157.Smith WB, Dowell JA, Pratt RD. Pharmacokinetics and pharmacodynamics of epoetin delta in two studies in healthy volunteers and two studies in patients with chronic kidney disease. Clin Ther. 2007;29(7):1368–80. doi: 10.1016/j.clinthera.2007.07.014. [DOI] [PubMed] [Google Scholar]
  • 158.Jones AJ, Papac DI, Chin EH, Keck R, Baughman SA, Lin YS, et al. Selective clearance of glycoforms of a complex glycoprotein pharmaceutical caused by terminal N-acetylglucosamine is similar in humans and cynomolgus monkeys. Glycobiology. 2007;17(5):529–40. doi: 10.1093/glycob/cwm017. [DOI] [PubMed] [Google Scholar]
  • 159.Perlman S, van den Hazel B, Christiansen J, Gram-Nielsen S, Jeppesen CB, Andersen KV, et al. Glycosylation of an N-terminal extension prolongs the half-life and increases the in vivo activity of follicle stimulating hormone. J Clin Endocrinol Metab. 2003;88(7):3227–35. doi: 10.1210/jc.2002-021201. [DOI] [PubMed] [Google Scholar]
  • 160.Barrios-De-Tomasi J, Timossi C, Merchant H, Quintanar A, Avalos JM, Andersen CY, et al. Assessment of the in vitro and in vivo biological activities of the human follicle-stimulating isohormones. Mol Cell Endocrinol. 2002;186(2):189–98. doi: 10.1016/s0303-7207(01)00657-8. [DOI] [PubMed] [Google Scholar]
  • 161.Macdougall IC. Optimizing the use of erythropoietic agents-- pharmacokinetic and pharmacodynamic considerations. Nephrol Dial Transplant. 2002;17 (Suppl 5):66–70. doi: 10.1093/ndt/17.suppl_5.66. [DOI] [PubMed] [Google Scholar]
  • 162.Ni H, Blajchman MA, Ananthanarayanan VS, Smith IJ, Sheffield WP. Mutation of any site of N-linked glycosylation accelerates the in vivo clearance of recombinant rabbit antithrombin. Thromb Res. 2000;99(4):407–15. doi: 10.1016/s0049-3848(00)00263-2. [DOI] [PubMed] [Google Scholar]
  • 163.Marinaro JA, Casley DJ, Bach LA. O-glycosylation delays the clearance of human IGF-binding protein-6 from the circulation. Eur J Endocrinol. 2000;142(5):512–6. doi: 10.1530/eje.0.1420512. [DOI] [PubMed] [Google Scholar]
  • 164.Chitlaru T, Kronman C, Zeevi M, Kam M, Harel A, Ordentlich A, et al. Modulation of circulatory residence of recombinant acetylcholinesterase through biochemical or genetic manipulation of sialylation levels. Biochem J. 1998;336 (Pt 3):647–58. doi: 10.1042/bj3360647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Saxena A, Raveh L, Ashani Y, Doctor BP. Structure of glycan moieties responsible for the extended circulatory life time of fetal bovine serum acetylcholinesterase and equine serum butyrylcholinesterase. Biochemistry. 1997;36(24):7481–9. doi: 10.1021/bi963156d. [DOI] [PubMed] [Google Scholar]
  • 166.Thotakura NR, Szkudlinski MW, Weintraub BD. Structure-function studies of oligosaccharides of recombinant human thyrotrophin by sequential deglycosylation and resialylation. Glycobiology. 1994;4(4):525–33. doi: 10.1093/glycob/4.4.525. [DOI] [PubMed] [Google Scholar]
  • 167.Otter M, Kuiper J, van Berkel TJ, Rijken DC. Mechanisms of tissue-type plasminogen activator (tPA) clearance by the liver. Ann N Y Acad Sci. 1992;667:431–42. doi: 10.1111/j.1749-6632.1992.tb51645.x. [DOI] [PubMed] [Google Scholar]
  • 168.Wasley LC, Timony G, Murtha P, Stoudemire J, Dorner AJ, Caro J, et al. The importance of N- and O-linked oligosaccharides for the biosynthesis and in vitro and in vivo biologic activities of erythropoietin. Blood. 1991;77(12):2624–32. [PubMed] [Google Scholar]
  • 169.Galway AB, Hsueh AJ, Keene JL, Yamoto M, Fauser BC, Boime I. In vitro and in vivo bioactivity of recombinant human follicle-stimulating hormone and partially deglycosylated variants secreted by transfected eukaryotic cell lines. Endocrinology. 1990;127(1):93–100. doi: 10.1210/endo-127-1-93. [DOI] [PubMed] [Google Scholar]
  • 170.Lucore CL, Fry ET, Nachowiak DA, Sobel BE. Biochemical determinants of clearance of tissue-type plasminogen activator from the circulation. Circulation. 1988;77(4):906–14. doi: 10.1161/01.cir.77.4.906. [DOI] [PubMed] [Google Scholar]
  • 171.Mirshahi M, Soria J, Soria C, Bertrand O, Mirshahi M, Basdevant A. Glycosylation of human fibrinogen and fibrin in vitro. Its consequences on the properties of fibrin(ogen) Thromb Res. 1987;48(3):279–89. doi: 10.1016/0049-3848(87)90440-3. [DOI] [PubMed] [Google Scholar]
  • 172.Elliott S, Egrie J, Browne J, Lorenzini T, Busse L, Rogers N, et al. Control of rHuEPO biological activity: the role of carbohydrate. Exp Hematol. 2004;32(12):1146–55. doi: 10.1016/j.exphem.2004.08.004. [DOI] [PubMed] [Google Scholar]
  • 173.Negri L, Melchiorri P, Rocchi R, Scolaro B. Opioid receptor affinity and analgesic activity of O- and C-glycosylated opioid peptides. Acta Physiol Hung. 1996;84(4):441–3. [PubMed] [Google Scholar]
  • 174.Fishburn CS. The pharmacology of PEGylation: balancing PD with PK to generate novel therapeutics. J Pharm Sci. 2008;97(10):4167–83. doi: 10.1002/jps.21278. [DOI] [PubMed] [Google Scholar]
  • 175.Raju TS. Terminal sugars of Fc glycans influence antibody effector functions of IgGs. Curr Opin Immunol. 2008;20(4):471–8. doi: 10.1016/j.coi.2008.06.007. [DOI] [PubMed] [Google Scholar]
  • 176.Crispin M, Bowden TA, Coles CH, Harlos K, Aricescu AR, Harvey DJ, et al. Carbohydrate and domain architecture of an immature antibody glycoform exhibiting enhanced effector functions. J Mol Biol. 2009;387(5):1061–6. doi: 10.1016/j.jmb.2009.02.033. [DOI] [PubMed] [Google Scholar]
  • 177.Barbey F, Hayoz D, Widmer U, Burnier M. Efficacy of enzyme replacement therapy in Fabry disease. Curr Med Chem Cardiovasc Hematol Agents. 2004;2(4):277–86. doi: 10.2174/1568016043356192. [DOI] [PubMed] [Google Scholar]
  • 178.Beck M. Agalsidase alfa for the treatment of Fabry disease: new data on clinical efficacy and safety. Expert Opin Biol Ther. 2009;9(2):255–61. doi: 10.1517/14712590802658428. [DOI] [PubMed] [Google Scholar]
  • 179.Lee K, Jin X, Zhang K, Copertino L, Andrews L, Baker-Malcolm J, et al. A biochemical and pharmacological comparison of enzyme replacement therapies for the glycolipid storage disorder Fabry disease. Glycobiology. 2003;13(4):305–13. doi: 10.1093/glycob/cwg034. [DOI] [PubMed] [Google Scholar]
  • 180.Keating GM, Simpson D. Spotlight on agalsidase beta in Fabry disease. BioDrugs. 2007;21(4):269–71. doi: 10.2165/00063030-200721040-00007. [DOI] [PubMed] [Google Scholar]
  • 181.Keating GM, Simpson D. Agalsidase Beta: a review of its use in the management of Fabry disease. Drugs. 2007;67(3):435–55. doi: 10.2165/00003495-200767030-00007. [DOI] [PubMed] [Google Scholar]
  • 182.Jurado Garcia JM, Torres Sanchez E, Olmos Hidalgo D, Alba Conejo E. Erythropoietin pharmacology. Clin Transl Oncol. 2007;9(11):715–22. doi: 10.1007/s12094-007-0128-y. [DOI] [PubMed] [Google Scholar]
  • 183.Takeuchi M, Kobata A. Structures and functional roles of the sugar chains of human erythropoietins. Glycobiology. 1991;1(4):337–46. doi: 10.1093/glycob/1.4.337. [DOI] [PubMed] [Google Scholar]
  • 184.Weenen C, Pena JE, Pollak SV, Klein J, Lobel L, Trousdale RK, et al. Long-acting follicle-stimulating hormone analogs containing N-linked glycosylation exhibited increased bioactivity compared with o-linked analogs in female rats. J Clin Endocrinol Metab. 2004;89(10):5204–12. doi: 10.1210/jc.2004-0425. [DOI] [PubMed] [Google Scholar]
  • 185.Creus S, Chaia Z, Pellizzari EH, Cigorraga SB, Ulloa-Aguirre A, Campo S. Human FSH isoforms: carbohydrate complexity as determinant of in-vitro bioactivity. Mol Cell Endocrinol. 2001;174(1–2):41–9. doi: 10.1016/s0303-7207(00)00453-6. [DOI] [PubMed] [Google Scholar]
  • 186.le Cotonnec JY, Loumaye E, Porchet HC, Beltrami V, Munafo A. Pharmacokinetic and pharmacodynamic interactions between recombinant human luteinizing hormone and recombinant human follicle-stimulating hormone. Fertil Steril. 1998;69(2):201–9. doi: 10.1016/s0015-0282(97)00503-7. [DOI] [PubMed] [Google Scholar]
  • 187.le Cotonnec JY, Porchet HC, Beltrami V, Munafo A. Clinical pharmacology of recombinant human luteinizing hormone: Part II. Bioavailability of recombinant human luteinizing hormone assessed with an immunoassay and an in vitro bioassay. Fertil Steril. 1998;69(2):195–200. doi: 10.1016/s0015-0282(97)00502-5. [DOI] [PubMed] [Google Scholar]
  • 188.le Cotonnec JY, Porchet HC, Beltrami V, Munafo A. Clinical pharmacology of recombinant human luteinizing hormone: Part I. Pharmacokinetics after intravenous administration to healthy female volunteers and comparison with urinary human luteinizing hormone. Fertil Steril. 1998;69(2):189–94. doi: 10.1016/s0015-0282(97)00501-3. [DOI] [PubMed] [Google Scholar]
  • 189.Klett D, Bernard S, Lecompte F, Leroux H, Magallon T, Locatelli A, et al. Fast renal trapping of porcine luteinizing hormone (pLH) shown by 123I-scintigraphic imaging in rats explains its short circulatory half-life. Reprod Biol Endocrinol. 2003;1:64. doi: 10.1186/1477-7827-1-64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Dhillon S, Keating GM. Lutropin alfa. Drugs. 2008;68(11):1529–40. doi: 10.2165/00003495-200868110-00005. [DOI] [PubMed] [Google Scholar]
  • 191.Baenziger JU, Kumar S, Brodbeck RM, Smith PL, Beranek MC. Circulatory half-life but not interaction with the lutropin/chorionic gonadotropin receptor is modulated by sulfation of bovine lutropin oligosaccharides. Proc Natl Acad Sci U S A. 1992;89(1):334–8. doi: 10.1073/pnas.89.1.334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Wakefield LM, Winokur TS, Hollands RS, Christopherson K, Levinson AD, Sporn MB. Recombinant latent transforming growth factor beta 1 has a longer plasma half-life in rats than active transforming growth factor beta 1, and a different tissue distribution. J Clin Invest. 1990;86(6):1976–84. doi: 10.1172/JCI114932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Hoglund M. Glycosylated and non-glycosylated recombinant human granulocyte colony-stimulating factor (rhG-CSF)--what is the difference? Med Oncol. 1998;15(4):229–33. doi: 10.1007/BF02787205. [DOI] [PubMed] [Google Scholar]
  • 194.Hovgaard D, Mortensen BT, Schifter S, Nissen NI. Comparative pharmacokinetics of single-dose administration of mammalian and bacterially-derived recombinant human granulocyte-macrophage colony-stimulating factor. Eur J Haematol. 1993;50(1):32–6. doi: 10.1111/j.1600-0609.1993.tb00071.x. [DOI] [PubMed] [Google Scholar]
  • 195.Sylvester RK. Clinical applications of colony-stimulating factors: a historical perspective. Am J Health Syst Pharm. 2002;59(7 Suppl 2):S6–12. doi: 10.1093/ajhp/59.suppl_2.S6. [DOI] [PubMed] [Google Scholar]
  • 196.Ziltener HJ, Clark-Lewis I, Jones AT, Dy M. Carbohydrate does not modulate the in vivo effects of injected interleukin-3. Exp Hematol. 1994;22(11):1070–5. [PubMed] [Google Scholar]
  • 197.Fukushima K, Watanabe H, Takeo K, Nomura M, Asahi T, Yamashita K. N-linked sugar chain structure of recombinant human lymphotoxin produced by CHO cells: the functional role of carbohydrate as to its lectin-like character and clearance velocity. Arch Biochem Biophys. 1993;304(1):144–53. doi: 10.1006/abbi.1993.1332. [DOI] [PubMed] [Google Scholar]
  • 198.Minta JO. The role of sialic acid in the functional activity and the hepatic clearance of C1-INH. J Immunol. 1981;126(1):245–9. [PubMed] [Google Scholar]
  • 199.Longhurst H. Rhucin, a recombinant C1 inhibitor for the treatment of hereditary angioedema and cerebral ischemia. Curr Opin Investig Drugs. 2008;9(3):310–23. [PubMed] [Google Scholar]
  • 200.van Doorn MB, Burggraaf J, van Dam T, Eerenberg A, Levi M, Hack CE, et al. A phase I study of recombinant human C1 inhibitor in asymptomatic patients with hereditary angioedema. J Allergy Clin Immunol. 2005;116(4):876–83. doi: 10.1016/j.jaci.2005.05.019. [DOI] [PubMed] [Google Scholar]
  • 201.Raju TS, Scallon BJ. Glycosylation in the Fc domain of IgG increases resistance to proteolytic cleavage by papain. Biochem Biophys Res Commun. 2006;341(3):797–803. doi: 10.1016/j.bbrc.2006.01.030. [DOI] [PubMed] [Google Scholar]
  • 202.Dissing-Olesen L, Thaysen-Andersen M, Meldgaard M, Hojrup P, Finsen B. The function of the human interferon-beta 1a glycan determined in vivo. J Pharmacol Exp Ther. 2008;326(1):338–47. doi: 10.1124/jpet.108.138263. [DOI] [PubMed] [Google Scholar]
  • 203.Bocci V, Di Francesco P, Pacini A, Pessina GP, Rossi GB, Sorrentino V. Renal metabolism of homologous serum interferon. Antiviral Res. 1983;3(1):53–8. doi: 10.1016/0166-3542(83)90014-1. [DOI] [PubMed] [Google Scholar]
  • 204.Iino M, Foster DC, Kisiel W. Functional consequences of mutations in Ser-52 and Ser-60 in human blood coagulation factor VII. Arch Biochem Biophys. 1998;352(2):182–92. doi: 10.1006/abbi.1998.0595. [DOI] [PubMed] [Google Scholar]
  • 205.Lillicrap D. Extending half-life in coagulation factors: where do we stand? Thromb Res. 2008;122 (Suppl 4):S2–8. doi: 10.1016/S0049-3848(08)70027-6. [DOI] [PubMed] [Google Scholar]
  • 206.Jain S, Hreczuk-Hirst DH, McCormack B, Mital M, Epenetos A, Laing P, et al. Polysialylated insulin: synthesis, characterization and biological activity in vivo. Biochim Biophys Acta. 2003;1622(1):42–9. doi: 10.1016/s0304-4165(03)00116-8. [DOI] [PubMed] [Google Scholar]
  • 207.Gregoriadis G, Fernandes A, McCormack B, Mital M, Zhang X. Polysialic acids: potential role in therapeutic constructs. Biotechnol Genet Eng Rev. 1999;16:203–15. doi: 10.1080/02648725.1999.10647975. [DOI] [PubMed] [Google Scholar]
  • 208.Gregoriadis G, Fernandes A, Mital M, McCormack B. Polysialic acids: potential in improving the stability and pharmacokinetics of proteins and other therapeutics. Cell Mol Life Sci. 2000;57(13–14):1964–9. doi: 10.1007/PL00000676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Stork R, Zettlitz KA, Muller D, Rether M, Hanisch FG, Kontermann RE. N-glycosylation as novel strategy to improve pharmacokinetic properties of bispecific single-chain diabodies. J Biol Chem. 2008;283(12):7804–12. doi: 10.1074/jbc.M709179200. [DOI] [PubMed] [Google Scholar]
  • 210.Fernandes AI, Gregoriadis G. Polysialylated asparaginase: preparation, activity and pharmacokinetics. Biochim Biophys Acta. 1997;1341(1):26–34. doi: 10.1016/s0167-4838(97)00056-3. [DOI] [PubMed] [Google Scholar]
  • 211.Fernandes AI, Gregoriadis G. The effect of polysialylation on the immunogenicity and antigenicity of asparaginase: implication in its pharmacokinetics. Int J Pharm. 2001;217(1–2):215–24. doi: 10.1016/s0378-5173(01)00603-2. [DOI] [PubMed] [Google Scholar]
  • 212.Jelkmann W. The enigma of the metabolic fate of circulating erythropoietin (Epo) in view of the pharmacokinetics of the recombinant drugs rhEpo and NESP. Eur J Haematol. 2002;69(5–6):265–74. doi: 10.1034/j.1600-0609.2002.02813.x. [DOI] [PubMed] [Google Scholar]
  • 213.Egrie JC, Dwyer E, Browne JK, Hitz A, Lykos MA. Darbepoetin alfa has a longer circulating half-life and greater in vivo potency than recombinant human erythropoietin. Exp Hematol. 2003;31(4):290–9. doi: 10.1016/s0301-472x(03)00006-7. [DOI] [PubMed] [Google Scholar]
  • 214.Gross AW, Lodish HF. Cellular trafficking and degradation of erythropoietin and novel erythropoiesis stimulating protein (NESP) J Biol Chem. 2006;281(4):2024–32. doi: 10.1074/jbc.M510493200. [DOI] [PubMed] [Google Scholar]
  • 215.Kuter DJ, Begley CG. Recombinant human thrombopoietin: basic biology and evaluation of clinical studies. Blood. 2002;100(10):3457–69. doi: 10.1182/blood.V100.10.3457. [DOI] [PubMed] [Google Scholar]
  • 216.Ruman JI, Pollak S, Trousdale RK, Klein J, Lustbader JW. Effects of long-acting recombinant human follicle-stimulating hormone analogs containing N-linked glycosylation on murine folliculogenesis. Fertil Steril. 2005;83 (Suppl 1):1303–9. doi: 10.1016/j.fertnstert.2004.12.027. [DOI] [PubMed] [Google Scholar]
  • 217.Trousdale RK, Yu B, Pollak SV, Husami N, Vidali A, Lustbader JW. Efficacy of native and hyperglycosylated follicle-stimulating hormone analogs for promoting fertility in female mice. Fertil Steril. 2009;91(1):265–70. doi: 10.1016/j.fertnstert.2007.11.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Fauser BC, Mannaerts BM, Devroey P, Leader A, Boime I, Baird DT. Advances in recombinant DNA technology: corifollitropin alfa, a hybrid molecule with sustained follicle-stimulating activity and reduced injection frequency. Hum Reprod Update. 2009;15(3):309–21. doi: 10.1093/humupd/dmn065. [DOI] [PubMed] [Google Scholar]
  • 219.Loutradis D, Drakakis P, Vlismas A, Antsaklis A. Corifollitropin alfa, a long-acting follicle-stimulating hormone agonist for the treatment of infertility. Curr Opin Investig Drugs. 2009;10(4):372–80. [PubMed] [Google Scholar]
  • 220.Balen AH, Mulders AG, Fauser BC, Schoot BC, Renier MA, Devroey P, et al. Pharmacodynamics of a single low dose of long-acting recombinant follicle-stimulating hormone (FSH-carboxy terminal peptide, corifollitropin alfa) in women with World Health Organization group II anovulatory infertility. J Clin Endocrinol Metab. 2004;89(12):6297–304. doi: 10.1210/jc.2004-0668. [DOI] [PubMed] [Google Scholar]
  • 221.Devroey P, Fauser BC, Platteau P, Beckers NG, Dhont M, Mannaerts BM. Induction of multiple follicular development by a single dose of long-acting recombinant follicle-Stimulating hormone (FSH-CTP, corifollitropin alfa) for controlled ovarian stimulation before in vitro fertilization. J Clin Endocrinol Metab. 2004;89(5):2062–70. doi: 10.1210/jc.2003-031766. [DOI] [PubMed] [Google Scholar]
  • 222.Duijkers IJ, Klipping C, Boerrigter PJ, Machielsen CS, De Bie JJ, Voortman G. Single dose pharmacokinetics and effects on follicular growth and serum hormones of a long-acting recombinant FSH preparation (FSH-CTP) in healthy pituitary-suppressed females. Hum Reprod. 2002;17(8):1987–93. doi: 10.1093/humrep/17.8.1987. [DOI] [PubMed] [Google Scholar]
  • 223.Wraith JE. Lysosomal disorders. Semin Neonatol. 2002;7(1):75–83. doi: 10.1053/siny.2001.0088. [DOI] [PubMed] [Google Scholar]
  • 224.Pohl S, Marschner K, Storch S, Braulke T. Glycosylation- and phosphorylation-dependent intracellular transport of lysosomal hydrolases. Biol Chem. 2009 doi: 10.1515/BC.2009.076. [DOI] [PubMed] [Google Scholar]
  • 225.Kornfeld S. Lysosomal enzyme targeting. Biochem Soc Trans. 1990;18(3):367–74. doi: 10.1042/bst0180367. [DOI] [PubMed] [Google Scholar]
  • 226.Grabowski GA, Hopkin RJ. Enzyme therapy for lysosomal storage disease: principles, practice, and prospects. Annu Rev Genomics Hum Genet. 2003;4:403–36. doi: 10.1146/annurev.genom.4.070802.110415. [DOI] [PubMed] [Google Scholar]
  • 227.Murray GJ. Lectin-specific targeting of lysosomal enzymes to reticuloendothelial cells. Methods Enzymol. 1987;149:25–42. doi: 10.1016/0076-6879(87)49041-1. [DOI] [PubMed] [Google Scholar]
  • 228.Stahl PD, Rodman JS, Miller MJ, Schlesinger PH. Evidence for receptor-mediated binding of glycoproteins, glycoconjugates, and lysosomal glycosidases by alveolar macrophages. Proc Natl Acad Sci U S A. 1978;75(3):1399–403. doi: 10.1073/pnas.75.3.1399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Furbish FS, Steer CJ, Barranger JA, Jones EA, Brady RO. The uptake of native and desialylated glucocerebrosidase by rat hepatocytes and Kupffer cells. Biochem Biophys Res Commun. 1978;81(3):1047–53. doi: 10.1016/0006-291x(78)91456-0. [DOI] [PubMed] [Google Scholar]
  • 230.Furbish FS, Steer CJ, Krett NL, Barranger JA. Uptake and distribution of placental glucocerebrosidase in rat hepatic cells and effects of sequential deglycosylation. Biochim Biophys Acta. 1981;673(4):425–34. doi: 10.1016/0304-4165(81)90474-8. [DOI] [PubMed] [Google Scholar]
  • 231.Steer CJ, Furbish FS, Barranger JA, Brady RO, Jones EA. The uptake of agalacto-glucocerebrosidase by rat hepatocytes and Kupffer cells. FEBS Lett. 1978;91(2):202–5. doi: 10.1016/0014-5793(78)81172-7. [DOI] [PubMed] [Google Scholar]
  • 232.McVie-Wylie AJ, Lee KL, Qiu H, Jin X, Do H, Gotschall R, et al. Biochemical and pharmacological characterization of different recombinant acid alpha-glucosidase preparations evaluated for the treatment of Pompe disease. Mol Genet Metab. 2008;94(4):448–55. doi: 10.1016/j.ymgme.2008.04.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Hopwood JJ, Bate G, Kirkpatrick P. Galsulfase. Nat Rev Drug Discov. 2006;5(2):101–2. doi: 10.1038/nrd1962. [DOI] [PubMed] [Google Scholar]
  • 234.Turner CT, Hopwood JJ, Brooks DA. Enzyme replacement therapy in mucopolysaccharidosis I: altered distribution and targeting of alpha-L-iduronidase in immunized rats. Mol Genet Metab. 2000;69(4):277–85. doi: 10.1006/mgme.2000.2979. [DOI] [PubMed] [Google Scholar]

RESOURCES