BioMedNet HomeLibraryDatabasesCollaborationsJob ExchangeShopping MallYour Room


FULL TEXT (TEXT + FULL FIGURES)

Current Opinion in Biotechnology
Vol. 7, No. 5, October 1996
Recombinant protein production in transgenic animals
[Review article]
Yann Echelard
Current Opinion in Biotechnology 1996, 7:536-540.
 
Viewing options [Help]

text only
 
 
Associated links
 
 
-> Publications by
Yann Echelard
 
 
-> Jump to this record in Evaluated MEDLINE
 
 
-> Related records from Evaluated MEDLINE
 
 
-> Related fulltext articles on BioMedNet
 


Outline


Abstract
The engineering of animals for recombinant protein production has gone beyond the stage of identifying proper regulatory sequences. Efforts are now spent on the generation of transgenic animals that process heterologous proteins more efficiently. Another line of research is the development of strategies aimed at bypassing pronuclear microinjection.

Abbreviations
CHO—Chinese hamster ovary;
ES—embryonic stem;
HbA—hemoglobin A;
hflX—human factor IX;
hGH—human growth hormone;
hPC—human protein C;
WAP—whey acidic protein.


Introduction

The commercial potential of transgenic animal bioreactors for the production of high-value proteins is appreciable. As a result, a growing body of work has focused on outlining the parameters of mammary gland expression of heterologous proteins. This review will discuss the following three subjects: milk-specific regulatory elements, post-translational modification capabilities of the mammary gland, and physiological effects of transgene expression. Recent developments in the related field of blood-borne production of recombinant proteins and peptides will also be summarized.

Pronuclear microinjection, which is still the only method routinely used to introduce foreign DNA in the germline of farm animals, is, at best, a costly, tedious, and inefficient endeavor. The last part of this review will describe some of the efforts aimed at bypassing pronuclear microinjection, either by developing pluripotent embryonic cell lines, or by developing somatic approaches similar to those currently used in gene therapy trials. These technologies could eventually improve our ability to fine tune the production characteristics of animal bioreactors.



Expression of foreign proteins in the mammary gland

The combination of large daily protein output, excellent post-translational processing abilities, ease of access to the recombinant protein (milking), and low capital cost of production plants (farms) as compared to industrial high-volume fermenters, has made the transgenic mammary gland an excellent candidate for the production of recombinant protein. Regulatory sequences of the major milk protein genes have been used, with variable success, to direct the expression of an array of heterologous proteins in the milk of transgenic mice (mostly), rats, rabbits, pigs, sheep, goats, and cows (for a recent review, see [1•]). Five promoters have been shown to yield high levels (>1mg ml-1) of foreign protein in milk: sheep beta-lactoglobulin [2] [3] [4], goat beta-casein ([5] [6] [7•]; H Meade et al., abstract 110, Meeting on Mouse Molecular Genetics, Cold Spring Harbor, New York, September 1994), bovine alphas1-casein [8] [9], rabbit whey acidic protein (WAP) [10] [11] [12], and mouse WAP [13] [14••]. Maximum expression levels reported with the mouse WAP promoter have been lower than with other promoters [5] [13] [14••] [15] [16] [17] [18] [19]. However, this could be a function of the heterologous gene expressed. When the 4.2 kilobase (kb) mouse WAP promoter [18] was used to express a human alpha1-antitrypsin genomic construct in milk of transgenic mice, levels comparable to that reported with other promoters [2] [10] [19] were observed (B Wilburn, J Williams, Y Echelard, unpublished data).

In most accounts, it has been difficult to achieve protein yields superior to 1mg ml-1 with cDNA derived constructs [3] [15] [16] [17] [19] [20] [21] [22] and sometimes the transgenes are transcriptionally silent. However, the use of a genomic construct does not always ensure high expression levels [23]. The weak performance of cDNA-containing transgenes could be caused by sensitivity to the silencing influences of chromosomal sequences surrounding the integration sites. Two groups [24] [25] have attempted to correct this situation by co-microinjecting highly expressed genomic constructs with poorly expressed cDNA constructs in order to obtain integration sites containing both silent and highly expressed transgenes. In both cases, although there was increased expression of the cDNA transgenes from co-integrated sites as compared to transgenic integration sites containing only cDNA sequences, the cDNA derived expression levels were still 2–3 orders of magnitude lower than those obtained from genomic transgenes. In another study, poorly expressed sheep beta-lactoglobulin driven human factor IX (hfIX) cDNA transgenes were shown to contain a cryptic 3' splice site [26••], leading to the accumulation of aberrantly spliced hfIX transcripts. With a construct eliminating the cryptic splice site, transgenic mice expressed hfIX to levels up to 60µg ml-1. This is much higher than with unmodified hfIX transgenes and close to commercially acceptable levels if the modified transgene is expressed similarly in large animals.

Proper post-translational modification of milk-produced recombinant proteins is also an issue of concern. A recent study [27••] compares the two N-glycosylation sites of human interferon-gamma (IFN-gamma; Asn25 and Asn97) in recombinant protein samples obtained with three expression systems: Chinese hamster ovary (CHO) cells, baculovirus-infected SF9 cells, and the mammary gland of transgenic mice. The transgenic mouse-derived IFN-gamma had predominantly complex sialylated biantennary N-glycans at Asn25, similar to the CHO cell-derived IFN-gamma. An increased proportion of oligomannose at Asn97 was found in transgenic mouse-derived material as compared to the CHO cell IFN-gamma. Increased incidence of oligomannose glycans could affect the clearance by the mannose receptor of the recombinant protein.

Transgenically produced human protein C (hPC) from both mouse and swine has also been well characterized [14••] [17] [28] [29]. Mature hPC is an extensively glycosylated heterodimer that requires the removal of prepropeptide and propeptide, beta-hydroxylation, as well as gamma-carboxylation of nine glutamic acid residues, for full activity. Analysis of transgenic mice and pig mammary gland-derived hPC, as compared to plasma-derived hPC, showed reduced anticoagulant activity as well as high proportions of single chain unprocessed precursors [17] [28] [29], indicating that the mammary gland carried post-translational modifications rather inefficiently. It is possible that the high-level expression of hPC was saturating the post-translational modification machinery of the mammary gland, because mammary gland expression of lower levels of another extensively modified protein, hfIX, does not seem to give rise to a high proportion of incompletely processed protein [26••]. In an effort to ameliorate the processing of recombinant hPC, bi-transgenic mice coexpressing hPC and the human serine protease furin in their mammary gland were generated [13]. These mice secreted low amounts (<5%) of unprocessed single chain hPC in contrast to the 40–60% encountered in animals solely expressing hPC [13] [17]. Coexpression of a modification enzyme with the protein of interest could also be applied to other post-translational modifications (N-thinspace and O-glycosylation, gamma-carboxylation). Moreover, it seems that the high efficiency of processing was obtained by using what should be a relatively inefficiently expressed human furin cDNA transgene, making this strategy even more attractive.

In another example [30••], the modification of the mouse mammary gland post-translational modification capability by the transgenic expression of a human alpha1,2-fucosyltransferase was exploited. Here, the objective was not to better process a recombinant protein, but to remodel the structure of milk glycoconjugates. This experiment showed that transgenic dairy animals could become an inexpensive source of otherwise hard to synthesize oligosaccharide of industrial or therapeutic value, as well as a source of recombinant pharmaceutical proteins.

The effect of the expression of the transgene on the production animal is another concern. In problem cases, the expression of the heterologous protein can either influence the physiology of the mammary gland and disrupt lactation [6] [31] [32] [33] [34] [35], or have a systemic effect on the transgenic animal, affecting its viability and reproductive performance [11] [35] [36] [37]. A number of milk proteins have been shown to be usually present in blood during lactation [38•] [39] [40]; hence the detection, in the serum, of a heterologous protein expressed in the mammary gland was not surprising when it was encountered [10] [11] [12] [19] [41] [42]. RNA analysis also showed the incidence of low-level ectopic expression of WAP-derived [10] [11] [12] [18] [23] [38•] [41] [42] [43] and beta-lactoglobulin-derived [44] [45] [46] [47] transgenes, whereas the expression of alphas1-casein and beta-casein transgenes (when examined) was restricted to the mammary gland [7•] [8] [9] [22] [48] [49] [50]. With the tighter regulation of alphas1-casein and beta-casein promoters, it may be argued that these sequences could be more appropriate than the current versions of the WAP and beta-lactoglobulin cassettes for driving the milk production of very active proteins such as human growth hormone (hGH). However, in terms of the health of the transgenic production animal, low-level ectopic expression of proteins, such as alpha1-antitrypsin, human serum albumin, or most monoclonal antibodies, is probably inconsequential.



Transgenic production of blood-borne peptides

The production of recombinant human hemoglobin A (HbA) has been achieved in transgenic mice and pigs [51] to levels up to 24% completely human HbA and 30% human alpha–pig beta hybrid using a construct containing human alpha-, epsi-, beta-globin genes linked to a human beta locus control region (LCR). The key to achieving a high expression level was the construction of a fusion beta-globin gene containing the pig beta-globin promoter linked to the human beta-globin coding region. Previous HbA constructs containing both the human promoter and beta-globin coding sequences permitted the production of, at most, 10% of human HbA in transgenic pigs [52]. Animals appeared to tolerate well the high levels of human HbA in their erythrocytes. Structural analysis of the swine-derived human recombinant HbA has shown that it was equivalent to protein purified from human blood [53].

Another procedure exploiting the recombinant hemoglobin expression system, and the fact that terminally extended alpha-globin can be functional, could become an alternative to the chemical synthesis of therapeutic peptides. Carboxy-terminal fusions of human alpha-globin with either alpha-endorphin (16-mer) or magainin (26-mer) were expressed in transgenic mice [54]. Fusion proteins were expressed to levels corresponding to 25% of the total hemoglobin, seemingly without ill effects on the transgenic mice. Magaining peptides were cleanly released by enterokinase digestion of purified fusion protein, using a cleavage site engineered between the two parts of the chimeric polypeptide. On the other hand, a cryptic enterokinase site caused partial degradation of alpha-endorphin during the digestion step, showing that the release of the recombinant peptides from the chimeric protein will not always be straightforward.



Alternatives to pronuclear microinjection

Approaches combining the potential of clonal expansion offered by nuclear transfer technology, with the possibilities of gene addition and replacement achievable in embryonic stem (ES) cell lines (or ES-like cells), could be particularly beneficial to the generation of transgenic bioreactors (potential advantages are reviewed in [55]).

Pluripotent ES cells able to contribute to the germline have only been derived in mice. Recently, there have been descriptions of chimeric animals generated with rat [56] and pig [57] ES cells, but, in both cases, no evidence of germline transmission from these cells has been reported. Cow fetuses obtained following the nuclear transfer of ES cell-derived nuclei in recipient oocytes died in utero, exhibiting major defects in placental development [58•]. In a related breakthrough, two healthy phenotypically female lambs were born from embryos generated by transferring nuclei isolated from embryo-derived epithelial cells into enucleated oocytes [59••]. Microsatellites analysis demonstrated that all the nuclear transfer lambs and fetuses were derived from the cell line. It is not yet clear whether foreign DNA can be easily introduced in this type of cell line, and there are questions about the health and reproductive fitness of the recovered lambs. However, this experiment is the first demonstration that it is possible to obtain large farm animals from embryo derived-cell lines, and it is certainly a step towards the possibility of performing gene replacement in large animals.

Another group reported a strategy that bypasses entirely the transgenic route to obtain recombinant protein production in animals [60]. Replication-defective retroviral vectors carrying the hGH cDNA were used to infect the mammary gland of goats hormonally induced to lactate. Expression levels of up to 118ng of hGH per milliliter of milk were obtained on the first lactation day. By day 3, expression had stabilized to 3–12 ng ml-1. The advantage of this method is that expression of the recombinant protein is obtained quickly, without the delay caused by the generation interval required to generate a producing transgenic animal (18 months for goats). However, production levels reported in this paper [60] are very low. In the conditions described, less than 1 mg of crude recombinant hGH would be obtained per lactation per goat. If this technology could also be applied to other potential production species (rabbit, sheep, pigs, and cows), it could permit the quick determination, for specific proteins, of which system is most appropriate for the high-level production of correctly processed polypeptides.



Conclusions

The expression of recombinant proteins in transgenic animals has come of age. At least two products (sheep-derived human alpha1-antitrypsin and goat-derived antithrombin III) will be submitted to phase I trials within two years. Transgenic animals have the ability to process highly modified proteins, although not always with total efficiency. There are species-specific and tissue-specific characteristics. However, it seems that transgenic systems are flexible and that, when needed, processing enzymes can be coexpressed with foreign products to obtain more humanized recombinant proteins. In the future, the further characterization of each bioreactor's capability, and an improved ability to introduce specific genetic modifications, should lead to more sophisticated transgenic production herds.



Acknowledgements

I wish to thank Barbara Davis for her help with manuscript preparation, and Harry Meade, Carol Ziomek, and Li-How Chen for comments and many stimulating discussions.



References and recommended reading
Papers of particular interest, published within the annual period of review, have been highlighted as:
• of special interest
•• of outstanding interest
  1. • Maga EA, Murray JD:
    Mammary gland expression of transgenes and the potential for altering the properties of milk.
    Bio-Technology 1995, 13: 1452–1457.
    This review provides a thorough summary of the mammary gland-specific expression of heterologous proteins.
    Return to citation reference [1]

  2. Wright G, Carver A, Cottom D, Reeves D, Scott A, Simons P, Wilmut I, Garner I, Colman A:
    High level expression of active human alpha1 antitrypsin in the milk of transgenic sheep.
    Bio-Technology 1991, 9: 830–834.
    Return to citation reference [1] [2]

  3. Shani M, Barash I, Nathan M, Ricca G, Searfoss GH, Dekel I, Faerman A, Givol D, Hurwitz DR:
    Expression of human serum albumin in the milk of transgenic mice.
    Transgenic Res 1992, 1: 195–208. [MEDLINE] [Cited by]
    Return to citation reference [1] [2]

  4. Hurwitz DR, Nathan M, Barash I, Ilan N, Shani M:
    Specific combinations of human serum albumin introns direct high level expression of albumin in transfected COS cells and in the milk of transgenic mice.
    Transgenic Res 1994, 3: 365–375. [MEDLINE] [Cited by]
    Return to citation reference [1]

  5. Ebert KM, Selgrath JP, DiTullio P, Denman J, Smith TE, Menon MA, Schindler JE, Monastersky GM, Vitale JA, Gordon K:
    Transgenic production of a variant of human tissue plasminogen activator in goat milk: generation of transgenic goats and analysis of expression.
    Bio-Technology 1991, 9: 835–838.
    Return to citation reference [1] [2]

  6. Ebert KM, Schindler JES:
    Transgenic farm animals: progress report.
    Theriogenology 1993, 39: 121–135. [Cited by]
    Return to citation reference [1] [2]

  7. • Persuy MA, Legrain S, Printz C, Stinnakre MG, Lepourry L, Brignon G, Mercier JC:
    High-level, stage-thinspace and mammary-tissue specific expression of a caprine kappa-casein-encoding minigene driven by a beta-casein promoter in transgenic mice.
    Gene 1995, 165: 291–296. [MEDLINE] [Cited by]
    First step toward the transgenic amelioration of milk through manipulation of the kappa-casein content.
    Return to citation reference [1] [2]

  8. Meade H, Gates L, Lacy E, Lonberg N:
    Bovine alphaS1-casein gene sequences direct expression of active human urokinase in mouse milk.
    Bio-Technology 1990, 8: 443–446.
    Return to citation reference [1] [2]

  9. Brem G, Hartl P, Besenfelder U, Wolf E, Zinovieva N, Pfaller R:
    Expression of synthetic cDNA sequences encoding human insulin-like growth factor-1 (IGF-1) in the mammary gland of transgenic rabbits.
    Gene 1994, 149: 351–355. [MEDLINE] [Cited by]
    Return to citation reference [1] [2]

  10. Bischoff R, Degryse E, Perraud F, Dalemans W, Ali-Hadji D, Thépot D, Devinoy E, Houdebine LM, Pavirani A:
    A 17.6 kbp region located upstream of the rabbit WAP gene directs high level expression of a functional human protein variant in transgenic mouse milk.
    FEBS Lett 1992, 305: 265–268. [MEDLINE] [Cited by]
    Return to citation reference [1] [2] [3] [4]

  11. Devinoy E, Thépot D, Stinnakre MG, Fontaine ML, Grabowski H, Puissant C, Pavirani A, Houdebine LM:
    High level production of human growth hormone in the milk of transgenic mice —thinspace the upstream region of the rabbit whey acidic protein (Wap) gene targets transgene expression to the mammary gland.
    Transgenic Res 1994, 3: 79–89. [MEDLINE] [Cited by]
    Return to citation reference [1] [2] [3] [4]

  12. Thépot D, Devinoy E, Fontaine ML, Stinnakre MG, Massoud M, Kann G, Houdebine LM:
    Rabbit whey acidic protein gene upstream region controls high-level expression of bovine growth hormone in the mammary gland of transgenic mice.
    Mol Reprod Dev 1995, 42: 261–267. [MEDLINE] [Cited by]
    Return to citation reference [1] [2] [3]

  13. Velander WH, Johnson JL, Page RL, Russell CG, Subramanian A, Wilkins TD, Gwazdauskas FC, Pittius C, Drohan WN:
    High level expression of a heterologous protein in the milk of transgenic swine using the cDNA encoding human protein C.
    Proc Natl Acad Sci USA 1992, 89: 12003–12007. [MEDLINE] [Cited by]
    Return to citation reference [1] [2] [3] [4]

  14. •• Drews R, Paleyanda RK, Lee TK, Chang RR, Rehemtulla A, Kaufman RJ, Drohan WN, Lubon HL:
    Proteolytic maturation of protein C upon engineering the mouse mammary gland to express furin.
    Proc Natl Acad Sci USA 1995, 92: 10462–10466. [MEDLINE] [Cited by]
    Coexpression of the serine protease furin with human protein C (hPC) in the mammary gland leads to higher conversion rates of the hPC precursor to mature protein. Example of enhancement of the protein processing capacity of the mammary gland, seemingly without deleterious effects on the transgenic mice.
    Return to citation reference [1] [2] [3]

  15. Gordon K, Lee E, Vitale JA, Smith AE, Westphal H, Hennighausen L:
    Production of human tissue plasminogen activator in transgenic mouse milk.
    Bio-Technology 1987, 5: 1183–1187.
    Return to citation reference [1] [2]

  16. Yu SH, Deen KC, Lee E, Hennighausen L, Sweet RW, Rosenberg M, Westphal H:
    Functional human CD4 protein produced in milk of transgenic mice.
    Mol Biol Med 1989, 6: 255–261. [MEDLINE] [Cited by]
    Return to citation reference [1] [2]

  17. Drohan WN, Zhang D, Paleyanda RK, Chang R, Wroble M, Velander WH, Lubon H:
    Inefficient processing of human protein C in the mouse mammary gland.
    Transgenic Res 1994, 3: 355–364. [MEDLINE] [Cited by]
    Return to citation reference [1] [2] [3] [4] [5]

  18. Hansson L, Edlund M, Edlund A, Johansson T, Marklund SL, Fromm S, Stromqvist M, Tornell J:
    Expression and characterization of biologically active human extracellular superoxide dismutase in milk of transgenic mice.
    J Biol Chem 1994, 269: 5358–5363. [MEDLINE] [Cited by]
    Return to citation reference [1] [2] [3]

  19. Rokkones E, Fromm SH, Kareem BN, Kungland H, Olstad OK, Høgset A, Iversen J, Bjøro T, Gautvik KM:
    Human parathyroid hormone as a secretory peptide in milk of transgenic mice.
    J Cell Biochem 1995, 59: 168–176. [MEDLINE] [Cited by]
    Return to citation reference [1] [2] [3] [4]

  20. Paleyanda RK, Zhang DW, Hennighausen L, McKnight RA, Lubon H:
    Regulation of human protein C gene expression by the mouse WAP promoter.
    Transgenic Res 1994, 3: 335–343. [MEDLINE] [Cited by]
    Return to citation reference [1]

  21. Whitelaw CBA, Archibald AL, Harris S, McClenaghan M, Simons JP, Clark AJ:
    Targeting expression to the mammary gland: intronic sequences can enhance the efficiency of gene expression in transgenic mice.
    Transgenic Res 1991, 1: 3–13. [MEDLINE] [Cited by]
    Return to citation reference [1]

  22. Platenburg GJ, Kootwijk EPA, Kooiman PM, Woloshuk SL, Nuijens JH, Krimpenfort PJA, Pieper FR, De Boer HA, Strijker R:
    Expression of human lactoferrin in milk of transgenic mice.
    Transgenic Res 1994, 3: 99–108. [MEDLINE] [Cited by]
    Return to citation reference [1] [2]

  23. Wei Y, Yarus S, Greenberg NM, Whitsett J, Rosen JM:
    Production of human surfactant protein C in the milk of transgenic mice.
    Transgenic Res 1995, 4: 232–240. [MEDLINE] [Cited by]
    Return to citation reference [1] [2]

  24. Clark AJ, Cowper A, Wallace R, Wright G, Simons JP:
    Rescuing transgene expression by co-integration.
    Bio-Technology 1992, 10: 1450–1454.
    Return to citation reference [1]

  25. McKnight RA, Wall RJ, Hennighausen L:
    Expression of genomic and cDNA after co-integration in transgenic mice.
    Transgenic Res 1995, 4: 39–45. [MEDLINE] [Cited by]
    Return to citation reference [1]

  26. •• Yull F, Wallace HR, Cowper A, Percy J, Cottingham I, Clark AJ:
    Fixing human factor IX (fIX): correction of a cryptic RNA splice enables the production of biologically active fIX in the mammary gland of transgenic mice.
    Proc Natl Acad Sci USA 1995, 82: 10899–10903. [Cited by]
    This work provides an alternative explanation for the low expression levels often encountered with cDNA transgenes.
    Return to citation reference [1] [2]

  27. •• James DC, Freedman RB, Hoare M, Ogonah OW, Roonez BC, Larionov OA, Dobrovolsky V, Lagutin OV, Jenkins N:
    N-glycosylation of recombinant human interferon-gamma produced in different animal expression systems.
    Bio-Technology 1995, 13: 592–596.
    This article underlines the organism-thinspace and tissue-specific nature of glycosylation and the need to thoroughly analyze the carbohydrate composition of recombinant proteins.
    Return to citation reference [1]

  28. Lee TK, Drohan WN, Lubon H:
    Proteolytic processing of human protein C in swine mammary gland.
    J Biochem 1995, 118: 81–87. [MEDLINE] [Cited by]
    Return to citation reference [1] [2]

  29. Medved LV, Orthner CL, Lubon H, Lee TK, Drohan WN, Ingham KC:
    Thermal stability and domain–domain interactions in natural and recombinant protein C.
    J Biol Chem 1995, 270: 13652–13659. [Full text] [MEDLINE] [Cited by]
    Return to citation reference [1] [2]

  30. •• Prieto PA, Mukerji P, Kelder B, Erney R, Gonzalez D, Yun JS, Smith DF, Moremen KW, Nardelli C, Pierce M et al.:
    Remodeling of mouse glycoconjugates by transgenic expression of human glycosyltransferase.
    J Biol Chem 1995, 270: 29515–29519. [Full text] [MEDLINE] [Cited by]
    This article introduces a stimulating twist to the transgenic milk concept. The mammary gland becomes an efficient neooligosaccharide-producing bioreactor.
    Return to citation reference [1]

  31. Jhappan C, Geiser AG, Kordon EC, Bagheri D, Hennighausen L, Roberts AB, Smith GH, Merlino G:
    Targeting expression of a transforming growth factor beta 1 transgene to the pregnant mammary gland inhibits alveolar development and lactation.
    EMBO J 1993, 12: 1835–1845. [MEDLINE] [Cited by]
    Return to citation reference [1]

  32. Pierce DF Jr, Johnson MD, Matsui Y, Robinson SD, Gold LI, Purchio AF, Daniels CW, Hogan BLM, Moses HL:
    Inhibition of mammary duct development but not alveolar outgrowth during pregnancy in transgenic mice expressing active TGF-beta1.
    Genes Dev 1993, 7: 2308–2317. [MEDLINE] [Cited by]
    Return to citation reference [1]

  33. Li B, Greenberg N, Stephens LC, Meyn R, Medina D, Rosen JM:
    Preferential overexpression of a 172ArgrarrLeu mutant p53 in the mammary gland of transgenic mice results in altered lobulalveolar development.
    Cell Growth Differ 1994, 5: 711–721. [MEDLINE] [Cited by]
    Return to citation reference [1]

  34. Kordon EC, McKnight RA, Jhappan C, Hennighausen L, Merlino G, Smith GH:
    Ectopic TGF-beta1 expression in the secretory mammary epithelium induces early senescence of the epithelial stem cell population.
    Dev Biol 1995, 168: 47–61. [MEDLINE] [Cited by]
    Return to citation reference [1]

  35. Bleck GT, Jimenez-Flores R, Bremel RD:
    Abnormal properties of milk from transgenic mice expression bovine beta-casein under the control of the bovine alpha-lactalbumin 5' flanking region.
    Int Dairy J 1995, 5: 619–632. [Cited by]
    Return to citation reference [1] [2]

  36. Reddy VB, Vitale JA, Wei C, Montoya-Zavala M, Stice SL, Balise J, Robl JM:
    Expression of human growth hormone in the milk of transgenic mice.
    Animal Biotechnol 1991, 2: 15–29.
    Return to citation reference [1]

  37. Jänne J, Hyttinen JM, Peura T, Tolvanen M, Alhonen L, Sinervirta R, Halmekytö M:
    Transgenic bioreactors.
    Int J Biochem 1994, 26: 859–870. [MEDLINE] [Cited by]
    Return to citation reference [1]

  38. • Wall RJ, Rexroad CE, Powell A, Shamay A, McKnight R, Hennighausen L:
    Synthesis and secretion of the mouse whey acidic protein in transgenic sheep.
    Transgenic Res 1996, 5: 67–72. [MEDLINE] [Cited by]
    Documents potential morbidity associated with the ectopic expression of the WAP protein in transgenic sheep.
    Return to citation reference [1] [2]

  39. Akers RM, McFadden TB, Beal WE, Guidry AJ, Farrell HM:
    Radioimmunoassay for measurement of bovine alpha lactalbumin in serum, milk and tissue culture media.
    J Dairy Res 1986, 53: 419–429. [MEDLINE] [Cited by]
    Return to citation reference [1]

  40. Grabowski H, Le Bars D, Chene N, Attal J, Maliénou-N'Gassa R, Puissant C, Houdebine LM:
    Rabbit whey acidic protein concentration in milk, serum, mammary gland extract and culture medium.
    J Dairy Sci 1991, 74: 4143–4154. [MEDLINE] [Cited by]
    Return to citation reference [1]

  41. Devinoy E, Stinnakre MG, Lavialle F, Thépot D, Ollivier-Bousquet M:
    Intracellular routing and release of caseins and growth hormone produced into the milk from transgenic mice.
    Exp Cell Res 1995, 221: 272–280. [MEDLINE] [Cited by]
    Return to citation reference [1] [2]

  42. Limonta JM, Castro FO, Martinez R, Puentes P, Ramos B, Aguillar A, Lleonart RL, De la Fuente J:
    Transgenic rabbits as bioreactors for the production of human growth hormone.
    J Biotechnol 1995, 40: 49–58. [MEDLINE] [Cited by]
    Return to citation reference [1] [2]

  43. Wen J, Kanawata Y, Hideaki T, Tanaka S, Tachi C:
    Expression of whey acidic protein (WAP) genes in tissue other than the mammary gland in normal and transgenic mice expressing mWAP/hGH fusion genes.
    Mol Reprod Dev 1995, 40: 386–390. [MEDLINE] [Cited by]
    Return to citation reference [1]

  44. Simons JP, McClenaghan M, Clark AJ:
    Alteration of the quality of milk by expression of sheep beta-lactoglobulin in transgenic mice.
    Nature 1987, 328: 530–532. [MEDLINE] [Cited by]
    Return to citation reference [1]

  45. Whitelaw CBA, Harris S, McClenaghan M, Simons JP, Clark AJ:
    Position-independent expression of the ovine beta-lactoglobulin gene in transgenic mice.
    Biochem J 1992, 286: 31–39. [Cited by]
    Return to citation reference [1]

  46. Barash I, Faerman A, Ratovitsky T, Puzis R, Nathan M, Hurwitz DR, Shani M:
    Ectopic expression of beta-lactoglobulin/human serum albumin fusion genes in transgenic mice: hormonal regulation and in situ localization.
    Transgenic Res 1994, 3: 141–151. [MEDLINE] [Cited by]
    Return to citation reference [1]

  47. Farini E, Whitelaw BA:
    Ectopic expression of beta-lactoglobulin transgenes.
    Mol Gen Genet 1995, 246: 734–738. [MEDLINE] [Cited by]
    Return to citation reference [1]

  48. Maga EA, Anderson GB, Huang MC, Murray JD:
    Expression of human lysozyme mRNA in the mammary gland of transgenic mice.
    Transgenic Res 1994, 3: 36–42. [MEDLINE] [Cited by]
    Return to citation reference [1]

  49. Clarke RE, Sokol D, Rigby N, Ward K, Murray JD, Mackinlay AG:
    Mammary gland-specific expression of bovine alphas1-casein-derived transgenes in mice.
    Transgenics 1994, 1: 313–319. [Cited by]
    Return to citation reference [1]

  50. Rjinkels M, Kooiman PM, Krimpenfort PJA, De Boer HA, Pieper FR:
    Expression analysis of the individual bovine beta-, alphas2-, and kappa-casein genes in transgenic mice.
    Biochem J 1995, 311: 929–937. [Cited by]
    Return to citation reference [1]

  51. Sharma A, Martin MJ, Okabe JF, Truglio RA, Dhanjal NK, Logan JS, Kumar R:
    An isologous porcine promoter permits high level expression of human hemoglobin in transgenic swine.
    Bio-Technology 1994, 12: 55–59.
    Return to citation reference [1]

  52. Swanson ME, Martin MJ, O'Donnell KJ, Hoover K, Lago W, Huntress V, Paron CT, Pinkert CK, Pilder S, Logan JS:
    Production of functional hemoglobin in transgenic swine.
    Bio-Technology 1992, 10: 557–559.
    Return to citation reference [1]

  53. Rao MJ, Schneider K, Chait BT, Chao TI, Keller H, Anderson S, Manjula BN, Kumar R, Acharya AS:
    Recombinant hemoglobin A produced in transgenic swine: structural equivalence with human hemoglobin A.
    Artif Cells Blood Substit and Immobil Biotechnol 1994, 22: 695–700.
    Return to citation reference [1]

  54. Sharma A, Khoury-Christianson AM, White SP, Dhanjal NP, Huang W, Paulhiac C, Friedman EJ, Mangula BN, Kumar R:
    High-efficiency synthesis of human alpha-endorphin and magainin in the erythrocytes of transgenic mice: a production system for therapeutic peptides.
    Proc Natl Acad Sci USA 1994, 91: 9337–9341. [MEDLINE] [Cited by]
    Return to citation reference [1]

  55. Clark AJ, Bissinger P, Bullock DW, Damak S, Wallace R, Whitelaw CBA, Yull F:
    Chromosomal position effects and the modulation of transgene expression.
    Reprod Fertil Dev 1994, 6: 589–598. [MEDLINE] [Cited by]
    Return to citation reference [1]

  56. Iannaccone PM, Taborn GU, Garton RL, Caplice MD, Brenin DR:
    Pluripotent embryonic stem cells from the rat are capable of producing chimeras.
    Dev Biol 1994, 163: 288–292. [MEDLINE] [Cited by]
    Return to citation reference [1]

  57. Wheeler MB:
    Development and validation of swine embryonic stem cells: a review.
    Reprod Fertil Dev 1994, 6: 563–568. [MEDLINE] [Cited by]
    Return to citation reference [1]

  58. • Stice SL, Strelchenko NS, Keefer CL, Matthews L:
    Pluripotent bovine embryonic cell lines direct embryonic development following nuclear transfer.
    Biol Reprod 1996, 54: 100–110. [MEDLINE] [Cited by]
    The most advanced report on cattle-derived ES cells.
    Return to citation reference [1]

  59. •• Campbell KHS, McWhir J, Ritchie WA, Wilmut I:
    Sheep cloned by nuclear transfer from a cultured cell line.
    Nature 1996, 380: 64–66. [MEDLINE] [Cited by]
    The authors describe an important breakthrough in the frustrating search for pluripotent embryo-derived cell lines in species other than the mouse.
    Return to citation reference [1]

  60. Archer JS, Kennan WA, Gould MN, Bremel RD:
    Human growth hormone (hGH) secretion in milk of goats after direct transfer of the hGH gene into the mammary gland by using replication defective retrovirus vectors.
    Proc Natl Acad Sci USA 1994, 91: 6840–6844. [MEDLINE] [Cited by]
    Return to citation reference [1] [2]



Author Contacts
Y Echelard, Genzyme Transgenics Corporation, One Mountain Road, Framingham, MA 01701-9322, USA; e-mail: ypgrech@world.std.com.
Return to author list


Copyright

Copyright © 1996 Current-Opinion.com
Help   Feedback      Search   Map
Black Line
© 1998-1999 BioMedNet. All rights reserved.
email: info@biomednet.com.